Next Article in Journal
Feasibility and Impact of Adapted Physical Activity (APA) in Cancer Outpatients Beginning Medical Anti-Tumoral Treatment: The UMA-CHAPA Study
Next Article in Special Issue
Handheld PET Probe for Pediatric Cancer Surgery
Previous Article in Journal
Treatment of Neuropathic Pain Directly Due to Cancer: An Update
Previous Article in Special Issue
Antiangiogenic Compound Axitinib Demonstrates Low Toxicity and Antitumoral Effects against Medulloblastoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Novel Targeted Therapeutic Strategies for Ewing Sarcoma

1
Group of Experimental Biotherapy and Diagnostic, Department of Advanced Materials, Institute for Regenerative Medicine, Sechenov First Moscow State Medical University, Moscow 119991, Russia
2
World-Class Research Center “Digital Biodesign and Personalized Healthcare”, Sechenov First Moscow State Medical University, Moscow 119991, Russia
3
National Cancer Institute, National Institutes of Health, Bethesda, MD 20814, USA
4
Tumor Immunology Lab, Department of Zoology, Institute of Science, Banaras Hindu University, Varanasi 221005, India
5
Endocrine Research Lab, Department of Zoology, Kutir Post Graduate College, Chakkey, Jaunpur 222146, India
6
Department of Advanced Materials, Institute for Regenerative Medicine, Sechenov First Moscow State Medical University, Moscow 119991, Russia
*
Author to whom correspondence should be addressed.
Cancers 2022, 14(8), 1988; https://doi.org/10.3390/cancers14081988
Submission received: 2 March 2022 / Revised: 3 April 2022 / Accepted: 11 April 2022 / Published: 14 April 2022
(This article belongs to the Special Issue Targeted Therapies for Pediatric Solid Tumors)

Abstract

:

Simple Summary

Ewing sarcoma is an uncommon cancer that arises in mesenchymal tissues and represents the second most widespread malignant bone neoplasm after osteosarcoma in children. Therapy has increased the 5-year survival rate in the last 40 years, although the recurrence rate has remained high. There is an immediate and unmet need for the development of novel Ewing sarcoma therapies. We offer new prospective targets for the therapy of Ewing sarcoma. The EWSR1/FLI1 fusion protein, which is identified in 85–90% of Ewing sarcoma tumors, and its direct targets are given special focus in this study. Experimantal therapy that targets multiple signaling pathways activated during ES progression, alone or in combination with existing regimens, may become the new standard of care for Ewing sarcoma patients, improving patient survival.

Abstract

Ewing sarcoma (ES) is an uncommon cancer that arises in mesenchymal tissues and represents the second most widespread malignant bone neoplasm after osteosarcoma in children. Amplifications in genomic, proteomic, and metabolism are characteristics of sarcoma, and targeting altered cancer cell molecular processes has been proposed as the latest promising strategy to fight cancer. Recent technological advancements have elucidated some of the underlying oncogenic characteristics of Ewing sarcoma. Offering new insights into the physiological basis for this phenomenon, our current review examines the dynamics of ES signaling as it related to both ES and the microenvironment by integrating genomic and proteomic analyses. An extensive survey of the literature was performed to compile the findings. We have also highlighted recent and ongoing studies integrating metabolomics and genomics aimed at better understanding the complex interactions as to how ES adapts to changing biochemical changes within the tumor microenvironment.

Graphical Abstract

1. Introduction

Ewing sarcoma (ES) is an aggressive tumor found often in adolescents, accounting for 10% to 15% of all bone sarcomas [1]. The “classic” Ewing bone sarcoma, extra-skeletal ES, malignant small cell tumor of the chest wall (Askin’s tumor), and primitive neuroectodermal tumors based on soft tissue (PNET) were all initially characterized by James Ewing in 1921. The highest incidence occurs in the second decade of life, with approximately 9–10 cases per million per year in patients aged 10–19 compared to an overall incidence of three cases per million per year in the population of the United States [2]. ES occurs predominantly in the Caucasian race and is very rare among African-Americans for unknown reasons [3]. There is a slight predominance of morbidity for men (sex ratio M:F 1.5:1) [3,4]. These tumors have a common genetic background: they are found in the pelvis, but they can appear in any bone or up to 30% of soft tissue [4,5]. Although sarcomas only account for 1% of all human malignancies [6], they are all aggressive [7,8,9] and can rapidly metastasize to the bone marrow, lungs, and other tissues [10,11].
This sarcoma subtype is likely derived from unique progenitor cells with similar genetic backgrounds belonging to endothelial, mesodermal, epithelial, and nerve cells [12,13], but most studies suggest that mesenchymal stem cells (MSCs) are the main precursor cells [14]. Ewing sarcoma belongs to the Ewing sarcoma family tumors (ESFT), which in turn belongs to the FET (FUS-EWSR1-TAF15) group of sarcomas and leukemias. The tumor group comprises more than 20 different tumor entities with ES as one of the most common types [15,16]. These tumors have similar non-random chromosomal translocations with one of the three genes, such as EWSR1, FUS, and TAF15, as 5′ partners and a large group of DNA binding transcription factor encoding genes as 3′ partners. Of note, these fusion proteins encode ETS biding factors that bind to similar DNA sequences. For ES the t(11;22)(q24;q12) translocation is present in 85–90% of tumors and EWSR1 and FLI1 gene fusion produces a fusion protein (EWSR1/FLI1). The t(11;22) (q24; q12) translocation is found in 85–90% of ES tumors, and the EWSR1 and FLI1 gene fusion results in a fusion protein (EWSR1/FLI1). Further, the t(21;22)(q22;q12), t(7;22)(q22;q12), t(7;22)(q22;q12), t(17;22)(q12;q12), t(2;22)(q33;q12) and others are less common translocations that result in the development of EWSR1/ERG, EWSR1/ETV1, EWSR1/ETV4 fusions of other genes and are present in the remaining 10–15% of the cases [11,17].
Local treatment and multi-agent adjuvant chemotherapy have increased the 5-year survival rate from less than 20% to more than 70% in the last 40 years, although the recurrence rate has remained high. Approximately 25% of people who have initially confined illness will have it recur at some point in their life. Because there is no standard therapy for recurrent and refractory ES, the 5-year OS for patients with a disease-free interval (DFI) > 2 years is around 30%, and the 5-year OS for those with a DFI of 2 years is about 7%. Given these factors, there is an immediate and unmet need for the development of novel ES therapies [18].

2. General Consideration: Hallmarks of Cancer

It is known that cancer cells are characterized by a set of features that distinguish them from non-neoplastic cells. The list of cancer hallmarks has been changed and refined over the years since it was first determined [19]. In 2022, Hanahan edited and expanded the list [20]. Nowadays, the authors of the concept propose nine hallmark capabilities: sustaining proliferative signaling, evading growth suppressors, activating invasion and metastasis, enabling replicative immortality, inducing angiogenesis, resisting cell death, avoiding immune destruction, deregulating cellular energetics, and unlocking phenotypic plasticity. In addition to these, there are also four enabling characteristics, by which cancer cells and tumors can adopt these functional capabilities: genome instability and mutations, tumor-promoting inflammation (with the effect of senescent cells), and non-mutational epigenetic reprogramming, and polymorphic microbiomes.
ES is, in contrast to most other sarcoma types, genetically stable, but the specific chromosomal translocation [11], for example, EWS-FLI1, is necessary for Ewing sarcoma tumorigenicity [21]. As shown earlier by Stolte et al. [21] FET tumors, besides gene fusion, are also characterized by wild type p53, suggesting a dependence of such tumors on DNA damage therapeutic stress, stabilized by p53 signaling. The only protein product of this gene can provide the transformation of cellular processes and the acquisition of hallmarks of cancer. Therefore, our review will focus on several genes and cellular proteins that hold promise for new therapeutic strategies. For ease of discussion and perception, we will consider the changed processes in three clusters: (1) Targeting of ES pressure on adhesion, migration, and invasion, (2) Targeting of Ewing sarcoma cells with a focus on proliferation, cell differentiation, and cell survival, and (3) Targeting of ES: Induction of apoptosis and cell cycle arrest.

3. Molecular Targets for ES Therapy

The two most common fusion proteins (EWSR1/FLI1 and EWSR1/ERG) [17] are involved in several cell signaling and regulatory pathways (Graphical Abstract). These fusion proteins usually act as transcription factors. For instance, Boulay et al. [22] identified a chromatin-binding factor (BAF) that interacts with EWSR1/FLI1 to activate gene expression in ES tumor cells and phenotypical changes. On the other hand, oncoprotein fusion can induce target genes via a partnership with GGAA microsatellites, as active enhancers [23] or by binding to RNA. Important discoveries in recent years have shown that major fractions of the ES fusion proteins bind to the SWI/SNF chromatin remodeling complex in tumor cells and that leads to the deregulation of gene expression, such as IGF-1 signaling [24] and epigenetic programming [15,16] towards retaining mesenchymal stem cell plasticity [25]. Blockade of these fusion proteins may be ideal for therapeutic targets. The feasibility of EWSR1/FLI1 targeting for ES therapy has been shown preclinically using chemical inhibitors [26,27,28] and siRNA technology (Bertrand et al. [29] Gauthier et al. [30] and Cervera et al. [31]) and via chemical targeting of EWSR1/FLI1 oncogenic fusion with chemical inhibitors [32]. Despite affecting Ewing sarcoma cell viability, multiple reports reveal induction of resistance factors contributing to the tumor’s survival and therapy escape [33,34,35]. Given this problem, it seems prudent to consider the combination of inhibitors against EWSR1/FLI1 or EWSR1/ERG with additional targeted therapies [36,37] that might overcome tumor cells’ resistance to the monotherapy and might demonstrate a less adverse effect due to dose reduction. One such drug is YK-4-279, a drug candidate active in downregulating transcription of EWS/FLI1 [38]. At the same time, YK-4-279 has been shown to inhibit ERG and ETV1 transcription in prostate cancer cells [39], therefore, it might also be active against Ewing sarcoma cells with commonly and less commonly presented fusion in the tumors with Ewing sarcoma. In our review, we want to propose potential targets for therapy that have not yet been used in treatment, along with inhibition of fusion proteins.

