Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2025 May 16;122(20):e2221823122. doi: 10.1073/pnas.2221823122

Phase separation of the oncogenic fusion protein EWS::FLI1 is modulated by its DNA-binding domain

Emily E Selig a,b, Erich J Sohn a,b, Aiola Stoja a,c, Alma K Moreno-Romero a,b, Shivani Akula a,b, Xiaoping Xu a,b, Alexander J R Bishop a,c, David S Libich a,b,1
PMCID: PMC12107149  PMID: 40377985

Significance

The oncogenic fusion protein, EWS::FLI1, responsible for more than 85% of Ewing sarcoma tumors combines the transactivation domain from EWS and the DNA-binding domain (DBD) from Friend leukemia integration 1 (FLI1). The fusion impacts the function of wild-type EWS and drives oncogenesis via aberrant transcriptional and splicing changes as well as defects in the DNA-damage response. Both components of the fusion are required for oncogenesis suggesting a synergistic function between the domains. Here, the authors describe the structural underpinnings of EWS::FLI1’s effect on EWS, mediated by the FLI1 DBD, that enhances the phase separation propensity of EWS and drives aberrant changes to the physical properties of condensates.

Keywords: RNA-binding protein, intrinsically disordered protein, NMR, EWS, EWS::FLI1 fusion

Abstract

Ewing sarcoma (EwS) is an aggressive cancer of bone and soft tissue that predominantly affects children and young adults. A chromosomal translocation joins the low-complexity domain (LCD) of the RNA-binding protein EWS (EWSLCD) with the DNA-binding domain of Friend leukemia integration 1 (FLI1DBD), creating EWS::FLI1, a potent fusion oncoprotein essential for EwS development and responsible for over 85% of EwS tumors. EWS::FLI1 forms biomolecular condensates in vivo and promotes tumorigenesis through mediation of aberrant transcriptional changes and by interfering with the normal functions of nucleic acid-binding proteins like EWS through a dominant-negative mechanism. In particular, the expression of EWS::FLI1 in EwS directly interferes with the biological functions of EWS leading to alternate splicing events and defects in DNA-damage repair pathways. Though the EWSLCD is capable of phase separation, here we report a direct interaction between FLI1DBD and EWSLCD that enhances condensate formation and alters the physical properties of the condensate. This effect was conserved for three related E-twenty-six transformation-specific (ETS) DNA-binding domains (DBDs) while DNA binding blocked the interaction with EWSLCD and inhibited EWS::FLI1 condensate formation. NMR spectroscopy and mutagenesis studies confirmed that ETS DBDs transiently interact with EWSLCD via the ETS DBDs “wings.” Together these results revealed that ETS DBDs, particularly FLI1DBD, enhance EWSLCD condensate formation and rigidity, supporting a model in which electrostatic and structural interactions drive condensate dynamics with implications for EWS::FLI1-mediated transcriptional regulation in EwS.


The oncogenic EWS::FLI1 fusion protein is the hallmark of a group of fusion proteins associated with Ewing sarcoma (EwS), a highly aggressive pediatric bone and soft tissue cancer (1). This fusion arises from a chromosomal translocation that joins the N-terminal low-complexity domain (LCD) of the RNA-binding protein EWS (EWSLCD) in-frame with the DNA-binding domain (DBD) of the E-twenty-six transformation-specific (ETS) family transcription factor Friend leukemia integration 1 (FLI1DBD). The resulting EWS::FLI1 fusion is responsible for approximately 85% of EwS tumors (1). Of the 28 ETS transcription factors, FLI1, transcriptional regulator ERG (ERG), and protein FEV (FEV) are the most common EWSLCD fusion partners identified in EwS; however, the reasons for this frequency remain unclear (2). More broadly, in the family of EwS-related primitive neuroectodermal tumors, the FET family of RNA-binding proteins, including fused-in-sarcoma (FUS), EWS, and TATA-binding protein associated factor 2 N (TAF15), are commonly found fused with the DBDs of ETS-family and other transcription factors (3).

The EWS::FLI1 fusion disrupts normal genetic programs by aberrantly binding to GGAA microsatellites and FLI1 consensus sites via the FLI1 DBD. Concurrently, the EWSLCD facilitates the recruitment of chromatin remodelers, epigenetic modifiers, and transcriptional machinery (48). Specifically, the EWSLCD plays a key role in DNA enhancer binding, and both domains are essential for recruiting ATP-dependent BRG1/BRM associated factor (BAF) chromatin remodeling complex and driving transcriptional activation (4, 914). However, these functions alone do not fully explain EWS::FLI1’s oncogenic role. Growing evidence suggests that it also exerts a dominant negative effect on the normal roles of EWS in transcriptional regulation and splicing (1518). Notably, recent findings indicate that the presence of EWS::FLI1 at transcriptionally active sites interferes with the release of Breast Cancer Type 1 Susceptibility Protein (BRCA1) from DNA-directed RNA polymerase II subunit RPB1 (RNA Pol II). This disruption leads to increased transcriptional stress, homologous recombination deficiency, and the accumulation of unresolved R-loops (15).

Biomolecular condensation, in which proteins and nucleic acids phase separate from the aqueous environment to form membraneless organelles, has emerged as a key mechanism of subcellular organization (1921). Condensates play essential roles in various cellular processes, including transcription (22, 23), splicing (24), DNA damage repair (25), and the stress response (26). The EWSLCD is intrinsically disordered, lacking stable secondary and tertiary structure, and is characterized by a repetitive, degenerate low complexity SYGQP-rich sequence, which promotes self-association and condensate formation (5, 2729). In cells, FET oncogenic fusions, including EWS::FLI1, form dynamic yet specific assemblies with other intrinsically disordered proteins/regions including EWS itself (28, 30), the C-terminal domain (CTD) of RNA Pol II (16, 31, 32), Bromodomain-containing protein 4, and components of the BAF complex (4) among others. Due to the physical features of the EWSLCD, these interactions are multivalent, dynamic, and transient, making high-resolution structural analysis of the complexes challenging (29, 33). Additionally, EWS::FLI1 undergoes phase separation and is thought to require self-association to stabilize its DNA binding and promote transcriptional activation. However, the exact nature of these interactions remains unresolved (4, 5, 28).

The oncogenic function of EWS::FLI1 requires features derived from both the EWSLCD and the FLI1DBD (4). To investigate their interplay in self-association and condensate formation, we used biophysical approaches including turbidity and thioflavin T (ThT) fluorescence assays, microscopy, and NMR to assess how EWS::FLI1 and the isolated FLI1DBD affect EWSLCD condensates. We investigated the rigidification of EWSLCD condensates by FLI1 when present as either a separate domain or in the oncogenic fusion, EWS::FLI1. We then sought to determine whether direct interactions between ETS DBDs and EWSLCD are responsible for the altered condensate properties. This study provides mechanistic insight into EWS::FLI1’s dominant negative regulation of wild-type EWS and supports the Goldilocks hypothesis of EWS::FLI1 toxicity (34). Furthermore, these studies reveal disease-associated regulatory mechanisms of condensate dynamics, uncovering the ETS DBDs as potent modulators of condensate liquidity through the possible promotion of cross-β structure formation.