3.1. Targeting of ES Pressure on Adhesion, Migration, and Invasion

Like most cancer types, the prognosis for ES patients with localized disease is much better than for patients with metastatic disease [40]. A major obstacle in the battle against metastatic disease continues to be an insufficient understanding of underlying processes; specifically, what processes, such as metabolic and others, drive cell adhesion, migration, and invasion during metastases.
Adhesion proteins, such as E-cadherin are regarded as tumor suppressors. The low level of epithelial proteins is associated with spheroid formation and migration [41]. The migration of cancer cells to new niches is a fundamental process underlying metastasis [42]. There are two main types of cancer migration; mesenchymal and amoeboid; each is dependent on complex intracellular signaling that governs the actin cytoskeleton. Actin is involved in several pathways, including Rho/Rac GTPases and the Hippo-pathway, which culminate in the transcriptional regulation of cytoskeletal and growth-promoting genes, respectively. The mesenchymal cell migration of sarcomas can involve both single cells and cells in chains [43]. For successful metastasis, migration is followed by invasion into secondary locations [44].
Cadherins are vital for the formation of cell-cell contacts. During tumor progression, E-cadherin expression is lost, which permits epithelial to mesenchymal transition, anchorage-independent growth, and spheroid formation [45]. Studies have shown that loss of E-cadherin promotes resistance to treatment [40] and acquisition of a mesenchymal-like phenotype in vitro. Ex vivo studies have shown that the increased expression of E-cadherin is associated with improved clinical outcomes in several types of sarcomas [46]. Thus, E-cadherin upregulation in ES cells is mediated by epigenetics [47], or by small molecules like MIL327 [40] or RNA interference to metalloproteinase type 9 (MMP9) [48].
Besides induction of the epithelial to mesenchymal transition (EMT)-developed cell signaling, ES metastasis requires cell migration and blood vessel invasion [49]. The Rho-associated kinases, ROCK1 and ROCK2 have been implemented in the regulation of metastases using various in vitro and in vivo models. Data by Roberto et al. [49] showed a positive correlation between miR-139-5p and ROCK1, where restoration of miR-139-5p impaired ES migration and invasion. Another approach is the application of ROCK2 inhibitors, such as SR3677 and hydroxyfasudil, in SK-ES-1 cells [42] to demonstrate an opportunity for simultaneous targeting of heterogenic ES cells.
Although E-cadherin is generally regarded as an EMT marker, several studies have implicated E-cadherin in tumor adhesion via Ras homolog family member A (Rho A) mediated activation and alterations in paxillin [48]. Although data from Gluer et al. [50] suggested that expression levels of various adhesion markers in ES-based tumor cells matter most, cell adhesion is seemingly determined by the properties of integrin proteins, cadherin 11 [51], and neural cell adhesion molecule (NCAM) [50]. Kang et al. [52] investigated the link between E-cadherin and Erb-B2 Receptor Tyrosine Kinase 4 (ERBB4) using ES spheroids and demonstrated that E-cadherin localizes at cell-to-cell junctions and influences β-catenin cellular localization.
Other cellular proteins, such as integrins and cluster of differentiation 99 (CD99, MIC2) also play roles in ES cell-cell attachment. For example, the affinity of integrin contacts is reduced by CD99, a cell surface protein, via dephosphorylation of focal adhesion kinase (FAK) [53]. Although this protein is not a direct target of EWSR1/FLI1, it is important in several signaling pathways [54]. In ES cells, the CD99 protein also impacts cellular adhesion [55], cell growth [56], differentiation [57,58], and tumor cell apoptosis [59] through multiple mechanisms, such as NOTCH-, NFkappa-B, or MAPK modulation. This could be a promising target for therapy since activation with specific monoclonal bodies induces micropinocytosis and leads to cancer cell killing through a caspase-independent, non-apoptotic pathway resembling methuosis [60]. Furthermore, this could be pivotal in tumors with resistance to canonical apoptosis-inducing agents.
The other therapeutic possibility to target ES cells came from understanding the molecular mechanisms that were impacted by the EWSR1/FLI1 fusion. In sarcoma cells, Katsching et al. [61] reported that EWSR1/FLI1 directly binds to the proteins belonging to the YAP1/TAZ pathways. Along with another mechanism of Yes-associated protein 1 (YAP)/TAZ regulation, such as AP-1 [62] and activation of Myocardin-related transcription factor B (MRTFB) and TEA domain family member 1 (TEAD) [61], effectors of RhoA and Hippo signaling, EWSR1/FLI1 has an oscillating mechanism to regulate the balance between the epithelial to mesenchymal transition (EMT), along with the reverse state (MET). These oscillations provide cancer cells with several advantages, including successful invasion, migration, and consequently metastasis. Another transcriptional mechanism contributing to ES metastases involves the Sonic Hedgehog pathway. Previous studies implicated EWSR1/FLI1 in the regulation of glioma-associated oncogene 1 (GLI1 for humans and GLI1 in mouse models) [63,64,65]. Earlier, micro-array analysis of 27 ES patients showed an association between the Sonic Hedgehog pathway (SHH) and metastasis [66]. Besides metastases, this pathway regulates normal cell growth and proliferation. However, abnormalities can lead to the development of tumors, metastasis, and the emergence of antitumor drug resistance [67]. The binding of an SHH ligand to the patched (Ptch) receptor leads to a cell response expressed in the migration of GLI1 into the nucleus where it binds to DNA.
The canonical and non-canonical activities of GLI1 lead to the expression of genes, such as PTCH and SMO, which regulate the cell cycle and proliferation [68]. It has been shown that patient-derived ES cell lines (CHLA9, CHLA10, TC32, CHLA258, and TC71) and tumor samples [69] are enriched with GLI1. Abnormalities in this pathway can lead to tumorigenesis, metastasis, and drug resistance [67]. Therefore, YAP / TAZ or SHH pathway blockade holds promise as a drug resistance-preventive strategy. It has been proposed by Joo et al. [65] that the chemical compound GANT58 interferes with shared transcriptional downstream targets between GLI1 and EWSR1/FLI1. GANT61, another synthetic compound derived from hexahydropyrimidine, has been characterized as an inhibitor of GLI1 transcription that binds to the GLI1-DNA complex [70]. It was shown that ES-derived tumor cells can exploit both GLI1 and GLI2 [64], which may minimize the effect of GLI1 inhibitors. Given the use of GANT61 for inhibition of SHH-dependent targets, such as GLI1 [71], additional GLI1 regulation or investigations towards stability offer opportunities to target cellular proteins vital for ES transformation.
Insulin-like growth factor (IGF-1) is associated with several oncogenic processes in cells [36] and the IGF-1 receptor (IGF-1R) is a pivotal receptor tyrosine kinase that regulates malignant tumor transformation of ES cells [72]. IGF-1 and IGF-1R are overexpressed in the majority of Ewing sarcoma cell lines and have been previously studied as potential therapeutic targets in tumor treatment. It was reported that the IGF-1R pathway is activated in several cancers, such as hepatocellular carcinoma [73], pancreatic ductal carcinoma [74], retinoblastoma [75], colorectal cancer [76], and Ewing sarcoma [77,78]. Chromosomal translocation of EWSR1/FLI1 leads to activation of the IGF-1R pathway and its downstream targets belonging to the Mitogen-Activated Protein Kinase (MAPK) and the Phosphoinositide 3-kinase (PI3K) pathways [79] to maintain the phenotype and viability of ES cells. Although the use of IGF-1R inhibitors was efficacious in vitro, no sustained clinical response was found for ES patients [80].
It was discovered that inhibition of the IGF pathway initiated aberrant compensatory mechanisms, such as pregnancy-associated plasma protein-A (PAPP-A) [81]. Research findings have shown that synergistic action of PAPP-A and IGF-1R was more effective in EWS [81] and hence, their combined treatment could play a potential role in EWS therapy. However, this study was preclinical, and lacked an immunocompetent EWS model. Further research using an EWS model will be helpful for a better understanding of the immunomodulating effects of anti-PAPP-A. PAPP-A is a zinc metalloproteinase that cleaves inhibitory IGF-1-binding proteins, thereby increasing IGF-1 availability for IGF receptor-mediated cell proliferation, migration, and survival. PPAP-A enhances local IGF-1, which is also associated with invasion and metastasis. Potential targetable cell antigens in ES include PAPP-A as one of the top five secreted metalloproteinase proteins overexpressed in ES [81]. Furthermore, it has been shown that therapy with transgenic T cells directed against PAPP-A, knockout of the PAPP-A gene, and complex therapy in which not only PAPP-A is directly inhibited, can also be effective [82].
The TGF-β and PDGF pathways play important roles in ES plasticity and tumor progression. The TGF-β co-receptor endoglin is routinely expressed by malignant cells, and it is associated with the upregulation of bone morphogenetic protein, integrin, focal adhesion kinase, and phosphoinositide-3-kinase signaling, which together work in concert to maintain tumor cell plasticity [43,83]. Like TGF-β, the platelet-derived growth factor (PDGF) pathway aids in maintaining a cancer stem cell-like phenotype in ES, but more notably, is involved in ES tumor neovascularization [84]. Importantly PDGF ligands and/or receptors are frequently upregulated in ES [66,85] and their expression correlates with the activity of the EWSR1/FLI1 fusion [86] in sarcoma tissues. Overall, the pathways involved in ES cellular invasion and migration are complex, and future work is needed to design an effective multitargeted approach against ES tumor cells [87].
These genes play a key role in several tumor processes for ES, especially in adhesion, migration, and invasion (Table 1). The regulation of these processes in a tumor is the first step toward the formation of a lesion secondary tumor growth-metastasis from a localized tumor. Metastasis is associated with poor clinical outcomes for patients, while the treatment of localized tumors currently has a relatively high success rate, it could be as high as 70% [18]. Therefore, these genes as targets for therapy are especially vital.

3.2. Targeting of Ewing Sarcoma Cells with a Focus on Proliferation, Cell Differentiation, and Cell Survival

Metastatic progression requires the migration and invasion of cancer cells to reach distant tissues. However, for malignant neoplasms to spread beyond their primary tumor, they must adapt and thrive in a new niche. In general, cancer cells are characterized by unlimited time spent on cell division coupled with high survivability [88]. The mechanism for preserving these cells in a poorly differentiated state is elaborate. Mutations and epigenetic changes trigger unregulated mitotic cycles, allowing cells to become insensitive to growth-inhibitory signals, and capable of evading programmed cell death. Cell cycle progression and cell division, cell death, and cellular senescence determine cell proliferation in a broad sense.
The expression of the ES fusion gene alters the expression of over 500 downstream targets, which collectively block differentiation and drive proliferation [89]. For example, EWSR1/FLI1 regulates several genes, including IGF-1, Homeobox protein Nkx (NKX2), T-LAK cell-originated protein kinase (TOPK), SRY-Box Transcription Factor 2 (SOX2), and Enhancer Of ZesteHomolog2 (EZH2) [90]. It is worth noting that EWS-FLI1 binds to the EZH2 promoter to activate embryonic tumor stem cell growth and metastatic spread [91]. These targets can be used in the treatment of Ewing sarcoma. For example, GSK126 inhibits EZH2 methyltransferase activity in ES cells. Firstly, it reduces the phenotypic heterogeneity of ES cells and their capability for self-renewal and tumorigenicity. In turn, the change in the concentration of EZH2 targets (non-protein neuroectodermal marker GD2) allows chimeric antigen receptor gene-modified T cell therapy to be applied to the tumor [92]. Ahmed et al. [93] assessed the biomarkers responsible for ES cell proliferation. Immunostaining of primary tissue revealed that the majority were positive for protein kinase B (AKT) (55%) and mTOR (77%), indicating activation of an AKT-mTOR axis in ES cells. Thus, mTOR could be a budding target, some clinical trials for it are in process (see Section 4). Approximately 33% of specimens also expressed YAP, establishing a link between proliferation and YAP. Additionally, YAP is critically associated with BMI-1 (a polycomb complex protein) which stabilizes YAP expression and activity [94] via controlled chromatin remodeling.
Cell survival is connected with another hallmark of cancer deregulating cellular energetics. Numerous metabolic, i.e., catabolic and anabolic processes are altered in ES cells. For instance, Tanner et al. [95] reported altered de novo serine synthesis and dependence on aerobic glycolysis increased instead of oxidative metabolism. Moreover, data from an investigation by Sen et al. [96] demonstrates a direct link between EWSR1/FLI1 proteins involved in serine biosynthesis and glutamine consumption. Interestingly, Issaq et al. [97] showed that the expression of 3-phosphoglycerate dehydrogenase (PHGDH) was regulated by EWSR1/FLI1. PHGDH is one of the main enzymes that catalyze3-phosphoglycerate to 3-phosphohydroxypyruvate [98] (serine de novo synthesis) that is required for cell proliferation and tumor growth [97]. Further, PHGDH knockdown decreased ES cell proliferation and inhibited xenograft tumorigenesis in orthotopic ES models, providing an additional link between EWSR1/FLI1 and ES carcinogenesis. Therefore, targeting serine metabolism could be important for novel anticancer approaches in ES.
It is well-known that several cancers [99,100] heavily rely upon glycolysis rather than oxidative metabolism since it provides metabolic plasticity to fuel tumor heterogeneity [101,102]. Ewing sarcoma is one tumor type [103] where a predominant fusion protein, EWS/FLI1, regulates glucose consumption as well as gene expression of glycolytic enzymes, such as lactate dehydrogenase (LDH) [32]. It has been shown recently that depletion of lactate dehydrogenase-A (LDHA) inhibits proliferation of ES cells and induces apoptosis, impacting tumor cell viability both in vitro and in vivo.
Additionally, for various cancers, a growing number of important regulators are being discovered within one group of molecules: long non-coding RNAs (lncRNAs) [104]. LncRNAs often increase cancer cell survival, proliferation, colony formation, migration, and invasion [104,105,106]. Their elevated expression contributes to the progression of the sarcoma. It was described, for example, for lncRNA taurine upregulated gene 1 (TUG1) [106] and lncRNA SOX2 [105]. Many of the biological regulatory mechanisms of lncRNAs in Ewing sarcoma are still elusive. Nevertheless, they have been shown to often act as competing endogenous RNAs to regulate other genes expression. In principle, the knockdown of lncRNAs or the selection of inhibitory proteins might help to suppress ES growth [105,106]. These approaches might be potential therapeutic options for treating Ewing sarcoma.
In summary, cell proliferation, differentiation, and survival are important, first of all, in the formation of a tumor in a primary or secondary lesion. These properties of cells are related to some extent to other tumor characteristics (Figure 1), for example, deregulated cellular energetics (Table 1). To regulate this group of genes and their products, EWSR1/FLI1 usually acts as a transcription factor and binds with RNA.

3.3. Targeting of ES: Induction of Apoptosis and Cell Cycle Arrest

The development of fusion event inhibitors to suppress the progression of Ewing sarcoma requires the understanding of multiple cellular processes which became active in the tumor cells. Since kinases are integral to tumor maintenance, intervention directed toward these family members is promising. Both apoptosis and cell division utilize several different kinases, such as ATR and CHK1.
ES cells exhibit increased levels of endogenous DNA replicative stress and are sensitive to inhibitors of ribonucleotide reductase (RNR), an enzyme that limits the rate of deoxyribonucleotide synthesis. ES cells are also dependent on the ataxia telangiectasia, the rad3-related protein (ATR), and the checkpoint kinase 1 (CHK1) pathway, which play key roles in orchestrating the cellular response to DNA replication stress for survival [107,108]. ES tumors are sensitive both in vitro and in vivo to ATR and CHK1 inhibitors as separate agents and in combination with other drugs [107,108,109,110,111]. The ATR-CHK1 pathway, when activated by DNA replication stress, orchestrates a multifaceted response that arrests cell cycle progression, suppresses the origin of replication, stabilizes replication forks, and promotes fork repair and restart [112].
However, ATR and CHK1 also have critical and unique functions outside of the S phase and the response to DNA replication stress. For example, ATR and/or CHK1 regulate chromosome segregation, the S/G2 checkpoint, the G2/M transition, double-strand DNA break repair, and the response to osmotic and mechanical stress [113,114]. Recently, Koppenhafer et al. [87] identified that the inhibition of the ATR-CHK1 pathway in ES cells under DNA replication stress leads to the aberrant activation of Cyclin-Dependent Kinase 2 (CDK2) and cell death. CDK1 and CDK2 are critical mediators of cell cycle progression that are regulated by the ATR-CHK1-CDC25A pathway. In the setting of DNA replication stress, ATR-CHK1 negatively regulates CDC25A, which de-phosphorylates and activates CDK1/2 to restrain cell cycle progression and promote DNA damage repair. A novel feedback intracellular loop in Ewing sarcoma cells has been discovered. In this loop, the inhibition of the ATR-CHK1 pathway, or the WEE1 kinase, during DNA replication stress leads to enhanced DNA replication stress, increased DNA damage, and apoptosis. Although most investigations in this field focus on the interaction between EWS/FLI1 and intracellular pathways, several others have explored the possibility of interfering with extracellular signaling paths that regulate EWS/FLI1, and consequently, tumor transformation.
Additionally, one of the enzymes regulating the work of the hereditary apparatus of the cell is Poly (ADP-ribose) polymerase (PARP), which is involved in the processes of DNA repair, maintaining the genetic stability of the cell and its programmed death, and as Ewing sarcoma cell lines are frequently defective in DNA break repair, they are susceptible to PARP inhibition [115,116,117]. Inhibition of PARP showed the effectiveness of this approach if cytostatic drugs were used, the effect of which was intensified [118]. Some studies indicate an increase in the effectiveness of treatment of Ewing sarcoma with PARP inhibitors alone [119] or with a combination, such as temozolomide [120,121]. Although preclinical in vitro models showed an acceptable result, the activity of PARP inhibitors as a single agent in preclinical in vivo models and clinical trials at an early stage of Ewing sarcoma did not demonstrate significant results [122]. The activity of PARP directly depends on its required substrate, nicotinamide adenine dinucleotide (NAD+), which is produced by nicotinamide phosphoribosyltransferase (NAMPT). Studies of the combined use of PARP and NAMPT inhibitors in vivo have shown great effectiveness. The combined therapy resulted in tumor regression, delayed disease progression, and increased survival [123]. Considering that Ewing sarcoma cells depend on functioning PARP, that PARP requires NAD+, and that NAD+ production depends on NAMPT, it seems appropriate to simultaneously inhibit these proteins, which are confirmed by recent studies [123].
Growth and differentiation factor 6 (GDF6), also known as bone morphogenetic protein 13 (BMP13), belongs to the TGF-superfamily’s BMP family. GDF6 has become an attractive target since its binding to its own receptor, such as CD99 [41]. GDF6 is vital to embryogenesis, particularly to the development of the neural and skeletal systems, and mutations in GDF6 are associated with abnormalities of the skeleton, the eyes [124], and other organs [125]. GDF6 is highly expressed in ES tumors and cell lines compared to mesenchymal stem cells and cells from other sarcoma subtypes.
Zhou et al. [41] described a GDF6 prodomain signaling pathway that regulates Src activity and ES tumor growth [41] via p21. A ChIP-sequencing showed binding of EWSR1/FLI1 to the GDF6 gene in ES cells, which implicates GDF6 as a direct target of EWSR1/FLI1 transcription activation. Considering the role of GDF6 in cell proliferation and differentiation, inactivation of gene expression [41] will offer a possibility to interfere with ES cell growth.
The anticancer properties of multiple ES-based therapeutic approaches have been extensively studied, often for their antiproliferative effects. The induction of apoptosis has been reported for several cancers, including Ewing sarcoma. However, ES cells are prone to developing treatment resistance, which contributes to disease recurrence [126]. Current ES-therapeutic strategies are promising, and the development of new therapeutics and combinations thereof with greater antitumoral properties has been proposed. For instance, Sonnemann et al. [127] investigated cell death mediated by histone deacetylase (HDAC) inhibitors in the presence of pro-apoptotic TNF-related apoptosis-inducing ligand (TRAIL). Additionally, Lu et al. [128] found that the proteasome inhibitor bortezomib synergizes with TRAIL in vitro using TC-71, to enhance cancer cell-related toxicity through apoptosis.
Previously, TRAIL, a member of the tumor necrosis factor (TNF) ligand superfamily (TNFLSF), has been found to induce apoptosis in cancer cells while sparing normal cells [129,130,131]. Multiple in vitro studies have shown that TRAIL and other death receptor agonists [132,133] are effective against sarcoma cell lines [131,134] with ES cell lines showing the greatest sensitivity [135]. Ubiquitin-specific protease 6 (USP6) may also contribute to the sensitization of ES cells to exogenous IFNs [136]. Henrich and colleagues [136] speculate that this negative feedback loop involves USP6, which serves to amplify Interferon Gamma (IFN)-mediated sensitivity to TRAIL. Indisputably, the molecular mechanism of TRAIL sensitivity warrants additional investigation to clarify the molecular basis for drug synergy against ES.
The metabolism of Ewing sarcoma involves many genes and metabolic pathways that may be potential targets for therapy (Table 1). The effect on some agents can lead to a static effect, and the activation or silencing of others can be expressed as a bright antitumor effect and lead to apoptosis. Some experimental approaches to stimulating cell cycle arrest [137,138,139] or apoptosis [140,141] in Ewing sarcoma cells showed effectiveness in preclinical settings, and therefore, might hold great promises in future clinical testing.

4. Current Clinical Trials

Standard chemotherapies, such as alkylating agents, topoisomerase, and tubulin inhibitors are non-specific and exert their effects on both tumor cells and normal cells. Currently, a search is underway to identify new drugs that are specifically targeted for ES cells and that are capable of eliminating the tumor cells and extending patient survival. Some clinical trials where traditional and experimental components of chemotherapy were used both separately and in combination with each other are presented (Table 2). The key to new and successful therapies may be the addition of standard treatment protocols with new, highly specific experimental drugs.

5. Conclusions

Ewing sarcoma is a cancer with metabolic processes and related pathology largely governed by fusion proteins. Unfortunately, inhibition of these proteins themselves has proved challenging and clinically unsuccessful, which necessitates combinations of new therapeutic approaches. After an extensive literature review, we chose 17 molecules that serve as promising targets for therapy that alter cell metabolism, and possess features crucial in tumorigenesis, including cell adhesion, migration, invasion, proliferation, differentiation, survival, apoptosis, and cell cycle arrest (Figure 1). Most of the genes we have described are direct targets of fusion proteins, therefore, successful indirect inhibition could have a cascading effect on cell survival and might have future clinical implications. In addition, many of the presented proteins are often highly expressed (some serve as markers for ES), so their inhibition could be readily available and exert a strong anti-cancer effect. Clinical trials are currently underway for some of these aforementioned target molecules. Therapies devoted to targeting them alone or in combination with current regimens could become the next standard of care for Ewing sarcoma patients.

Author Contributions

D.F., Conceptualization, Methodology, Investigation, Writing—Original Draft, Writing—Review & Editing; S.T., Conceptualization, Writing—Original Draft, Visualization; M.S., Writing—Review & Editing; B.A.S., Writing—Review & Editing; Writing—Original Draft, Writing—Review & Editing; A.A., Writing—Original Draft, Writing—Review & Editing; R.K.K., Writing—Original Draft, Writing—Review & Editing; N.K., Writing—Review & Editing; P.T., Writing—Review & Editing; I.U., Conceptualization, Supervision, Writing—Original Draft, Writing—Review & Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Russian Science Foundation under grant No 21-15-00213 (IU, description of molecular mechanisms and metabolism feature of ES).

Conflicts of Interest

The authors declare no financial conflict.

Abbreviation

CD99cluster of differentiation 99
EMTepithelial to mesenchymal transition
ERBB4Erb-B2 Receptor Tyrosine Kinase 4
ESEwing sarcoma
ESFTEwing sarcoma family tumors
EZH2Enhancer Of Zeste 2 Polycomb Repressive Complex 2 Subunit
GDF6growth and differentiation factor 6
FETfamily of genes, Fused in sarcoma, Ewing sarcoma breakpoint region 1, TATA-box binding protein associated factor 15
GLI1glioma-associated oncogene 1
IGF-1Insulin-like growth factor
IGF-1Rinsulin-like growth factor 1 receptor
LDHlactate dehydrogenase
lncRNAslong non-coding RNAs
MAPKMitogen-Activated Protein Kinase
METmesenchymal to epithelial transition
MMP9metalloproteinase type 9
MRTFBof Myocardin-related transcription factor B
NAMPTnicotinamide phosphoribosyl transferase
NCAMneural cell adhesion molecule
PAPP-APregnancy-associated plasma protein A
PARPPoly (ADP-ribose) polymerase
PDGFPlatelet-derived growth factor
PHGDH3-phosphoglycerate dehydrogenase
PI3KPhosphoinositide 3-kinases
ROCKRho-associated coiled-coil kinase
RhoARas homolog family member A
SHHSonic Hedgehog pathway
SOX2SRY-Box Transcription Factor 2
SRBCTSmall Round Blue Cell Tumors
TEADTEA domain family member 1
TNFtumor necrosis factor
mTORmammalian target of rapamycin
TRAILTNF-Related Apoptosis Inducing Ligand
PNETPrimitive neuroectodermal tumor
VEGF-AVascular endothelial growth factor A
YAPYes-associated protein.