Results

EWS::FLI1 Forms Rigidified Condensates.

We generated phase diagrams for full-length EWS and EWS::FLI1 (Fig. 1A and SI Appendix, Fig. S1) using turbidity measurements. Consistent with previous findings for FUS, EWS condensate formation was inhibited at high NaCl concentrations, suggesting that electrostatic interactions between N- and C-terminal residues play a role (35, 36) (Fig. 1B). In contrast, high NaCl concentrations promoted condensate formation of EWS::FLI1 (Fig. 1B), suggesting that the hydrophobic interactions that drive EWSLCD condensate formation also significantly contribute in EWS::FLI1 (33, 37, 38). In vivo, EWS and EWS::FLI1 function in multicomponent biomolecular condensates formed by multivalent LCD–LCD interactions with each other as well as with other RNA/DNA binding proteins with similar LCDs, such as the BAF complex components (6, 29, 30). To determine whether multicomponent biomolecular condensates comprising EWS wild-type and EWS::FLI1 have similar material properties, we prepared condensates by mixing 200 µM EWSLCD (EWS 1 to 264, SI Appendix, Fig. S1), which is common to both EWS and EWS::FLI1, with full-length EWS, EWS::FLI1, or EWSLCD (Fig. 1C). Full-length EWS and EWS::FLI1 readily form dual-component condensates with EWSLCD (Fig. 1G), like in vivo observations in the EwS cell line A673 (39). We assessed condensate dynamics by monitoring fluorescence recovery after photobleaching (FRAP) of the fluorescently labeled molecules within the condensates (Fig. 1E). Condensates containing full-length EWS + EWSLCD showed rapid FRAP, while condensates containing full-length EWS::FLI1 + EWSLCD recovered more slowly (Fig. 1E), indicating that condensates containing EWS::FLI1 are less liquid-like and more rigid. To determine whether this observation is recapitulated in the complex cellular environment, we used a system that fuses the 52-heptad (YSPTSPS) repeat CTD of RNA Pol II (RNAPIICTD52) to the lac repressor (LacI) and a fluorescent reporter (CFP) (40) (Fig. 1D). Expression of this CFP-LacI-RNAPIICTD52 construct in U2OS cells modified with an array of lac operons (~50,000 copies) stably inserted into chromosome 1 creates a dense foci of RNAPIICTD52, akin to intracellular condensates, capable of recruiting biomolecular constituents including EWS and EWS::FLI1 (Fig. 1 D and H). These cellular puncta recover fluorescence rapidly when bleached and overexpression of EWS did not reduce their liquidity (Fig. 1F). Similar to our in vitro experiments, overexpression and recruitment of EWS::FLI1 to the RNAPIICTD52 foci (Fig. 1H) resulted in reduced FRAP (Fig. 1F), suggesting the condensate rigidification by EWS::FLI1 also affects multicomponent condensates.

Fig. 1.

Fig. 1.

Biomolecular condensation of EWS and EWS::FLI1. (A) Cartoons of EWS, FLI1, and EWS::FLI1 domains, the EWS, and FLI1 breakpoints are marked with black triangles. (B) Phase diagrams for EWS and EWS::FLI1 measured via turbidity. Experimental design for (C) in vitro and (D) in cellulo FRAP. FRAP (E and F) and micrographs (G and H) of condensates containing (E and G) 200 µM EWSLCD plus 9 µM EWS488 (Left), 5 µM EWS::FLI1488 (Center), or 25 µM EWSLCD,488 (Right), or (F and H) CFP-LacI-RNAPIICTD52 foci in U2OS cells with wild-type EWS549 (Left), EWS::FLI1549 (Center), or only CFP-LacI-RNAPIICTD52 (control, Right). Superscripts 488 or 549 indicate fluorescent labeling, (Scale bars, 25 µm.) FRAP data are means of triplicates ± SEM.

EWS::FLI1 and FLI1DBD Alter Condensate Formation of EWSLCD.

Next we tested the effect of EWS::FLI1 and FLI1DBD on the phase separation propensity of EWSLCD using turbidity measurements and microscopy (Fig. 2A). In physiological buffer conditions (100 mM NaCl), 25 μM EWSLCD will form condensates (Fig. 2A). The addition of equimolar concentrations of FLI1DBD (25 µM) significantly increased the turbidity of the sample and dramatically altered the morphology of the condensates which showed reduced coalescence and became fusion-impaired over the 5-h equilibration (Fig. 2A, cf. Movies S1 and S2). The addition of EWS::FLI1 to EWSLCD condensates had a larger effect, significantly increasing the turbidity of the sample and bright-field microscopy revealed structures that appeared more aggregate-like than liquid-like droplets (Fig. 2A). At lower ratios of EWS::FLI1 to EWSLCD, structures that resemble condensates can still be observed (Fig. 2B) indicating that the alteration of EWSLCD condensate morphology is dependent on the concentration of EWS::FLI1. Turbidity measurements for EWSLCD, EWSLCD + FLI1DBD, and EWSLCD + EWS::FLI1 stabilize after ~2 h, with 5-h aged condensates maintaining the trends in turbidity seen in freshly prepared samples (SI Appendix, Fig. S2A). At 25 μM, full-length EWS::FLI1 alone formed rigid puncta rather than liquid-like droplets (Fig. 2A), providing further evidence that the FLI1DBD has an adverse effect on the liquidity of the EWSLCD condensates. To determine whether this effect was specific to the FLI1DBD, we also tested equimolar concentrations of constructs including the RNA-recognition motif (RRM) of EWS (EWSRRM or EWSRRM-RGG2, 25 µM) (Fig. 2A and Table 1 and SI Appendix, Fig. S1). EWSRRM had no significant effect on the turbidity of EWSLCD, while EWSRRM-RGG2 induced a slight increase in turbidity, and both EWSRRM and EWSRRM-RGG2 resulted in EWSLCD condensates that appeared more spherical and retained robust coalescence (Fig. 2A and Movies S3 and S4). Therefore, the FLI1DBD found in EWS::FLI1 increases the phase separation propensity of EWSLCD and changes the material properties of the condensates causing them to rapidly become fusion impaired, which may drive the different effect on the morphology of the condensates compared to the RNA-binding domains that are naturally found in EWS.

Fig. 2.

Fig. 2.

FLI1DBD alters the phase separation propensity of EWSLCD. (A) Turbidity and micrographs of 25 µM EWSLCD plus 25 µM of either FLI1DBD, EWSRRM, EWSRRM-RGG2, or EWS::FLI1, and controls (no EWSLCD) measured in phase separating conditions (100 mM NaCl) after 5 h equilibration. EWS::FLI1 alone is shown for comparison. ThT assay and micrographs of (B) 50 µM EWSLCD, 5 µM EWS::FLI1 alone, and 50 µM EWSLCD plus 0.1, 1, or 5 µM EWS::FLI1, and (C) 50 µM EWSLCD, 5 µM FLI1DBD (alone), and 50 µM EWSLCD plus 1 or 5 µM FLI1DBD in phase-separating conditions (150 mM NaCl). For (B and C) Images were taken at T > 10 h (Scale bars, 25 µm). ThT (turbidity) data are means (averages) of triplicates ± SEM. P-values: *P < 0.05; **P < 0.01; ****P < 0.0001.

Table 1.

Recombinant protein constructs of EWS, EWS::FLI1, and ETS DBDs

Construct name Affinity Tag MW (kDa) Margins*
EWS (full-length) 8× His 68.5 EWS 1–656
FUS (full-length) His-MBP 53.4 FUS 1–526

EWS::FLI1

(type 2, full-length)

8× His or 8× His-GFP 54.3 EWS 1–264 + FLI1 22 –452 (F408A)
EWSLCD 8× His 27.9 EWS 2–264
EWSRRM 8× His 9.9 EWS 359–447
EWSRRM-RGG2 8× His 16.3 EWS 359–513
EWSRGG3 10× His-MBP 8.73 EWS 549–638
FLI1DBD 8× His 14.6 FLI1 276–399 (F362A)
FLI1DBD C299S S390C 8× His 14.6 FLI1 276–399 (C299S, F362A, S390C)
FLI1DBD R2L2 8× His 14.4 FLI1 276–399 (R337L, R340L, F362A)
FLI1DBD Δα4 8× His 10.3 FLI1 276–361
ERGDBD 8× His 14.5 ERG 306–429 (F392A)
ETV1DBD 8× His 15.1 ETV1 332–458
PU.1DBD 8× His 12.3 PU.1 168–270
PU.1DBD, ETV1 swap 8× His 12.2 PU.1 168–270 (K204I, H205H, K206P, G281del, K221P, K222A, T224N, K245A, K246G, insertE247, K247R, L248Y)
PU.1DBD, no +ve charge 8× His 12.1 PU.1 168–270 (K204I, H205H, K206P, G281del, R220N, K221P, K222A, T224N, K245A, K246G, insertE247, K247R, L248Y)

*Canonical sequences from Uniprot.

Mutations EWS::FLI1 (F408A), FLI1 (F362A), and ERG (F392A) are equivalent (SI Appendix, Fig. S6).

Biomolecular condensates formed by LCD-containing proteins similar in composition to EWSLCD age spontaneously over time, becoming less liquid-like (41, 42). Cross-β structure has been shown to stabilize the condensed state of FUS (43), and condensate rigidification is concurrent with the formation of fibrillar structures that initially form at the interface between condensed and dilute phases and sometimes protrude from the condensates (41, 42). To investigate whether the changes in condensate morphology induced by EWS::FLI1 and FLI1DBD are due to the formation of cross-β structure within EWSLCD condensates, we developed ThT fluorescence assays. Under phase separating conditions (150 mM NaCl), EWSLCD condensates rigidify slowly over approximately 10 h, with a small concomitant increase in ThT fluorescence (Fig. 2 B and C). The addition of EWS::FLI1 or FLI1DBD, even at substoichiometric concentrations (as low as 1:500 molar ratio), to EWSLCD condensates significantly accelerated condensate rigidification in a manner dependent on the concentration of EWS::FLI1 or FLI1DBD (Fig. 2 B and C). FLI1DBD also increased ThT fluorescence of full-length EWS under phase separating conditions (SI Appendix, Fig. S2B). ThT-positive structures are not formed by FLI1DBD in the absence of EWSLCD at the low concentrations used in this assay (Fig. 2C). Under these conditions, we did not observe fibril formation by transmission electron microscopy (SI Appendix, Fig. S2C), indicating that the observed increase in ThT fluorescence arises from the formation of cross-β structure that has not had time, or is incapable of organizing into bona fide amyloid fibrils. This initial formation of protofibrillary species may be occurring at the interface between the condensed and dilute phase, as observed for FUS, which may account for the rigidification of condensates and the impaired ability of the condensates to fuse (42). The FLI1DBD is not known to form cross-β structure, whereas FET family proteins have been shown to form cross-β structures (27, 41, 42, 44, 45); therefore, it is logical to conclude that the FLI1DBD drives cross-β structure formation by EWSLCD but does not bind ThT itself. Together, these results indicate that the FLI1DBD found in EWS::FLI1 increases the rate at which EWSLCD condensates rigidify, possibly caused via the formation of ThT-positive cross-β structure.

DNA Binding to FLI1DBD Inhibits EWSLCD Condensate Rigidification.

The ability of EWS::FLI1 and FLI1DBD to rigidify EWSLCD condensates suggests that the proteins directly interact. To determine whether the interaction site involves the FLI1 DNA-binding site, we conducted turbidity assays in the presence of DNA (Fig. 3A). At 5 µM, EWS::FLI1 undergoes rapid condensate formation and/or aggregation and the samples become turbid immediately (Fig. 3A). The addition of equimolar concentrations of double-stranded, high-affinity (HA) DNA (46) reduced the turbidity of the sample, indicating that condensate formation and/or aggregation was inhibited (Fig. 3A and Table 2). A double-stranded DNA oligonucleotide containing 10 GGAA (GGAA10) repeats known to bind FLI1DBD in vitro (14, 47), exerted the same effect on EWS::FLI1 aggregation (SI Appendix, Fig. S3A). ThT assays demonstrated that an equimolar amount of HA DNA reduced EWS::FLI1-induced rigidification of EWSLCD condensates (Fig. 3B).

Fig. 3.

Fig. 3.

DNA binding by EWS::FLI1 or FLI1DBD inhibits phase separation and condensate rigidification. (A) Turbidity of 5 µM EWS::FLI1 (filled) or plus 5 µM HA DNA (open), and HA DNA after 5 h equilibration. (B) ThT assay of 50 µM EWSLCD, plus 1 µM of EWS::FLI1 (filled), or EWS::FLI1 + HA DNA (open). (C) Turbidity of 25 µM EWSLCD, plus 25 µM of FLI1DBD (filled), or FLI1DBD + HA DNA (open), and HA DNA, or FLI1DBD + HA DNA after 5 h equilibration. (D) ThT assay of 50 µM EWSLCD plus 5 µM of FLI1DBD (filled), or FLI1DBD + 3 (open) or 5 (light open) µM HA DNA. Fluorescence and DIC microscopy of freshly prepared samples of (E) EWSLCD + EWS::FLI1488 (Top) and EWSLCD + EWS::FLI1488 + HA DNA650 (Bottom), and (F) EWSLCD,650 + FLI1DBD,488 (Top) or EWSLCD,488 + FLI1DBD + HA DNA650 (Bottom). Scale bars, 10 µm, superscripts 488 or 650 indicate fluorescent dye labeling. Fluorescence intensity ratio (Iin/Iout) and FRAP of (G and H) EWS::FLI1488 in condensates composed of EWSLCD + EWS::FLI1488 ± HA DNA650, or (I and J) of FLI1DBD,488 in condensates composed of EWSLCD + FLI1DBD,488 + increasing amounts of HA DNA650. Experiments were conducted under phase separating conditions (100 or 150 mM NaCl), turbidity, ThT, FRAP (AC, D, H, and J), or intensity (G and H) data are shown as means of triplicates or 16 replicates ± SEM. P-values: **P < 0.01; ****P < 0.0001.