References

  1. Mer, A.S.; Ba-Alawi, W.; Smirnov, P.; Wang, Y.X.; Brew, B.; Ortmann, J.; Tsao, M.S.; Cescon, D.W.; Goldenberg, A.; Haibe-Kains, B. Integrative Pharmacogenomics Analysis of Patient-Derived Xenografts. Cancer Res. 2019, 79, 4539–4550. [Google Scholar] [CrossRef] [PubMed]
  2. Esiashvili, N.; Goodman, M.; Marcus, R.B., Jr. Changes in incidence and survival of Ewing sarcoma patients over the past 3 decades: Surveillance Epidemiology and End Results data. J. Pediatr. Hematol. Oncol. 2008, 30, 425–430. [Google Scholar] [CrossRef] [PubMed]
  3. Jawad, M.U.; Cheung, M.C.; Min, E.S.; Schneiderbauer, M.M.; Koniaris, L.G.; Scully, S.P. Ewing sarcoma demonstrates racial disparities in incidence-related and sex-related differences in outcome: An analysis of 1631 cases from the SEER database, 1973–2005. Cancer 2009, 115, 3526–3536. [Google Scholar] [CrossRef] [PubMed]
  4. Ordonez, J.L.; Osuna, D.; Herrero, D.; de Alava, E.; Madoz-Gurpide, J. Advances in Ewing’s sarcoma research: Where are we now and what lies ahead? Cancer Res. 2009, 69, 7140–7150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. 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]
  6. Clark, M.A.; Fisher, C.; Judson, I.; Thomas, J.M. Soft-tissue sarcomas in adults. N. Engl. J. Med. 2005, 353, 701–711. [Google Scholar] [CrossRef] [Green Version]
  7. Cabral, A.N.F.; Rocha, R.H.; Amaral, A.; Medeiros, K.B.; Nogueira, P.S.E.; Diniz, L.M. Cutaneous angiosarcoma: Report of three different and typical cases admitted in a unique dermatology clinic. An. Bras. Dermatol. 2017, 92, 235–238. [Google Scholar] [CrossRef] [Green Version]
  8. Kruse, A.J.; Croce, S.; Kruitwagen, R.F.; Riedl, R.G.; Slangen, B.F.; Van Gorp, T.; Van de Vijver, K.K. Aggressive behavior and poor prognosis of endometrial stromal sarcomas with YWHAE-FAM22 rearrangement indicate the clinical importance to recognize this subset. Int. J. Gynecol. Cancer 2014, 24, 1616–1622. [Google Scholar] [CrossRef]
  9. Coates, S.J.; Ogunrinade, O.; Lee, H.J.; Desman, G. Epidermotropic metastatic epithelioid sarcoma: A potential diagnostic pitfall. J Cutan Pathol 2014, 41, 672–676. [Google Scholar] [CrossRef]
  10. Lin, P.P.; Wang, Y.; Lozano, G. Mesenchymal Stem Cells and the Origin of Ewing’s Sarcoma. Sarcoma 2011, 2011, 276463. [Google Scholar] [CrossRef] [Green Version]
  11. Riggi, N.; Stamenkovic, I. The Biology of Ewing sarcoma. Cancer Lett. 2007, 254, 1–10. [Google Scholar] [CrossRef] [PubMed]
  12. Lawlor, E.R.; Lim, J.F.; Tao, W.; Poremba, C.; Chow, C.J.; Kalousek, I.V.; Kovar, H.; MacDonald, T.J.; Sorensen, P.H. The Ewing tumor family of peripheral primitive neuroectodermal tumors expresses human gastrin-releasing peptide. Cancer Res. 1998, 58, 2469–2476. [Google Scholar] [PubMed]
  13. O’Regan, S.; Diebler, M.F.; Meunier, F.M.; Vyas, S. A Ewing’s sarcoma cell line showing some, but not all, of the traits of a cholinergic neuron. J. Neurochem. 1995, 64, 69–76. [Google Scholar] [CrossRef] [PubMed]
  14. 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] [Green Version]
  15. Linden, M.; Thomsen, C.; Grundevik, P.; Jonasson, E.; Andersson, D.; Runnberg, R.; Dolatabadi, S.; Vannas, C.; Luna Santamariotaa, M.; Fagman, H.; et al. FET family fusion oncoproteins target the SWI/SNF chromatin remodeling complex. EMBO Rep. 2019, 20, e45766. [Google Scholar] [CrossRef] [PubMed]
  16. Linden, M.; Vannas, C.; Osterlund, T.; Andersson, L.; Osman, A.; Escobar, M.; Fagman, H.; Stahlberg, A.; Aman, P. FET fusion oncoproteins interact with BRD4 and SWI/SNF chromatin remodelling complex subtypes in sarcoma. Mol. Oncol. 2022. [Google Scholar] [CrossRef]
  17. Trancau, I.O. Chromosomal translocations highlighted in Primitive Neuroectodermal Tumors (PNET) and Ewing sarcoma. J. Med. Life 2014, 7, 44–50. [Google Scholar]
  18. Van Mater, D.; Wagner, L. Management of recurrent Ewing sarcoma: Challenges and approaches. Onco. Targets Ther. 2019, 12, 2279–2288. [Google Scholar] [CrossRef] [Green Version]
  19. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef] [Green Version]
  20. Hanahan, D. Hallmarks of Cancer: New Dimensions. Cancer Discov. 2022, 12, 31–46. [Google Scholar] [CrossRef]
  21. Stolte, B.; Iniguez, A.B.; Dharia, N.V.; Robichaud, A.L.; Conway, A.S.; Morgan, A.M.; Alexe, G.; Schauer, N.J.; Liu, X.; Bird, G.H.; et al. Genome-scale CRISPR-Cas9 screen identifies druggable dependencies in TP53 wild-type Ewing sarcoma. J. Exp. Med. 2018, 215, 2137–2155. [Google Scholar] [CrossRef] [PubMed]
  22. 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]
  23. Marchetto, A.; Ohmura, S.; Orth, M.F.; Knott, M.M.L.; Colombo, M.V.; Arrigoni, C.; Bardinet, V.; Saucier, D.; Wehweck, F.S.; Li, J.; et al. Oncogenic hijacking of a developmental transcription factor evokes vulnerability toward oxidative stress in Ewing sarcoma. Nat. Commun. 2020, 11, 2423. [Google Scholar] [CrossRef] [PubMed]
  24. Cironi, L.; Riggi, N.; Provero, P.; Wolf, N.; Suva, M.L.; Suva, D.; Kindler, V.; Stamenkovic, I. IGF1 is a common target gene of Ewing’s sarcoma fusion proteins in mesenchymal progenitor cells. PLoS ONE 2008, 3, e2634. [Google Scholar] [CrossRef] [PubMed]
  25. Riggi, N.; Suva, M.L.; De Vito, C.; Provero, P.; Stehle, J.C.; Baumer, K.; Cironi, L.; Janiszewska, M.; Petricevic, T.; Suva, 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]
  26. 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] [Green Version]
  27. Welch, D.; Kahen, E.; Fridley, B.; Brohl, A.S.; Cubitt, C.L.; Reed, D.R. Small molecule inhibition of lysine-specific demethylase 1 (LSD1) and histone deacetylase (HDAC) alone and in combination in Ewing sarcoma cell lines. PLoS ONE 2019, 14, e0222228. [Google Scholar] [CrossRef]
  28. Grohar, P.J.; Woldemichael, G.M.; Griffin, L.B.; Mendoza, A.; Chen, Q.R.; Yeung, C.; Currier, D.G.; Davis, S.; Khanna, C.; Khan, J.; et al. Identification of an inhibitor of the EWS-FLI1 oncogenic transcription factor by high-throughput screening. J. Natl. Cancer Inst. 2011, 103, 962–978. [Google Scholar] [CrossRef] [Green Version]
  29. Bertrand, J.R.; Pioche-Durieu, C.; Ayala, J.; Petit, T.; Girard, H.A.; Malvy, C.P.; Le Cam, E.; Treussart, F.; Arnault, J.C. Plasma hydrogenated cationic detonation nanodiamonds efficiently deliver to human cells in culture functional siRNA targeting the Ewing sarcoma junction oncogene. Biomaterials 2015, 45, 93–98. [Google Scholar] [CrossRef]
  30. Gauthier, F.; Claveau, S.; Bertrand, J.R.; Vasseur, J.J.; Dupouy, C.; Debart, F. Gymnotic delivery and gene silencing activity of reduction-responsive siRNAs bearing lipophilic disulfide-containing modifications at 2′-position. Bioorg. Med. Chem. 2018, 26, 4635–4643. [Google Scholar] [CrossRef]
  31. Cervera, S.T.; Rodriguez-Martin, C.; Fernandez-Tabanera, E.; Melero-Fernandez de Mera, R.M.; Morin, M.; Fernandez-Penalver, S.; Iranzo-Martinez, M.; Amhih-Cardenas, J.; Garcia-Garcia, L.; Gonzalez-Gonzalez, L.; et al. Therapeutic Potential of EWSR1-FLI1 Inactivation by CRISPR/Cas9 in Ewing Sarcoma. Cancers 2021, 13, 3783. [Google Scholar] [CrossRef] [PubMed]
  32. Yeung, C.; Gibson, A.E.; Issaq, S.H.; Oshima, N.; Baumgart, J.T.; Edessa, L.D.; Rai, G.; Urban, D.J.; Johnson, M.S.; Benavides, G.A.; et al. Targeting Glycolysis through Inhibition of Lactate Dehydrogenase Impairs Tumor Growth in Preclinical Models of Ewing Sarcoma. Cancer Res. 2019, 79, 5060–5073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Heisey, D.A.R.; Lochmann, T.L.; Floros, K.V.; Coon, C.M.; Powell, K.M.; Jacob, S.; Calbert, M.L.; Ghotra, M.S.; Stein, G.T.; Maves, Y.K.; et al. The Ewing Family of Tumors Relies on BCL-2 and BCL-XL to Escape PARP Inhibitor Toxicity. Clin. Cancer Res. 2019, 25, 1664–1675. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Palombo, R.; Verdile, V.; Paronetto, M.P. Poison-Exon Inclusion in DHX9 Reduces Its Expression and Sensitizes Ewing Sarcoma Cells to Chemotherapeutic Treatment. Cells 2020, 9, 328. [Google Scholar] [CrossRef] [Green Version]
  35. Radic-Sarikas, B.; Tsafou, K.P.; Emdal, K.B.; Papamarkou, T.; Huber, K.V.; Mutz, C.; Toretsky, J.A.; Bennett, K.L.; Olsen, J.V.; Brunak, S.; et al. Combinatorial Drug Screening Identifies Ewing Sarcoma-specific Sensitivities. Mol. Cancer Ther. 2017, 16, 88–101. [Google Scholar] [CrossRef] [Green Version]
  36. Lessnick, S.L.; Ladanyi, M. Molecular pathogenesis of Ewing sarcoma: New therapeutic and transcriptional targets. Annu. Rev. Pathol. 2012, 7, 145–159. [Google Scholar] [CrossRef] [Green Version]
  37. Spriano, F.; Chung, E.Y.L.; Gaudio, E.; Tarantelli, C.; Cascione, L.; Napoli, S.; Jessen, K.; Carrassa, L.; Priebe, V.; Sartori, G.; et al. The ETS Inhibitors YK-4-279 and TK-216 Are Novel Antilymphoma Agents. Clin. Cancer Res. 2019, 25, 5167–5176. [Google Scholar] [CrossRef]
  38. Lamhamedi-Cherradi, S.E.; Menegaz, B.A.; Ramamoorthy, V.; Aiyer, R.A.; Maywald, R.L.; Buford, A.S.; Doolittle, D.K.; Culotta, K.S.; O’Dorisio, J.E.; Ludwig, J.A. An Oral Formulation of YK-4-279: Preclinical Efficacy and Acquired Resistance Patterns in Ewing Sarcoma. Mol. Cancer Ther. 2015, 14, 1591–1604. [Google Scholar] [CrossRef] [Green Version]
  39. Rahim, S.; Minas, T.; Hong, S.H.; Justvig, S.; Celik, H.; Kont, Y.S.; Han, J.; Kallarakal, A.T.; Kong, Y.; Rudek, M.A.; et al. A small molecule inhibitor of ETV1, YK-4-279, prevents prostate cancer growth and metastasis in a mouse xenograft model. PLoS ONE 2014, 9, e114260. [Google Scholar] [CrossRef] [Green Version]