Table 2.

DNA oligonucleotides used for binding studies with ETS DBDs

Construct name Sequence
HA DNA TTTACCGGAAGTGTTT
10 x GGAA DNA CGCGGATCCGGAAGGAAGGAAGGAAGGAAGGAAGGAAGGAAGGAAGGAACTGCAGTTTT
Scrambled DNA TTGAGAGAGAGAGATT

Consistent with the findings for EWS::FLI1, the addition of HA DNA inhibited the ability of FLI1DBD to enhance EWSLCD condensate formation as evidenced by a reduction in turbidity relative to EWSLCD + FLI1DBD in the absence of DNA (Fig. 3C). Additionally, DNA binding inhibited the ability of FLI1DBD to drive EWSLCD condensate rigidification in a concentration-dependent manner (Fig. 3D). At the highest concentrations of HA DNA tested (50 µM corresponding to a 10:1 DNA:FLI1DBD ratio), the effect of FLI1DBD on EWSLCD condensates was almost entirely abolished (SI Appendix, Fig. S3B). The GGAA10 oligonucleotide also inhibited the effect FLI1DBD exerts on EWSLCD condensates (SI Appendix, Fig. S3C). A scrambled dsDNA control sequence (Table 2) had no inhibitory effect, demonstrating that the inhibitory effect of DNA in these ThT assays is due to DNA binding by FLI1 (SI Appendix, Fig. S3C). Collectively, the ThT assays and turbidity data demonstrate that DNA binding by EWS::FLI1 and FLI1DBD inhibits condensate/aggregate formation of EWS::FLI1 and reduces the interaction between FLI1DBD and EWSLCD.

We then used fluorescence microscopy to characterize the colocalization of the protein and DNA components comprising the condensates. To quantify the colocalization of the fluorescently labeled components to the EWSLCD, we calculated fluorescence intensity ratios inside versus outside (Iin/Iout) of the condensates. Fluorescent EWS::FLI1488 (labeled with DyLight 488) was mixed with EWSLCD without (Fig. 3 E, Top) or with (Fig. 3 E, Bottom) HA DNA650 (labeled with Cy5) under phase separating conditions. Fluorescence corresponding to EWS::FLI1488 was localized in the condensates, indicating that EWS::FLI1 interacts with EWSLCD in the condensates. Surprisingly, HA DNA was largely excluded from EWSLCD + EWS::FLI1 condensates with an intensity ratio of 0.5 ± 0.1 (Fig. 3 E, Bottom). Similarly, we mixed the EWSLCD,650 (EWSLCD labeled with DyLight 650) with FLI1DBD,488 under phase separating conditions and the FLI1DBD,488 signal spatially overlapped with the EWSLCD,650 condensates, indicating partitioning of FLI1DBD into EWSLCD condensates (Fig. 3 F, Top). Consistent with the behavior of multicomponent (EWSLCD + EWS::FLI1 + HA DNA) condensates, mixing EWSLCD,488 condensates with unlabeled FLI1DBD and HA DNA650 revealed that HA DNA was moderately excluded from the condensates (Fig. 3 F, Bottom) with an intensity ratio of 0.8 ± 0.02. The partitioning of the FLI1DBD to EWSLCD condensates is specific and further reinforces that these domains interact since unrelated proteins such as green fluorescent protein (GFP) do not colocalize to EWSLCD condensates unless fused to EWSLCD (SI Appendix, Fig. S4A). Additionally, FLI1DBD,488 does not form visible condensates under these conditions (SI Appendix, Fig. S4B).

The fluorescence intensity ratio for EWS::FLI1488 colocalization to EWSLCD condensates in the absence of DNA was 5.4 ± 0.9, indicating that EWS::FLI1 is enriched in the condensates (Fig. 3G). Addition of HA DNA resulted in reduced colocalization of EWS::FLI1 to EWSLCD condensates with a measured intensity ratio of 2.5 ± 0.1 (Fig. 3G). FRAP of EWS::FLI1488 within EWSLCD condensates reached ~46% of the steady-state intensity 30 s after photobleaching, while in the presence of HA DNA, fluorescence recovery reached ~63% after the same interval, indicating that the addition of DNA increases the liquidity of EWS::FLI1488-containing condensates (Fig. 3H). This observation may be due to a reduction of the concentration of EWS::FLI1 in the condensates due to DNA binding (Fig. 3G), thereby reducing its adverse effect on condensate liquidity.

Analogous to the observations for EWS::FLI1, fluorescence microscopy revealed that in the absence of DNA, the Iin/Iout ratio of FLI1DBD,488 was ~5.8 ± 2.0 (Fig. 3I). As the concentration of HA DNA was increased, the intensity of FLI1DBD,488 in condensates became noticeably reduced (SI Appendix, Fig. S5A), and the Iin/Iout intensity ratio reduced to 1.6 ± 0.2 for EWSLCD,650 with FLI1DBD,488 and 50 μM HA DNA (Fig. 3I), consistent with the effect observed for EWS::FLI1 (cf. Fig. 3G). Therefore, DNA binding by EWS::FLI1 or FLI1DBD outcompetes the interactions between EWSLCD and FLI1DBD that drive condensate colocalization, with DNA possibly acting as a sink for EWS::FLI1 and FLI1DBD, precluding them from entering the condensate. The EWSLCD,650 within condensates containing only EWSLCD underwent rapid recovery after photobleaching, reaching ~91% of the initial fluorescence, indicating liquid-like condensate behavior (Fig. 3J). In contrast, freshly prepared condensates formed in the presence of FLI1DBD displayed slower fluorescence recovery, reaching only ~40% recovery indicating that the dynamics of EWSLCD molecules within the condensates change in the presence of FLI1DBD (Fig. 3J and SI Appendix, Fig. S5B). With the addition of HA DNA to EWSLCD + FLI1DBD condensates, fluorescence recovered more rapidly in a DNA concentration-dependent manner, consistent with DNA binding inhibiting the effect of FLI1DBD on EWSLCD (Fig. 3J).

Exclusion of DNA from EWSLCD condensates may be due to a lack of charge neutralization of the DNA phosphate backbone within the condensates since the EWSLCD contains only 2 positively charged residues (out of 264) in contrast with 6 negatively charged residues and 27 tyrosines with delocalized π electrons in their sidechain rings. Pelleting assays revealed that although condensates are readily visible via microscopy, the majority (>95%) of the EWSLCD remains in the dilute phase (supernatant) in the absence of the FLI1DBD (SI Appendix, Fig. S5C). However, in the presence of the FLI1DBD, the partitioning of EWSLCD into the pellet fraction was noticeably higher (~50% of total EWSLCD), consistent with the FLI1DBD enhancing EWSLCD condensate formation (Fig. 2A and SI Appendix, Fig. S5C). The FLI1DBD also partitioned into the pellet fraction (~55% of total FLI1DBD), indicating direct association with the EWSLCD (SI Appendix, Fig. S5C). Addition of HA DNA and FLI1DBD reduced the partitioning of both EWSLCD and FLI1DBD (~90% of each protein remains in the dilute phase) into the pellet further indicating that DNA inhibits the EWSLCD–FLI1DBD interaction (SI Appendix, Fig. S5C). The HA DNA alone had no effect on EWSLCD condensate formation and the FLI1DBD incubated alone remained in the dilute phase (SI Appendix, Fig. S5C).

The FLI1 DNA-Binding and Dimer Interfaces Are Not Involved in the Interaction with EWSLCD.

The FLI1DBD DNA recognition helix (α3) harbors two highly conserved arginine residues that contact core bases of the ETS consensus sequence in all the reported crystal structures of ETS DBDs complexed with DNA (4852) (Fig. 4A and SI Appendix, Fig. S6A). To test whether these arginines participate in Arg-Tyr π-cation interactions with EWSLCD as a possible mechanism that promotes rapid condensate rigidification, we mutated both arginines to leucines (FLI1DBD,R2L2, SI Appendix, Fig. S1). This mutant is incapable of binding HA DNA (SI Appendix, Fig. S7A) and its effect on EWSLCD condensates was assessed using ThT assays (Fig. 4B). Surprisingly, the FLI1DBD,R2L2 mutant retained the same ability to enhance the rate of EWSLCD condensate rigidification in the presence or absence of DNA (Fig. 4B). Together these results indicate that the positively charged arginines located on the DNA recognition helix of FLI1 are not involved in the EWSLCD–FLI1DBD interaction. Furthermore, in line with how HA DNA affects the ThT-positive response of EWSLCD + FLI1DBD condensates, the inability of FLI1DBD,R2L2 to bind DNA correlated with its failure to inhibit the increase in ThT positivity within EWSLCD condensates, even when DNA was present.

Fig. 4.

Fig. 4.

FLI1DBD helices α3 and α4 do not contribute to the rigidification of EWSLCD condensates. (A) FLI1DBD structure highlighting the DNA recognition helix (α3), conserved R337 and R340 that contact DNA (SI Appendix, Fig. S6), and C-terminal helix α4 (5e8g). ThT assay of 50 µM EWSLCD alone (black), and plus 5 µM of either FLI1DBD (red), or (B) FLI1DBD + HA DNA (red open), FLI1DBD,R2L2 (cyan), or FLI1DBD,R2L2 + HA DNA (cyan open), or (C) FLI1DBD,Δα4 mutant (blue). ThT assays were conducted under phase separating conditions (150 mM NaCl), with data shown as means of triplicates ± SEM.

The structures of ETS DBDs solved to date share a common winged helix–turn–helix fold (SI Appendix, Fig. S6B). However, the reported boundaries of the ETS domains vary. Some studies have used an 85 amino acid construct containing only three α-helices (17, 5355), while others have employed a longer construct that includes a fourth α-helix (9, 17, 49). A recent crystallographic study showed the FLI1DBD forming a dimer, with the fourth α-helix participating in the dimer interface (49). In this study, we used a FLI1DBD construct that includes the fourth α-helix (residues 362 to 369), but to prevent dimerization we introduced the F408A (EWS::FLI1 sequence numbering) mutation and included ~30 C-terminal disordered residues (SI Appendix, Figs. S1 and S6A). To assess whether the fourth α-helix, which contains exposed hydrophobic residues, interacts with EWSLCD, we performed ThT assays using FLI1DBD,Δα4, a FLI1DBD construct lacking residues beyond 361 (Fig. 4C). As expected, this truncated construct retained DNA binding activity, as confirmed by electrophoretic mobility shift assays (SI Appendix, Fig. S7B). Moreover, in ThT assays, it affected EWSLCD condensates similarly to the long construct (Fig. 4C). These findings define the EWSLCD interaction site responsible for inducing EWSLCD rigidification within condensates, narrowing it to the folded core of FLI1 (residues 276-361) while excluding the arginines of the DNA recognition helix (α3, Fig. 4A).

Enhancement of EWSLCD Condensate Rigidification Is a General Property of ETS DBDs.

Beyond EWS::FLI1, other EWS–ETS fusions cause EwS. To determine whether FLI1DBD condensate rigidification activity is linked to the ETS DBD fold, we performed ThT assays with three additional ETS DBDs and three constructs containing segments of EWS’s C-terminal RNA-binding domain (EWSRRM, EWSRRM-RGG2, EWSRGG3, Fig. 5 and SI Appendix, Fig. S1). Two of these ETS DBDs (ERG, and ETS translocation variant 1, ETV1) are also found in oncogenic fusions with EWSLCD, while the third, transcription factor PU.1 (PU.1) is oncogenic but not known to fuse with EWS in EwS. All constructs were properly folded, as judged from circular dichroism spectra (SI Appendix, Fig. S8). Bright-field microscopy images were captured at the end of the ThT assay (T > 10 h) (Fig. 5). Compared with EWSLCD alone, EWSLCD + EWSRRM or EWSRRM-RGG2 had minimal impact on the rate of ThT fluorescence increase, with condensates remaining well-dispersed and predominantly spherical (Fig. 5 A, C, and E). In contrast, all four ETS DBDs significantly enhanced EWSLCD condensate rigidification (Fig. 5 B, D, F, and H), leading to irregular morphologies, particularly in the presence of PU.1DBD and ETV1DBD (Fig. 5 B and H). In the absence of EWSLCD condensates none of the constructs induced a ThT signal (SI Appendix, Fig. S9A). Notably, PU.1DBD induced the fastest ThT fluorescence increase, disrupting condensate coalescence and fusion even more severely than FLI1DBD (Movie S5). Surprisingly, EWSRGG3 also increased ThT positivity in a concentration-dependent manner (Fig. 5G and SI Appendix, Fig. S9B). However, unlike ETS DBD-induced condensates, these did not develop the same irregular shapes. Instead, they rapidly coalesced and spread out across the chamber bottom, resembling surface wetting (Fig. 5G and Movie S6).

Fig. 5.

Fig. 5.

Rigidification effect on EWSLCD condensates is conserved for other ETS DBDs. ThT assays (Left) and micrographs of aged samples (T ~ 24 h, Right) of (A) 50 µM EWSLCD incubated alone or with (B) 5 µM of PU.1DBD, (C) EWSRRM, (D) FLI1DBD, (E) EWSRRM-RGG2, (F) ERGDBD, (G) EWSRGG3, and (H) ETV1DBD. Samples were prepared under phase separating conditions (150 mM NaCl). Dashed lines estimate half maximum signal. Images were taken at T > 10 h (Scale bars, 50 µm). ThT data are means of triplicates ± SEM.

The addition of HA DNA with ERGDBD and PU.1DBD to EWSLCD condensates inhibited rigidification, similar to FLI1DBD, suggesting a structurally conserved mechanism (SI Appendix, Fig. S9C). To rule out His-tag effects we tested a His-tag free version of PU.1DBD (SI Appendix, Fig. S9D). Since His-tagged EWSRRM and EWSRRM-RGG2 did not induce condensate rigidification, and the His-tag free PU.1DBD behaved identically to its tagged counterpart, the His-tag was not a contributing factor (SI Appendix, Fig. S9D). These results indicate that the impact of FLI1DBD on EWSLCD condensates is conserved across ETS transcription factors and is linked to the fold of the winged helix–turn–helix motif.