  40. Rellinger, E.J.; Padmanabhan, C.; Qiao, J.; Appert, A.; Waterson, A.G.; Lindsley, C.W.; Beauchamp, R.D.; Chung, D.H. ML327 induces apoptosis and sensitizes Ewing sarcoma cells to TNF-related apoptosis-inducing ligand. Biochem. Biophys. Res. Commun. 2017, 491, 463–468. [Google Scholar] [CrossRef]
  41. Zhou, F.; Elzi, D.J.; Jayabal, P.; Ma, X.; Chiu, Y.C.; Chen, Y.; Blackman, B.; Weintraub, S.T.; Houghton, P.J.; Shiio, Y. GDF6-CD99 Signaling Regulates Src and Ewing Sarcoma Growth. Cell. Rep. 2020, 33, 108332. [Google Scholar] [CrossRef] [PubMed]
  42. Pinca, R.S.; Manara, M.C.; Chiadini, V.; Picci, P.; Zucchini, C.; Scotlandi, K. Targeting ROCK2 rather than ROCK1 inhibits Ewing sarcoma malignancy. Oncol. Rep. 2017, 37, 1387–1393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ehnman, M.; Chaabane, W.; Haglund, F.; Tsagkozis, P. The Tumor Microenvironment of Pediatric Sarcoma: Mesenchymal Mechanisms Regulating Cell Migration and Metastasis. Curr. Oncol. Rep. 2019, 21, 90. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Codman, E.A. The classic: Registry of bone sarcoma: Part I.—Twenty-five criteria for establishing the diagnosis of osteogenic sarcoma. Part II.—Thirteen registered cases of “five year cures” analyzed according to these criteria. 1926. Clin. Orthop. Relat. Res. 2009, 467, 2771–2782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Jolly, M.K.; Ware, K.E.; Xu, S.; Gilja, S.; Shetler, S.; Yang, Y.; Wang, X.; Austin, R.G.; Runyambo, D.; Hish, A.J.; et al. E-Cadherin Represses Anchorage-Independent Growth in Sarcomas through Both Signaling and Mechanical Mechanisms. Mol. Cancer Res. 2019, 17, 1391–1402. [Google Scholar] [CrossRef] [Green Version]
  46. Machado, I.; Navarro, S.; Lopez-Guerrero, J.A.; Alberghini, M.; Picci, P.; Llombart-Bosch, A. Epithelial marker expression does not rule out a diagnosis of Ewing’s sarcoma family of tumours. Virchows Arch. 2011, 459, 409–414. [Google Scholar] [CrossRef]
  47. Hurtubise, A.; Bernstein, M.L.; Momparler, R.L. Preclinical evaluation of the antineoplastic action of 5-aza-2′-deoxycytidine and different histone deacetylase inhibitors on human Ewing’s sarcoma cells. Cancer Cell Int. 2008, 8, 16. [Google Scholar] [CrossRef] [Green Version]
  48. Sanceau, J.; Truchet, S.; Bauvois, B. Matrix metalloproteinase-9 silencing by RNA interference triggers the migratory-adhesive switch in Ewing’s sarcoma cells. J. Biol. Chem. 2003, 278, 36537–36546. [Google Scholar] [CrossRef] [Green Version]
  49. Roberto, G.M.; Delsin, L.E.A.; Vieira, G.M.; Silva, M.O.; Hakime, R.G.; Gava, N.F.; Engel, E.E.; Scrideli, C.A.; Tone, L.G.; Brassesco, M.S. ROCK1-PredictedmicroRNAs Dysregulation Contributes to Tumor Progression in Ewing Sarcoma. Pathol. Oncol. Res. 2020, 26, 133–139. [Google Scholar] [CrossRef]
  50. Gluer, S.; Zense, M.; von Schweinitz, D. Cell adhesion molecules and intermediate filaments on embryonal childhood tumors. Pathol. Res. Pract. 1998, 194, 773–780. [Google Scholar] [CrossRef]
  51. Hatano, M.; Matsumoto, Y.; Fukushi, J.; Matsunobu, T.; Endo, M.; Okada, S.; Iura, K.; Kamura, S.; Fujiwara, T.; Iida, K.; et al. Cadherin-11 regulates the metastasis of Ewing sarcoma cells to bone. Clin. Exp. Metastasis 2015, 32, 579–591. [Google Scholar] [CrossRef] [PubMed]
  52. Kang, H.G.; Jenabi, J.M.; Zhang, J.; Keshelava, N.; Shimada, H.; May, W.A.; Ng, T.; Reynolds, C.P.; Triche, T.J.; Sorensen, P.H. E-cadherin cell-cell adhesion in ewing tumor cells mediates suppression of anoikis through activation of the ErbB4 tyrosine kinase. Cancer Res. 2007, 67, 3094–3105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lee, K.J.; Kim, Y.; Yoo, Y.H.; Kim, M.S.; Lee, S.H.; Kim, C.G.; Park, K.; Jeoung, D.; Lee, H.; Ko, I.Y.; et al. CD99-Derived Agonist Ligands Inhibit Fibronectin-Induced Activation of beta1 Integrin through the Protein Kinase A/SHP2/Extracellular Signal-Regulated Kinase/PTPN12/Focal Adhesion Kinase Signaling Pathway. Mol. Cell Biol. 2017, 37, e00675-16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Franzetti, G.A.; Laud-Duval, K.; Bellanger, D.; Stern, M.H.; Sastre-Garau, X.; Delattre, O. MiR-30a-5p connects EWS-FLI1 and CD99, two major therapeutic targets in Ewing tumor. Oncogene 2013, 32, 3915–3921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Cerisano, V.; Aalto, Y.; Perdichizzi, S.; Bernard, G.; Manara, M.C.; Benini, S.; Cenacchi, G.; Preda, P.; Lattanzi, G.; Nagy, B.; et al. Molecular mechanisms of CD99-induced caspase-independent cell death and cell-cell adhesion in Ewing’s sarcoma cells: Actin and zyxin as key intracellular mediators. Oncogene 2004, 23, 5664–5674. [Google Scholar] [CrossRef] [Green Version]
  56. Huijbers, E.J.M.; van der Werf, I.M.; Faber, L.D.; Sialino, L.D.; van der Laan, P.; Holland, H.A.; Cimpean, A.M.; Thijssen, V.; van Beijnum, J.R.; Griffioen, A.W. Targeting Tumor Vascular CD99 Inhibits Tumor Growth. Front. Immunol. 2019, 10, 651. [Google Scholar] [CrossRef]
  57. De Feo, A.; Sciandra, M.; Ferracin, M.; Felicetti, F.; Astolfi, A.; Pignochino, Y.; Picci, P.; Care, A.; Scotlandi, K. Exosomes from CD99-deprived Ewing sarcoma cells reverse tumor malignancy by inhibiting cell migration and promoting neural differentiation. Cell Death Dis. 2019, 10, 471. [Google Scholar] [CrossRef]
  58. Rocchi, A.; Manara, M.C.; Sciandra, M.; Zambelli, D.; Nardi, F.; Nicoletti, G.; Garofalo, C.; Meschini, S.; Astolfi, A.; Colombo, M.P.; et al. CD99 inhibits neural differentiation of human Ewing sarcoma cells and thereby contributes to oncogenesis. J. Clin. Investig. 2010, 120, 668–680. [Google Scholar] [CrossRef]
  59. Sohn, H.W.; Choi, E.Y.; Kim, S.H.; Lee, I.S.; Chung, D.H.; Sung, U.A.; Hwang, D.H.; Cho, S.S.; Jun, B.H.; Jang, J.J.; et al. Engagement of CD99 induces apoptosis through a calcineurin-independent pathway in Ewing’s sarcoma cells. Am. J. Pathol. 1998, 153, 1937–1945. [Google Scholar] [CrossRef]
  60. Manara, M.C.; Terracciano, M.; Mancarella, C.; Sciandra, M.; Guerzoni, C.; Pasello, M.; Grilli, A.; Zini, N.; Picci, P.; Colombo, M.P.; et al. CD99 triggering induces methuosis of Ewing sarcoma cells through IGF-1R/RAS/Rac1 signaling. Oncotarget 2016, 7, 79925–79942. [Google Scholar] [CrossRef] [Green Version]
  61. 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] [PubMed] [Green Version]
  62. Tomazou, E.M.; Sheffield, N.C.; Schmidl, C.; Schuster, M.; Schonegger, A.; Datlinger, P.; Kubicek, S.; Bock, C.; Kovar, H. Epigenome mapping reveals distinct modes of gene regulation and widespread enhancer reprogramming by the oncogenic fusion protein EWS-FLI1. Cell Rep. 2015, 10, 1082–1095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zwerner, J.P.; Joo, J.; Warner, K.L.; Christensen, L.; Hu-Lieskovan, S.; Triche, T.J.; May, W.A. The EWS/FLI1 oncogenic transcription factor deregulates GLI1. Oncogene 2008, 27, 3282–3291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Beauchamp, E.; Bulut, G.; Abaan, O.; Chen, K.; Merchant, A.; Matsui, W.; Endo, Y.; Rubin, J.S.; Toretsky, J.; Uren, A. GLI1 is a direct transcriptional target of EWS-FLI1 oncoprotein. J. Biol. Chem. 2009, 284, 9074–9082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Joo, J.; Christensen, L.; Warner, K.; States, L.; Kang, H.G.; Vo, K.; Lawlor, E.R.; May, W.A. GLI1 is a central mediator of EWS/FLI1 signaling in Ewing tumors. PLoS ONE 2009, 4, e7608. [Google Scholar] [CrossRef] [PubMed]
  66. Schaefer, K.L.; Eisenacher, M.; Braun, Y.; Brachwitz, K.; Wai, D.H.; Dirksen, U.; Lanvers-Kaminsky, C.; Juergens, H.; Herrero, D.; Stegmaier, S.; et al. Microarray analysis of Ewing’s sarcoma family of tumours reveals characteristic gene expression signatures associated with metastasis and resistance to chemotherapy. Eur. J. Cancer 2008, 44, 699–709. [Google Scholar] [CrossRef]
  67. Giroux-Leprieur, E.; Costantini, A.; Ding, V.W.; He, B. Hedgehog Signaling in Lung Cancer: From Oncogenesis to Cancer Treatment Resistance. Int. J. Mol. Sci. 2018, 19, 2835. [Google Scholar] [CrossRef] [Green Version]
  68. Niyaz, M.; Khan, M.S.; Mudassar, S. Hedgehog Signaling: An Achilles’ Heel in Cancer. Transl. Oncol. 2019, 12, 1334–1344. [Google Scholar] [CrossRef]
  69. Mullard, M.; Cade, M.; Morice, S.; Dupuy, M.; Danieau, G.; Amiaud, J.; Renault, S.; Lezot, F.; Brion, R.; Thepault, R.A.; et al. Sonic Hedgehog Signature in Pediatric Primary Bone Tumors: Effects of the GLI Antagonist GANT61 on Ewing’s Sarcoma Tumor Growth. Cancers 2020, 12, 3438. [Google Scholar] [CrossRef]
  70. Matsumoto, T.; Tabata, K.; Suzuki, T. The GANT61, a GLI inhibitor, induces caspase-independent apoptosis of SK-N-LO cells. Biol. Pharm. Bull. 2014, 37, 633–641. [Google Scholar] [CrossRef] [Green Version]
  71. Bacelar Sacramento de Araujo, T.; de Oliveira Siquara da Rocha, L.; Torres Andion Vidal, M.; Cerqueira Coelho, P.L.; Galvao Dos Reis, M.; Solano de Freitas Souza, B.; Botelho Pereira Soares, M.; Almeida Pereira, T.; Della Coletta, R.; Pereira Bezerra, D.; et al. GANT61 Reduces Hedgehog Molecule (GLI1) Expression and Promotes Apoptosis in Metastatic Oral Squamous Cell Carcinoma Cells. Int. J. Mol. Sci. 2020, 21, 6076. [Google Scholar] [CrossRef] [PubMed]