ETS DBDs Interact with the EWSLCD via Residues Adjacent to the DNA-Binding Face.

We used solution NMR to map the interaction of ETS DBDs with the EWSLCD. Given the transient nature of the interaction and the challenges of studying an aggregation-prone system, we selected the PU.1DBD for these experiments. Backbone resonances of PU.1 were assigned using standard approaches (56). 1H,15N heteronuclear single quantum coherence spectra were recorded for a sample containing an ~3:1 molar ratio of EWSLCD to 15N PU.1DBD (SI Appendix, Fig. S10). Addition of higher concentrations of EWSLCD were not possible because PU.1DBD induced EWSLCD condensate formation even in low salt conditions. Nevertheless, we observed small chemical shift perturbations (CSP) (δΔ ~ 0.04 ppm) for PU.1 resonances (Fig. 6A and SI Appendix, Fig. S10). The signal intensity uniformly decreased between a sample of 15N PU.1 and the 3:1 sample likely due to the formation of complexes between PU.1DBD and EWSLCD that are beyond the size-limits for detection with solution NMR (Fig. 6B). However, a few peaks were differentially broadened, and coincided with or were located near residues with CSPs (Fig. 6 A and B). We plotted residues with CSPs greater than one SD from the mean and residues with differential signal intensities less than one SD from the mean onto the AlphaFold (57) structure of human PU.1 (Fig. 6C). Notably, these residues clustered to one face of the DBD, including residues in the ETS DBD “wings” (loops 3, 4, and 6) that contact the phosphate backbone of DNA (50); however, no shifts or differential broadening were observed for the DNA recognition helix (α3), consistent with the effects observed with the FLI1DBD,R2L2 mutant (Fig. 4B). Furthermore, we observed no shifts or peak broadening on the opposite face of the DBD (Fig. 6C). The CSPs and differential broadening of residues in the disordered C-terminal tail were considered nonspecific to condensate rigidification. This is because the C-terminally truncated ETS domain (FLI1DBD, Δα4) which lacks these residues, still retained condensate rigidification activity (Fig. 4C and SI Appendix, Fig. S1) and this region is poorly conserved among the ETS DBDs (SI Appendix, Fig. S6A).

Fig. 6.

Fig. 6.

Residues in the DBD wings of PU.1 are involved in the interaction with EWSLCD. (A) CSP and (B) signal broadening of 15N-PU.1DBD residues at a 3:1 molar ratio with EWSLCD, SD marked with red dashed lines. Black (red) asterisks indicate overlapped/ambiguously assigned (proline) residues. Residues with CSPs (intensity differences) greater (less) than 1 SD are orange filled, and (C) mapped to the AlphaFold model of the human PU.1DBD. (D) Sequence alignment and net charge of ETSDBD loops 3, 4, and 6, conserved positive charges are bolded (SI Appendix, Fig. S6). (E) ThT assay and (F) micrographs of 50 µM EWSLCD alone (black), with 5 µM of PU.1DBD (blue), PU.1DBD, ETV1 swap (ETV sequence in loops, green open circles), or PU.1DBD, no +ve charge (all charged residues in loops are mutated to neutral, teal). Images were taken at T > 10 h (Scale bars, 50 µm). ThT data are means of triplicates ± SEM.

A sequence comparison of loops 3, 4, and 6 from the four ETS DBDs tested in the ThT assays (Fig. 5) revealed that the total net charge varies between +8 to 0 (PU.1 > FLI1 = ERG > ETV1, Fig. 6D). The loops in PU.1 are dynamic (SI Appendix, Fig. S11) and enriched in lysine residues, giving them a net charge of +8. Notably, loop 3 is the only loop containing positively charged residues and lacks any negatively charged residues, resulting in a net charge of +2 (Fig. 6D). Loops 3, 4, and 6 of FLI1 and ERG are identical with a net charge of +2, and in ETV1 they are depleted in Lys with a net charge of 0. Loops 4 and 6 in all the ETS DBDs tested contain one highly conserved, positively charged residue (Arg or Lys) at 220 and 247 (PU.1 sequence numbering, Fig. 6D). ThT assays showed that the PU.1DBD rigidified condensates the fastest, followed by FLI1DBD and ERGDBD, and finally ETV1DBD which had the slowest effect (Fig. 5 B, D, F, and H respectively). The net charge of loops 3, 4, and 6 directly correlates with the rate of ETS DBD-induced EWSLCD condensate rigidification, suggesting that electrostatic interactions play a role. To test this, we constructed two PU.1DBD mutants: one in which loops 3, 4, and 6 were replaced with those from ETV1 (PU.1DBD,ETV1 swap) and another where all positive charges were removed from loops 3, 4, and 6 (PU.1DBD,no +ve charge) (SI Appendix, Fig. S1). The PU.1DBD,ETV1 swap mutant rigidified EWSLCD condensates at a slower rate than wild-type PU.1DBD and had a less severe effect on condensate morphology (Fig. 6 E and F). Subsequently, the PU.1DBD,no +ve charge mutant further reduced the rate of ThT positivity and the condensates appeared even more similar to EWSLCD-only condensates, although the difference in the rate of ThT-positivity between the two mutants appears minimal (Fig. 6 E and F). These findings support a loop-mediated condensate rigidification mechanism involving specific interactions with the EWSLCD. However, the PU.1DBD,ETV swap mutant still rigidified condensates faster than ETV1DBD, suggesting loops 3, 4, and 6 do not completely account for the rate differences observed between PU.1DBD and ETV1DBD. Electrostatic interactions alone do not determine condensate rigidity. The FLI1DBD,R2L2 mutant rigidified EWSLCD condensates at the same rate as FLI1DBD (Fig. 4B), and although EWSRRM-RGG2 is rich in positively charged Arg residues, it did not rigidify EWSLCD condensates like the ETS DBDs (Fig. 5). Similarly, EWSRGG3, which has a net charge of +10, also induced rigidification but to a lesser extent than the ETS DBDs (Fig. 5 and SI Appendix, Fig. S9B).

To further examine the role of positive charge in the condensate rigidification, we tested 30- and 50-residue poly-L-lysine peptides for their effect on ThT positivity and condensate formation. At the same molar ratio (1:10) where ETS DBDs induced a rapid ThT response, neither peptide produced strong ThT positivity (SI Appendix, Fig. S9E). Poly-L-lysine 30 slightly impaired condensate coalescence, with some droplets failing to merge (SI Appendix, Fig. S9E and Movie S7), whereas poly-L-lysine 50 maintained robust coalescence (SI Appendix, Fig. S9E and Movie S8). These findings suggest that positive charge alone does not fully explain the rigidifying effect of ETS DBDs on EWSLCD condensates. Instead, the structural arrangement of the three loops and their ability to adopt conformations favorable for EWSLCD may contribute to the liquid-to-solid transition.

Discussion

EWS has roles in transcriptional regulation, homologous recombination, DNA damage response, and splicing, though its precise functions remain unclear. Notably, knockout of EWS is postnatally lethal in mice (15, 5860). The expression of EWS::FLI1 mimics the effects of EWS knockdown in HeLa and EwS cell, supporting the hypothesis that it acts as a dominant-negative regulator of EWS function (15, 17, 18, 61). This was recently confirmed in EwS cells, where EWS::FLI1 disrupts EWS’s regulation of RNA Pol II phosphorylation by CKD7/9 (15). We demonstrated that the FLI1DBD in EWS::FLI1 enhances EWSLCD-mediated condensate formation, partitions into EWSLCD condensates, and accelerates their rigidification. This effect was conserved across three other ETS DBDs, suggesting a shared mechanism common to ETS DBDs. Boulay et al. recently reported that β-isoxazole induced precipitation of EWS::FLI1 in EwS cell lines was stronger than that of EWS (4), further supporting the hypothesis that the FLI1DBD interaction increases the intrinsic phase separation propensity of the EWSLCD. Studies in prostate cancer cell lines identified intermolecular interactions between EWS and ERG, ETV1, ETV4, and ETV5 that were necessary and sufficient for oncogenesis, reinforcing the idea that FET–ETS interactions can have oncogenic consequences (38). These findings suggest that EWS is recruited to ETS-bound DNA enhancers through regions flanking the ETS DBD, highlighting the need for further investigation into EWS–ETS interactions (38). Furthermore, coimmunoprecipitation experiments demonstrated a direct interaction between PU.1 and FUS that inhibits the normal regulatory functions of FUS in splicing (62, 63). Therefore, mislocalization of ETS domains either as fusions or due to aberrant regulation appears to have a dual effect of enhancing self-association and negatively impacting the functions of FET-family proteins and possibly other nucleic acid-binding proteins. We found that the interaction between ETS DBDs and EWSLCD involves residues in the “wings” (loops 3, 4, and 6) that are partially occluded by the DNA phosphate backbone in the DNA-bound state. As a result, DNA binding reduces the interactions between EWS::FLI1 or ETS DBDs and EWSLCD that promote condensation (Fig. 7 A and B). The in vivo situation is more complicated since other proteins, posttranslational modifications, and nucleic acids surely contribute to and modulate condensate properties. We confirmed that EWS::FLI1 disrupts condensate liquidity in a complex cellular environment, but more work is required to determine how other condensate components may modulate ETS DBD-induced rigidity.

Fig. 7.

Fig. 7.

ETS DBDs enhance the rigidification of biomolecular condensates containing EWS::FLI1 or the EWSLCD. Without ETS DBDs, EWSLCD exists in an equilibrium between monomers, liquid-like condensates, and rigid condensates. (A) Colocalization of ETS DBDs enhances EWSLCD phase separation and accelerates rigidification, forming ThT-positive cross-β structures. (B) DNA binding by free ETS DBDs prevents their colocalization with EWSLCD reducing their impact condensate formation.

EWSLCD self-association stabilizes EWS::FLI1 binding to DNA enhancers, aids in recruiting RNA Pol II and the BAF complex, and is essential for oncogenic gene activation (4, 6, 28). Chong et al. overexpressed exogenous EWSLCD in EwS patient–derived cells to artificially increase LCD–LCD and assess their impact on EWS::FLI1 driven transcription. They found that excess EWSLCD reduced transcriptional activation likely by increasing LCD–LCD interactions (39), suggesting transcription requires a finely tuned level of these interactions. Our findings align with this model, as excess EWSLCD counteracted EWS::FLI1 effects and altered condensate liquidity. This raises the question of whether EWS::FLI1’s tunable transcriptional activity depends on the number of molecules in transcriptional hubs or on the regulation of intermolecular dynamics within them. These findings also provide mechanistic insight into the proposed Goldilocks effect of EWS::FLI1 toxicity where too high expression of the fusion possibly results in uncontrollable aggregation and eventual cellular death (34). Following this logic, certain ETS fusions such as EWS::PU.1 are not observed since they may produce a protein too toxic to be tolerated by the cell. Given that FLI1DBD affects EWSLCD condensate liquidity, further investigation is needed in the context of full-length EWS::FLI1 rather than isolated EWSLCD constructs. We note that such experiments are extremely challenging.

The three FET family proteins can form condensates and hydrogels (27, 28). Hydrogels formed by their LCDs contain cross-β structure, which facilitate copolymerization with binding partners like the RNA Pol II CTD and hnRNPA1/2 (5, 27, 42). Given this, it is unsurprising that the EWSLCD can form ThT-positive condensates, indicative of cross-β structures (64). High-resolution structural models of cross-β structures exist for short segments of FUS (65, 66) and for the full FUS LCD (43). Recent findings show that FET fibrilization begins at the condensate surface, forming a rigid, protofibrillary shell, stabilized by ThT-positive cross-β structures (42)—features resembling EWS::FLI1 condensates. If EWSLCD forms cross-β structures through parallel, in-register β-strands, it would create “ladders” of identical sidechains stacked at intervals matching the interstrand distance (67). These stabilizing ladders form either through π–π stacking of aromatic tyrosine residues or by complementary hydrogen bonding of glutamine/asparagine residues (67). The exact mechanism by which ETS DBDs disrupt the intra- and intermolecular interactions governing phase separation and cross-β formation remains unclear. However, tyrosine residues in the EWSLCD are essential for functional self-association (4, 33, 68). Indeed, removing all tyrosine residues from the LCD abolishes its ability to form condensates (29, 33), prevents EWS::FLI1 from recruiting the BAF complex (4), and eliminates the transformative activity of EWS::FLI1 (68). Our ThT assays, NMR, and mutagenesis studies indicated that residues in loops 3, 4, and 6 of the ETS DBDs contribute positively charged residues (e.g., lysine in PU.1) to the interaction with the EWSLCD. Supporting this theory, charged residues such as lysine have been demonstrated to be important for the interaction of the CTD of RNA Pol II with FET family proteins, TAF15 and FUS (69, 70). Since the highly positively charged EWSRRM-RGG2 and poly-L-lysine constructs had only a minimal effect on EWSLCD condensate rigidity, the conformation of the ETS DBD loops relative to each other must also contribute to condensate rigidification. The ThT increase we observed with EWSRGG3 suggests additional factors may contribute to ThT positivity, beyond cross-β formation. Changes in condensate viscosity, known to affect ThT fluorescence, could also play a role (71).

While fluorescence microscopy revealed, unexpectedly, that HA DNA is mostly excluded from condensates formed by EWSLCD + EWS::FLI1 or FLI1DBD, posttranslational modifications, or additional binding partners such as EWS or other nucleic acid binding proteins may provide sufficient charge neutralization to allow for colocalization of DNA to condensates containing EWS::FLI1. Zuo et al. reported GFP-tagged EWS::FLI1 formed condensates that colocalized with DNA (6). Here, untagged EWS::FLI1 formed ThT-positive aggregates too quickly without condensate partners, such as EWSLCD, preventing accurate study of its condensate formation with DNA, consistent with the findings of Ryan et al. (72). The ligation of a highly soluble domain, such as GFP, to EWS::FLI1 may reduce the aggregation rate sufficiently to better study the effect of DNA or other additives on condensate formation without EWSLCD present (73). Future studies with relevant protein and nucleic acid partners will be needed to further investigate the biophysical properties of complex multicomponent condensates containing EWS::FLI1.