  72. Karakus, R.; Karakus, E.; Emir, S.; Kacar, A.; Ozyoruk, D. Insulin-like growth factor-1 receptor expression in pediatric tumors: A comparative immunohistochemical study. Turk. J. Med. Sci. 2018, 48, 419–423. [Google Scholar] [CrossRef] [PubMed]
  73. Zhang, R.; Lian, Y.; Xie, K.; Cai, Y.; Pan, Y.; Zhu, Y. Ropivacaine suppresses tumor biological characteristics of human hepatocellular carcinoma via inhibiting IGF-1R/PI3K/AKT/mTOR signaling axis. Bioengineered 2021, 12, 9162–9173. [Google Scholar] [CrossRef]
  74. Dobre, M.; Herlea, V.; Vladut, C.; Ciocirlan, M.; Balaban, V.D.; Constantinescu, G.; Diculescu, M.; Milanesi, E. Dysregulation of miRNAs Targeting the IGF-1R Pathway in Pancreatic Ductal Adenocarcinoma. Cells 2021, 10, 1856. [Google Scholar] [CrossRef] [PubMed]
  75. Guo, L.; Bai, Y.; Ni, T.; Li, Y.; Cao, R.; Ji, S.; Li, S. MicroRNA1533p suppresses retinoblastoma cell growth and invasion via targeting the IGF1R/Raf/MEK and IGF1R/PI3K/AKT signaling pathways. Int. J. Oncol. 2021, 59, 47. [Google Scholar] [CrossRef]
  76. Yang, C.; Zhang, Y.; Segar, N.; Huang, C.; Zeng, P.; Tan, X.; Mao, L.; Chen, Z.; Haglund, F.; Larsson, O.; et al. Nuclear IGF1R interacts with NuMA and regulates 53BP1dependent DNA doublestrand break repair in colorectal cancer. Oncol. Rep. 2021, 46, 168. [Google Scholar] [CrossRef]
  77. Toretsky, J.A.; Steinberg, S.M.; Thakar, M.; Counts, D.; Pironis, B.; Parente, C.; Eskenazi, A.; Helman, L.; Wexler, L.H. Insulin-like growth factor type 1 (IGF-1) and IGF binding protein-3 in patients with Ewing sarcoma family of tumors. Cancer 2001, 92, 2941–2947. [Google Scholar] [CrossRef]
  78. Scotlandi, K.; Manara, M.C.; Serra, M.; Marino, M.T.; Ventura, S.; Garofalo, C.; Alberghini, M.; Magagnoli, G.; Ferrari, S.; Lopez-Guerrero, J.A.; et al. Expression of insulin-like growth factor system components in Ewing’s sarcoma and their association with survival. Eur. J. Cancer 2011, 47, 1258–1266. [Google Scholar] [CrossRef]
  79. Worrall, C.; Nedelcu, D.; Serly, J.; Suleymanova, N.; Oprea, I.; Girnita, A.; Girnita, L. Novel mechanisms of regulation of IGF-1R action: Functional and therapeutic implications. Pediatr. Endocrinol. Rev. 2013, 10, 473–484. [Google Scholar]
  80. de Groot, S.; Rottgering, B.; Gelderblom, H.; Pijl, H.; Szuhai, K.; Kroep, J.R. Unraveling the Resistance of IGF-Pathway Inhibition in Ewing Sarcoma. Cancers 2020, 12, 3568. [Google Scholar] [CrossRef]
  81. Heitzeneder, S.; Sotillo, E.; Shern, J.F.; Sindiri, S.; Xu, P.; Jones, R.; Pollak, M.; Noer, P.R.; Lorette, J.; Fazli, L.; et al. Pregnancy-Associated Plasma Protein-A (PAPP-A) in Ewing Sarcoma: Role in Tumor Growth and Immune Evasion. J. Natl. Cancer Inst. 2019, 111, 970–982. [Google Scholar] [CrossRef] [PubMed]
  82. Kirschner, A.; Thiede, M.; Grunewald, T.G.; Alba Rubio, R.; Richter, G.H.; Kirchner, T.; Busch, D.H.; Burdach, S.; Thiel, U. Pappalysin-1 T cell receptor transgenic allo-restricted T cells kill Ewing sarcoma in vitro and in vivo. Oncoimmunology 2017, 6, e1273301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Pardali, E.; van der Schaft, D.W.; Wiercinska, E.; Gorter, A.; Hogendoorn, P.C.; Griffioen, A.W.; ten Dijke, P. Critical role of endoglin in tumor cell plasticity of Ewing sarcoma and melanoma. Oncogene 2011, 30, 334–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hamdan, R.; Zhou, Z.; Kleinerman, E.S. Blocking SDF-1alpha/CXCR4 downregulates PDGF-B and inhibits bone marrow-derived pericyte differentiation and tumor vascular expansion in Ewing tumors. Mol. Cancer Ther. 2014, 13, 483–491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Uren, A.; Merchant, M.S.; Sun, C.J.; Vitolo, M.I.; Sun, Y.; Tsokos, M.; Illei, P.B.; Ladanyi, M.; Passaniti, A.; Mackall, C.; et al. Beta-platelet-derived growth factor receptor mediates motility and growth of Ewing’s sarcoma cells. Oncogene 2003, 22, 2334–2342. [Google Scholar] [CrossRef] [Green Version]
  86. Zwerner, J.P.; May, W.A. PDGF-C is an EWS/FLI induced transforming growth factor in Ewing family tumors. Oncogene 2001, 20, 626–633. [Google Scholar] [CrossRef] [Green Version]
  87. Koppenhafer, S.L.; Goss, K.L.; Terry, W.W.; Gordon, D.J. Inhibition of the ATR-CHK1 Pathway in Ewing Sarcoma Cells Causes DNA Damage and Apoptosis via the CDK2-Mediated Degradation of RRM2. Mol. Cancer Res. 2020, 18, 91–104. [Google Scholar] [CrossRef]
  88. Zhang, K.; Zhang, M.; Luo, Z.; Wen, Z.; Yan, X. The dichotomous role of TGF-beta in controlling liver cancer cell survival and proliferation. J. Genet. Genom. 2020, 47, 497–512. [Google Scholar] [CrossRef]
  89. Pridgeon, M.G.; Grohar, P.J.; Steensma, M.R.; Williams, B.O. Wnt Signaling in Ewing Sarcoma, Osteosarcoma, and Malignant Peripheral Nerve Sheath Tumors. Curr. Osteoporos. Rep. 2017, 15, 239–246. [Google Scholar] [CrossRef]
  90. Ross, K.A.; Smyth, N.A.; Murawski, C.D.; Kennedy, J.G. The biology of ewing sarcoma. ISRN Oncol 2013, 2013, 759725. [Google Scholar] [CrossRef] [Green Version]
  91. Burdach, S.; Plehm, S.; Unland, R.; Dirksen, U.; Borkhardt, A.; Staege, M.S.; Muller-Tidow, C.; Richter, G.H. Epigenetic maintenance of stemness and malignancy in peripheral neuroectodermal tumors by EZH2. Cell Cycle 2009, 8, 1991–1996. [Google Scholar] [CrossRef] [PubMed]
  92. Kailayangiri, S.; Altvater, B.; Lesch, S.; Balbach, S.; Gottlich, C.; Kuhnemundt, J.; Mikesch, J.H.; Schelhaas, S.; Jamitzky, S.; Meltzer, J.; et al. EZH2 Inhibition in Ewing Sarcoma Upregulates GD2 Expression for Targeting with Gene-Modified T Cells. Mol. Ther. 2019, 27, 933–946. [Google Scholar] [CrossRef] [PubMed]
  93. Ahmed, A.A.; Abedalthagafi, M.; Anwar, A.E.; Bui, M.M. Akt and Hippo Pathways in Ewing’s Sarcoma Tumors and Their Prognostic Significance. J. Cancer 2015, 6, 1005–1010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hsu, J.H.; Lawlor, E.R. BMI-1 suppresses contact inhibition and stabilizes YAP in Ewing sarcoma. Oncogene 2011, 30, 2077–2085. [Google Scholar] [CrossRef] [Green Version]
  95. Tanner, J.M.; Bensard, C.; Wei, P.; Krah, N.M.; Schell, J.C.; Gardiner, J.; Schiffman, J.; Lessnick, S.L.; Rutter, J. EWS/FLI is a Master Regulator of Metabolic Reprogramming in Ewing Sarcoma. Mol. Cancer Res. 2017, 15, 1517–1530. [Google Scholar] [CrossRef] [Green Version]
  96. Sen, N.; Cross, A.M.; Lorenzi, P.L.; Khan, J.; Gryder, B.E.; Kim, S.; Caplen, N.J. EWS-FLI1 reprograms the metabolism of Ewing sarcoma cells via positive regulation of glutamine import and serine-glycine biosynthesis. Mol. Carcinog. 2018, 57, 1342–1357. [Google Scholar] [CrossRef]
  97. Issaq, S.H.; Mendoza, A.; Kidner, R.; Rosales, T.I.; Duveau, D.Y.; Heske, C.M.; Rohde, J.M.; Boxer, M.B.; Thomas, C.J.; DeBerardinis, R.J.; et al. EWS-FLI1-regulated Serine Synthesis and Exogenous Serine are Necessary for Ewing Sarcoma Cellular Proliferation and Tumor Growth. Mol. Cancer Ther. 2020, 19, 1520–1529. [Google Scholar] [CrossRef]
  98. Sullivan, M.R.; Mattaini, K.R.; Dennstedt, E.A.; Nguyen, A.A.; Sivanand, S.; Reilly, M.F.; Meeth, K.; Muir, A.; Darnell, A.M.; Bosenberg, M.W.; et al. Increased Serine Synthesis Provides an Advantage for Tumors Arising in Tissues Where Serine Levels Are Limiting. Cell Metab. 2019, 29, 1410–1421.e4. [Google Scholar] [CrossRef]
  99. Sanchez-Sanchez, A.M.; Antolin, I.; Puente-Moncada, N.; Suarez, S.; Gomez-Lobo, M.; Rodriguez, C.; Martin, V. Melatonin Cytotoxicity Is Associated to Warburg Effect Inhibition in Ewing Sarcoma Cells. PLoS ONE 2015, 10, e0135420. [Google Scholar] [CrossRef]
  100. Nie, X.; Wang, H.; Wei, X.; Li, L.; Xue, T.; Fan, L.; Ma, H.; Xia, Y.; Wang, Y.D.; Chen, W.D. LRP5 promotes gastric cancer via activating canonical Wnt/beta-catenin and glycolysis pathways. Am. J. Pathol. 2021, 192, 503–517. [Google Scholar] [CrossRef]
  101. Cortese, N.; Capretti, G.; Barbagallo, M.; Rigamonti, A.; Takis, P.G.; Castino, G.F.; Vignali, D.; Maggi, G.; Gavazzi, F.; Ridolfi, C.; et al. Metabolome of Pancreatic Juice Delineates Distinct Clinical Profiles of Pancreatic Cancer and Reveals a Link between Glucose Metabolism and PD-1(+) Cells. Cancer Immunol. Res. 2020, 8, 493–505. [Google Scholar] [CrossRef] [PubMed]
  102. van der Schaft, D.W.; Hillen, F.; Pauwels, P.; Kirschmann, D.A.; Castermans, K.; Egbrink, M.G.; Tran, M.G.; Sciot, R.; Hauben, E.; Hogendoorn, P.C.; et al. Tumor cell plasticity in Ewing sarcoma, an alternative circulatory system stimulated by hypoxia. Cancer Res. 2005, 65, 11520–11528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Dasgupta, A.; Trucco, M.; Rainusso, N.; Bernardi, R.J.; Shuck, R.; Kurenbekova, L.; Loeb, D.M.; Yustein, J.T. Metabolic modulation of Ewing sarcoma cells inhibits tumor growth and stem cell properties. Oncotarget 2017, 8, 77292–77308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Chen, Z.; Wang, X.; Wang, G.; Xiao, B.; Ma, Z.; Huo, H.; Li, W. A seven-lncRNA signature for predicting Ewing’s sarcoma. PeerJ 2021, 9, e11599. [Google Scholar] [CrossRef]
  105. Ma, L.; Sun, X.; Kuai, W.; Hu, J.; Yuan, Y.; Feng, W.; Lu, X. LncRNA SOX2 overlapping transcript acts as a miRNA sponge to promote the proliferation and invasion of Ewing’s sarcoma. Am. J. Transl. Res. 2019, 11, 3841–3849. [Google Scholar]
  106. Li, H.; Huang, F.; Liu, X.Q.; Liu, H.C.; Dai, M.; Zeng, J. LncRNA TUG1 promotes Ewing’s sarcoma cell proliferation, migration, and invasion via the miR-199a-3p-MSI2 signaling pathway. Neoplasma 2021, 68, 590–601. [Google Scholar] [CrossRef]
  107. Goss, K.L.; Koppenhafer, S.L.; Harmoney, K.M.; Terry, W.W.; Gordon, D.J. Inhibition of CHK1 sensitizes Ewing sarcoma cells to the ribonucleotide reductase inhibitor gemcitabine. Oncotarget 2017, 8, 87016–87032. [Google Scholar] [CrossRef] [Green Version]
  108. 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]
  109. Lowery, C.D.; Dowless, M.; Renschler, M.; Blosser, W.; VanWye, A.B.; Stephens, J.R.; Iversen, P.W.; Lin, A.B.; Beckmann, R.P.; Krytska, K.; et al. Broad Spectrum Activity of the Checkpoint Kinase 1 Inhibitor Prexasertib as a Single Agent or Chemopotentiator across a Range of Preclinical Pediatric Tumor Models. Clin. Cancer Res. 2019, 25, 2278–2289. [Google Scholar] [CrossRef]
  110. Henssen, A.G.; Reed, C.; Jiang, E.; Garcia, H.D.; von Stebut, J.; MacArthur, I.C.; Hundsdoerfer, P.; Kim, J.H.; de Stanchina, E.; Kuwahara, Y.; et al. Therapeutic targeting of PGBD5-induced DNA repair dependency in pediatric solid tumors. Sci. Transl. Med. 2017, 9, eaam9078. [Google Scholar] [CrossRef] [Green Version]
  111. Koppenhafer, S.L.; Goss, K.L.; Terry, W.W.; Gordon, D.J. mTORC1/2 and Protein Translation Regulate Levels of CHK1 and the Sensitivity to CHK1 Inhibitors in Ewing Sarcoma Cells. Mol. Cancer Ther. 2018, 17, 2676–2688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Saldivar, J.C.; Cortez, D.; Cimprich, K.A. The essential kinase ATR: Ensuring faithful duplication of a challenging genome. Nat. Rev. Mol. Cell Biol. 2017, 18, 622–636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Saldivar, J.C.; Cimprich, K.A. A new mitotic activity comes into focus. Science 2018, 359, 30–31. [Google Scholar] [CrossRef] [PubMed]
  114. Kabeche, L.; Nguyen, H.D.; Buisson, R.; Zou, L. A mitosis-specific and R loop-driven ATR pathway promotes faithful chromosome segregation. Science 2018, 359, 108–114. [Google Scholar] [CrossRef] [Green Version]
  115. Wang, Z.; Wang, F.; Tang, T.; Guo, C. The role of PARP1 in the DNA damage response and its application in tumor therapy. Front. Med. 2012, 6, 156–164. [Google Scholar] [CrossRef]
  116. Vormoor, B.; Curtin, N.J. Poly(ADP-ribose) polymerase inhibitors in Ewing sarcoma. Curr. Opin. Oncol. 2014, 26, 428–433. [Google Scholar] [CrossRef]
  117. Stewart, E.; Goshorn, R.; Bradley, C.; Griffiths, L.M.; Benavente, C.; Twarog, N.R.; Miller, G.M.; Caufield, W.; Freeman, B.B., 3rd; Bahrami, A.; et al. Targeting the DNA repair pathway in Ewing sarcoma. Cell Rep. 2014, 9, 829–841. [Google Scholar] [CrossRef] [Green Version]
  118. Vormoor, B.; Schlosser, Y.T.; Blair, H.; Sharma, A.; Wilkinson, S.; Newell, D.R.; Curtin, N. Sensitizing Ewing sarcoma to chemo- and radiotherapy by inhibition of the DNA-repair enzymes DNA protein kinase (DNA-PK) and poly-ADP-ribose polymerase (PARP) 1/2. Oncotarget 2017, 8, 113418–113430. [Google Scholar] [CrossRef] [Green Version]
  119. Garnett, M.J.; Edelman, E.J.; Heidorn, S.J.; Greenman, C.D.; Dastur, A.; Lau, K.W.; Greninger, P.; Thompson, I.R.; Luo, X.; Soares, J.; et al. Systematic identification of genomic markers of drug sensitivity in cancer cells. Nature 2012, 483, 570–575. [Google Scholar] [CrossRef] [Green Version]
  120. Engert, F.; Schneider, C.; Weibeta, L.M.; Probst, M.; Fulda, S. PARP Inhibitors Sensitize Ewing Sarcoma Cells to Temozolomide-Induced Apoptosis via the Mitochondrial Pathway. Mol. Cancer Ther. 2015, 14, 2818–2830. [Google Scholar] [CrossRef] [Green Version]
  121. Gill, S.J.; Travers, J.; Pshenichnaya, I.; Kogera, F.A.; Barthorpe, S.; Mironenko, T.; Richardson, L.; Benes, C.H.; Stratton, M.R.; McDermott, U.; et al. Combinations of PARP Inhibitors with Temozolomide Drive PARP1 Trapping and Apoptosis in Ewing’s Sarcoma. PLoS ONE 2015, 10, e0140988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Smith, M.A.; Hampton, O.A.; Reynolds, C.P.; Kang, M.H.; Maris, J.M.; Gorlick, R.; Kolb, E.A.; Lock, R.; Carol, H.; Keir, S.T.; et al. Initial testing (stage 1) of the PARP inhibitor BMN 673 by the pediatric preclinical testing program: PALB2 mutation predicts exceptional in vivo response to BMN 673. Pediatr. Blood Cancer 2015, 62, 91–98. [Google Scholar] [CrossRef] [Green Version]
  123. Heske, C.M.; Davis, M.I.; Baumgart, J.T.; Wilson, K.; Gormally, M.V.; Chen, L.; Zhang, X.; Ceribelli, M.; Duveau, D.Y.; Guha, R.; et al. Matrix Screen Identifies Synergistic Combination of PARP Inhibitors and Nicotinamide Phosphoribosyltransferase (NAMPT) Inhibitors in Ewing Sarcoma. Clin. Cancer Res. 2017, 23, 7301–7311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Bademci, G.; Abad, C.; Cengiz, F.B.; Seyhan, S.; Incesulu, A.; Guo, S.; Fitoz, S.; Atli, E.I.; Gosstola, N.C.; Demir, S.; et al. Long-range cis-regulatory elements controlling GDF6 expression are essential for ear development. J. Clin. Investig. 2020, 130, 4213–4217. [Google Scholar] [CrossRef] [PubMed]
  125. Williams, L.A.; Bhargav, D.; Diwan, A.D. Unveiling the bmp13 enigma: Redundant morphogen or crucial regulator? Int. J. Biol. Sci. 2008, 4, 318–329. [Google Scholar] [CrossRef] [PubMed]
  126. 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. Primers 2018, 4, 5. [Google Scholar] [CrossRef]
  127. Sonnemann, J.; Dreyer, L.; Hartwig, M.; Palani, C.D.; Hong, L.T.T.; Klier, U.; Broker, B.; Volker, 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]
  128. Lu, G.; Punj, V.; Chaudhary, P.M. Proteasome inhibitor Bortezomib induces cell cycle arrest and apoptosis in cell lines derived from Ewing’s sarcoma family of tumors and synergizes with TRAIL. Cancer Biol. Ther. 2008, 7, 603–608. [Google Scholar] [CrossRef] [Green Version]
  129. Lohberger, B.; Bernhart, E.; Stuendl, N.; Glaenzer, D.; Leithner, A.; Rinner, B.; Bauer, R.; Kretschmer, N. Periplocin mediates TRAIL-induced apoptosis and cell cycle arrest in human myxofibrosarcoma cells via the ERK/p38/JNK pathway. Phytomedicine 2020, 76, 153262. [Google Scholar] [CrossRef]
  130. Du, J.; Wang, Y.; Chen, D.; Ji, G.; Ma, Q.; Liao, S.; Zheng, Y.; Zhang, J.; Hou, Y. BAY61-3606 potentiates the anti-tumor effects of TRAIL against colon cancer through up-regulating DR4 and down-regulating NF-kappaB. Cancer Lett. 2016, 383, 145–153. [Google Scholar] [CrossRef]
  131. Hanikoglu, F.; Cort, A.; Ozben, H.; Hanikoglu, A.; Ozben, T. Epoxomicin Sensitizes Resistant Osteosarcoma Cells to TRAIL Induced Apoptosis. Anticancer Agents Med. Chem. 2015, 15, 527–533. [Google Scholar] [CrossRef] [PubMed]
  132. Subbiah, V.; Brown, R.E.; Buryanek, J.; Trent, J.; Ashkenazi, A.; Herbst, R.; Kurzrock, R. Targeting the apoptotic pathway in chondrosarcoma using recombinant human Apo2L/TRAIL (dulanermin), a dual proapoptotic receptor (DR4/DR5) agonist. Mol. Cancer Ther. 2012, 11, 2541–2546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Plummer, R.; Attard, G.; Pacey, S.; Li, L.; Razak, A.; Perrett, R.; Barrett, M.; Judson, I.; Kaye, S.; Fox, N.L.; et al. Phase 1 and pharmacokinetic study of lexatumumab in patients with advanced cancers. Clin. Cancer Res. 2007, 13, 6187–6194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hennessy, M.; Wahba, A.; Felix, K.; Cabrera, M.; Segura, M.G.; Kundra, V.; Ravoori, M.K.; Stewart, J.; Kleinerman, E.S.; Jensen, V.B.; et al. Bempegaldesleukin (BEMPEG; NKTR-214) efficacy as a single agent and in combination with checkpoint-inhibitor therapy in mouse models of osteosarcoma. Int. J. Cancer 2021, 148, 1928–1937. [Google Scholar] [CrossRef]
  135. Gamie, Z.; Kapriniotis, K.; Papanikolaou, D.; Haagensen, E.; Da Conceicao Ribeiro, R.; Dalgarno, K.; Krippner-Heidenreich, A.; Gerrand, C.; Tsiridis, E.; Rankin, K.S. TNF-related apoptosis-inducing ligand (TRAIL) for bone sarcoma treatment: Pre-clinical and clinical data. Cancer Lett. 2017, 409, 66–80. [Google Scholar] [CrossRef]
  136. Henrich, I.C.; Young, R.; Quick, L.; Oliveira, A.M.; Chou, M.M. USP6 Confers Sensitivity to IFN-Mediated Apoptosis through Modulation of TRAIL Signaling in Ewing Sarcoma. Mol. Cancer Res. 2018, 16, 1834–1843. [Google Scholar] [CrossRef] [Green Version]
  137. Robles, A.J.; Dai, W.; Haldar, S.; Ma, H.; Anderson, V.M.; Overacker, R.D.; Risinger, A.L.; Loesgen, S.; Houghton, P.J.; Cichewicz, R.H.; et al. Altertoxin II, a Highly Effective and Specific Compound against Ewing Sarcoma. Cancers 2021, 13, 6176. [Google Scholar] [CrossRef]
  138. Kerschner-Morales, S.L.; Kuhne, M.; Becker, S.; Beck, J.F.; Sonnemann, J. Anticancer effects of the PLK4 inhibitors CFI-400945 and centrinone in Ewing’s sarcoma cells. J. Cancer Res. Clin. Oncol. 2020, 146, 2871–2883. [Google Scholar] [CrossRef]
  139. Ma, Y.; Baltezor, M.; Rajewski, L.; Crow, J.; Samuel, G.; Staggs, V.S.; Chastain, K.M.; Toretsky, J.A.; Weir, S.J.; Godwin, A.K. Targeted inhibition of histone deacetylase leads to suppression of Ewing sarcoma tumor growth through an unappreciated EWS-FLI1/HDAC3/HSP90 signaling axis. J. Mol. Med. (Berl.) 2019, 97, 957–972. [Google Scholar] [CrossRef]
  140. Flores, G.; Everett, J.H.; Boguslawski, E.A.; Oswald, B.M.; Madaj, Z.B.; Beddows, I.; Dikalov, S.; Adams, M.; Klumpp-Thomas, C.A.; Kitchen-Goosen, S.M.; et al. CDK9 Blockade Exploits Context-dependent Transcriptional Changes to Improve Activity and Limit Toxicity of Mithramycin for Ewing Sarcoma. Mol. Cancer Ther. 2020, 19, 1183–1196. [Google Scholar] [CrossRef] [Green Version]
  141. Wang, S.; Hwang, E.E.; Guha, R.; O’Neill, A.F.; Melong, N.; Veinotte, C.J.; Conway Saur, A.; 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]
Figure 1. Schematic illustration of cancer pathways under control of ES’s fusion proteins: The figure summarizes our review of research active molecules over the past 10 years. The various domain of ES-produced fusion oncoproteins required for the activation of gene expression, and their products, such as RNA or proteins regulate cellular proliferation, apoptosis and migration, for example. All together with induced cell signaling cells gain oncogenic traits and transformation. Small circles are active molecules, large circles are cellular processes (see the hallmark of cancer). Colors in the figure are correlated to organizational divisions in the review for ease of perception. There are three colored clusters: purple (Section 3.1. Targeting of ES Pressure on Adhesion, Migration, and Invasion), blue (Section 3.2; Targeting of Ewing Sarcoma Cells with a Focus on Proliferation, Cell Differentiation, and Cell Survival) and red (Section 3.3; Targeting of ES: Induction of Apoptosis and Cell Cycle Arrest).
Figure 1. Schematic illustration of cancer pathways under control of ES’s fusion proteins: The figure summarizes our review of research active molecules over the past 10 years. The various domain of ES-produced fusion oncoproteins required for the activation of gene expression, and their products, such as RNA or proteins regulate cellular proliferation, apoptosis and migration, for example. All together with induced cell signaling cells gain oncogenic traits and transformation. Small circles are active molecules, large circles are cellular processes (see the hallmark of cancer). Colors in the figure are correlated to organizational divisions in the review for ease of perception. There are three colored clusters: purple (Section 3.1. Targeting of ES Pressure on Adhesion, Migration, and Invasion), blue (Section 3.2; Targeting of Ewing Sarcoma Cells with a Focus on Proliferation, Cell Differentiation, and Cell Survival) and red (Section 3.3; Targeting of ES: Induction of Apoptosis and Cell Cycle Arrest).
Cancers 14 01988 g001
Table 1. Targeting molecules in ES pathogenesis.
Table 1. Targeting molecules in ES pathogenesis.
Targetable MoleculesMain PathwaysTumor Effects
CD99IGF-1R and RAS-Rac1 signalingInduces caspase-independent cell death, endocytosis, cell aggregation, micropinocytosis, cell adhesion, migration, invasion, metastasis, differentiation
GDF6GDF6 prodomain signaling pathwayCell proliferation, tumor growth, differentiation, apoptosis
E-cadherinMAPK PathwayAnchorage-independent growth and spheroid formation, cell-cell adhesion
EndoglinTGFβ signalingTumor cell plasticity, patient survival, invasion, anchorage-independent growth, progression of aggressive tumors
EZH2EpigeneticCell differentiation, phenotypic heterogeneity, self-renewal
GLI1Sonic Hedgehog (SHH) pathwayCell proliferation, cell cycle control, apoptosis, cell viability, metastasis, invasion, migration, clonogenicity
PDGF family membersPDGF pathwaySelf-renewal, invasion, chemotherapy resistance, primary tumor growth, metastasis, drug resistance, poor clinical outcome
ROCK2RhoA-ROCK pathwayMigration, invasion, proliferation, clonogenic capacity, tumor growth
YAP proteinsYAP/TAZ pathway, Hippo signaling, WNT/β-catenin signalingMigration, cell proliferation, metastasis, anchorage-independent colony formation
PAPP-AIGF signalingCell proliferation, migration, cell survival, tumor growth, invasion, metastasis
PARP familyDNA repair, replicationApoptosis
TRAILTRAIL-pathwayInduces caspase-independent cell death, apoptosis
ATR/CHK1ATR-CHK1 pathwayCell cycle regulation, cell cycle arrest
LDHaerobic glycolysisCell proliferation, apoptosis, tumor growth, cell survival
PHGDHSerine synthesisCell proliferation
lncRNA SOX2WNT/β-catenin signalingCell proliferation, invasion, apoptosis, tumor growth
lncRNA TUG1TUG-miR-145-5p-TRPC6 pathwayCell proliferation, migration, invasion
Table 2. Clinical trials.
Table 2. Clinical trials.
S.N.NumberNumber of PatientsDiseaseDrug/TargetResults
1NCT0412915118Ewing Sarcoma
Recurrent
Palbociclib/CDK4 and CDK6
Ganitumab/IGF-1R
Active
2NCT0254654416Relapsed Ewing Sarcoma
Refractory Ewing Sarcoma
Linsitinib/IGF-1RDisease progression, limited therapeutical effect
3NCT0094932524Soft Tissue and Bone SarcomaTemsirolimus/mTOR
Doxorubicin/topoisomerase II
The response rate was 53%, found a correlation between inhibition of mTOR and therapeutical effect 10.1186/s13569-018-0107-9
4NCT00987636907Ewing sarcomaZoledronic acid/osteoclast apoptosis
Busulfan/guanine N7
Treosulfan/guanine N7
Melphalan/guanine N7
BuMel treatment was more successful than standard chemotherapy -vincristine, dactinomycin, and ifosfamide (VAI)
5NCT0061881335Ewing SarcomaRadiation therapy
therapeutic conventional surgery
etoposide/topoisomerase II
ifosfamide/DNA
doxorubicin hydrochloride/topoisomerase II
cyclophosphamide/guanine N7
vincristine sulfate/tubulin
topotecan hydrochloride/topoisomerase I
filgrastim/Granulocyte
No incidence of death was recorded in 37 weeks of treatment
6NCT005162957Ewing Sarcoma of Bone
Extraosseous Ewing Sarcoma
Peripheral Primitive Neuroectodermal Tumor
Recurrent Ewing Sarcoma/Peripheral Primitive Neuroectodermal Tumor
Topotecan hydrochloride/topoisomerase I
cyclophosphamide/guanine N7
vincristine sulfate/tubulin
bevacizumab/VEGF-A
Days of event free survival—442
7NCT0047027510Recurrent or Refractory Ewing SarcomaCytarabine/DNALack of efficacy
8NCT0265700545Relapsed or Refractory Ewing SarcomaTK216/EWS-FLI1Active
9NCT0006189338Ewing Sarcoma Family of TumorsRadiation therapy
conventional surgery
etoposide/topoisomerase II
ifosfamide/DNA
doxorubicin hydrochloride/topoisomerase II
cyclophosphamide/guanine N7
vincristine sulfate/tubulin
topotecan hydrochloride/topoisomerase I
filgrastim/granulocyte
vinblastine sulfate/tubulin
MESNA/urotoxic metabolites
24-month event free survival was 35%: 71% for the seven with isolated pulmonary metastases, 26% for all others.
10NCT0251113222Ewing SarcomaVigil/TGF-β
Temozolomide/guanine
Irinotecan/topoisomerase I
1 case report of complete response to therapy
11NCT0158354312Recurrent/Metastatic Ewing’s SarcomaOlaparib/PARPNo significant responses or durable disease control was seen
12NCT0133113518Ewing sarcoma,
osteosarcoma,
malignant
peripheral nerve
sheath tumor,
rhabdoid tumor,
retinoblastoma
Sirolimus/mTORThe combination of sirolimus with metronomic chemotherapy is well tolerated in children. A phase II trial of this combination is ongoing.
13NCT0042827224Ewing Sarcoma
Osteosarcoma
Neuroblastoma
Rhabdomyosarcoma
Lexatumumab/TRAIL-2RThe drug seems to mediate some clinical activity in pediatric solid tumors and may work with radiation to enhance antitumor effects.
14NCT02306161312Metastatic Ewing Sarcoma
Metastatic Malignant Neoplasm in the Bone
Metastatic Malignant Neoplasm in the Bone Marrow
Metastatic Malignant Neoplasm in the Lung
Metastatic Peripheral Primitive Neuroectodermal Tumor of Bone
Peripheral Primitive Neuroectodermal Tumor of Soft Tissues
Cyclophosphamide/guanine N7
Doxorubicin/topoisomerase II
Etoposide/topoisomerase II
Ganitumab/IGevent-freeF-1R
Ifosfamide/DNA
Vincristine/tubulin
Active
15NCT0406711545Ewing SarcomaTrabectedin/guanine N2
Irinotecan/topoisomerase I
Recruiting
16NCT0007010950Rhabdomyosarcoma
Recurrent Childhood Rhabdomyosarcoma
Recurrent Childhood Soft Tissue Sarcoma
Recurrent Ewing Sarcoma
Peripheral Primitive Neuroectodermal Tumor
Trabectedin/guanine N2
17NCT0360064950Ewing Sarcoma
Myxoid Liposarcoma
Sarcoma, Soft Tissue
Desmoplastic Small Round Cell Tumor
Extraskeletal Myxoid Chondrosarcoma
Angiomatoid Fibrous Histiocytoma
Clear Cell Sarcoma
Primary Pulmonary Myxoid Sarcoma
Myoepithelial Tumor
Sclerosing Epithelioid Fibrosarcoma
Fibromyxoid Tumor
Cyclophosphamide/guanine N7
Topotecan/topoisomerase I
Seclidemstat/LSD1
Recruiting
18NCT0349137156Osteosarcoma
Ewing sarcoma
Chondrosarcoma
Soft tissue sarcoma
Methylsulfonic apatinib/VEGFR-2No data
19NCT0469072529Osteosarcoma
Ewing sarcoma
Chondrosarcoma
TQB3525/PI 3-kinasesActive
20NCT016105708Ewing Sarcoma
Sarcoma
Mithramycin/EWS-FLI1The trial was closed to enrollment, due to inability to safely achieve the desired mithramycin exposure
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fayzullina, D.; Tsibulnikov, S.; Stempen, M.; Schroeder, B.A.; Kumar, N.; Kharwar, R.K.; Acharya, A.; Timashev, P.; Ulasov, I. Novel Targeted Therapeutic Strategies for Ewing Sarcoma. Cancers 2022, 14, 1988. https://doi.org/10.3390/cancers14081988

AMA Style

Fayzullina D, Tsibulnikov S, Stempen M, Schroeder BA, Kumar N, Kharwar RK, Acharya A, Timashev P, Ulasov I. Novel Targeted Therapeutic Strategies for Ewing Sarcoma. Cancers. 2022; 14(8):1988. https://doi.org/10.3390/cancers14081988

Chicago/Turabian Style

Fayzullina, Daria, Sergey Tsibulnikov, Mikhail Stempen, Brett A. Schroeder, Naveen Kumar, Rajesh Kumar Kharwar, Arbind Acharya, Peter Timashev, and Ilya Ulasov. 2022. "Novel Targeted Therapeutic Strategies for Ewing Sarcoma" Cancers 14, no. 8: 1988. https://doi.org/10.3390/cancers14081988

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