EWS::FLI1 drives oncogenesis through dysregulation of genes downstream of DNA enhancers, aberrant transcription, and deregulation of alternative splicing programs, all of which are dependent on an intact DBD (1, 60, 74). As part of its oncogenic function, EWS::FLI1 acts as a dominant-negative against the wild-type copy of EWS (15, 30), and associates with a multitude of nucleic acid binding proteins such as FUS, and RNA Pol II (7, 16, 32). The structural biology and biological implications of these interactions remain poorly understood, due to their heterogenous and transient nature. Here, we demonstrated that the ETS DBDs interact transiently yet specifically with the EWSLCD, altering the intermolecular interactions that govern its condensate formation and subsequently promoting condensate rigidification (Fig. 7). Our findings provide a mechanism that explains how EWS::FLI1 exerts a dominant negative influence over the normal functions of EWS, alters the properties of condensates and potentially interferes with the native functions of other RNA-binding proteins. In EwS, the mislocalization of the FLI1DBD (or other ETS DBDs) in FET-fusion proteins promotes EWS self-association and aggregation, disrupting local protein network dynamics and interfering with essential intermolecular interactions.

Materials and Methods

Briefly, U2OS-lac cells were cultured and transfected with Halo-tagged EWS and EWS::FLI1 constructs, FRAP was used to assess in vitro and in cellulo condensate properties. Protein constructs were expressed in Escherichia coli, purified via immobilized metal affinity and size-exclusion chromatography, and labeled for fluorescence imaging. Turbidity and ThT assays were performed to assess condensate formation and rigidification. NMR spectroscopy was used to analyze ETS DBD and EWSLCD interactions. Detailed descriptions of the methods used in this study are available in SI Appendix.

Supplementary Material

Appendix 01 (PDF)

Movie S1.

Dynamics of condensates formed by 25 μM EWSLCD in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 30 μm.

Download video file (9MB, mp4)
Movie S2.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM FLI1DBD in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 30 μm.

Download video file (8.6MB, mp4)
Movie S3.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM EWSRRM in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (9.4MB, mp4)
Movie S4.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM EWSRRMRGG2 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (8.9MB, mp4)
Movie S5.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM PU.1DBD in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10x magnification. Video speed is increased 6×. Scale bar = 30 μm.

Download video file (7.4MB, mp4)
Movie S6.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM EWSRGG3 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (10.6MB, mp4)
Movie S7.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM poly-L-lysine 30 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (8.9MB, mp4)
Movie S8.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM poly-L-lysine 50 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (9MB, mp4)

Acknowledgments

This work was supported by the St. Baldrick’s Foundation’s Shohet Family Fund for Ewing Sarcoma Research 634706 (D.S.L.), CPRIT Research Training Award RP170345 (E.E.S.), F31CA295030 (E.J.S.), Greehey Graduate Fellowship in Children’s Health and T32CA291696 (A.S.), Cancer Research UK Stand Up 2 Cancer RT6187, R01CA241554 (A.J.R.B.), the Welch Foundation AQ-2001-20190330, and R01GM140127 (D.S.L.). We would like to thank Drs. Richard Young and Shasha Chong for their generous gifts of reagents; Ms. Barbara Hunter for assistance with transmission electron microscopy; and Dr. Kristin Cano for NMR technical assistance and valuable discussions. This work is based upon research conducted in the Structural Biology Core Facilities, a part of the Institutional Research Cores at the University of Texas Health Science Center at San Antonio supported by the Office of the Vice President for Research and the Mays Cancer Center Drug Discovery and Structural Biology Shared Resource (NIH P30 CA054174).

Author contributions

E.E.S. and D.S.L. designed research; E.E.S., A.S., A.K.M.-R., S.A., and X.X. performed research; E.E.S., E.J.S., A.S., A.K.M.-R., S.A., X.X., and D.S.L. analyzed data; A.J.R.B. and D.S.L. acquired funding; and E.E.S., E.J.S., A.S., A.J.R.B., and D.S.L. wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

This article is a PNAS Direct Submission.

Data, Materials, and Software Availability

Expression plasmids data have been deposited in Addgene (188042188053) (7585).

Supporting Information

References

  • 1.Grunewald T. G. P., et al. , Ewing sarcoma. Nat. Rev. Dis. Primers 4, 5 (2018). [DOI] [PubMed] [Google Scholar]
  • 2.Hollenhorst P. C., McIntosh L. P., Graves B. J., Genomic and biochemical insights into the specificity of ETS transcription factors. Annu. Rev. Biochem. 80, 437–471 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Fisher C., The diversity of soft tissue tumours with EWSR1 gene rearrangements: A review. Histopathology 64, 134–150 (2014). [DOI] [PubMed] [Google Scholar]
  • 4.Boulay G., et al. , Cancer-specific retargeting of BAF complexes by a prion-like domain. Cell 171, 163–178.e19 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Kwon I., et al. , Phosphorylation-regulated binding of RNA polymerase II to fibrous polymers of low-complexity domains. Cell 155, 1049–1060 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Zuo L., et al. , Loci-specific phase separation of FET fusion oncoproteins promotes gene transcription. Nat. Commun. 12, 1491 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Petermann R., et al. , Oncogenic EWS-Fli1 interacts with hsRPB7, a subunit of human RNA polymerase II. Oncogene 17, 603–610 (1998). [DOI] [PubMed] [Google Scholar]
  • 8.Vasileva E., et al. , Origin of Ewing sarcoma by embryonic reprogramming of neural crest to mesoderm. bioRxiv [Preprint] (2024). 10.1101/2024.10.27.620438 (Accessed 20 March 2025). [DOI]
  • 9.May W. A., et al. , Ewing sarcoma 11;22 translocation produces a chimeric transcription factor that requires the DNA-binding domain encoded by FLI1 for transformation. Proc. Natl. Acad. Sci. U.S.A. 90, 5752–5756 (1993). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.May W. A., et al. , The Ewing’s sarcoma EWS/FLI-1 fusion gene encodes a more potent transcriptional activator and is a more powerful transforming gene than FLI-1. Mol. Cell Biol. 13, 7393–7398 (1993). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Johnson K. M., et al. , Role for the EWS domain of EWS/FLI in binding GGAA-microsatellites required for Ewing sarcoma anchorage independent growth. Proc. Natl. Acad. Sci. U.S.A. 114, 9870–9875 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Lessnick S. L., Braun B. S., Denny C. T., May W. A., Multiple domains mediate transformation by the Ewing’s sarcoma EWS/FLI-1 fusion gene. Oncogene 10, 423–431 (1995). [PubMed] [Google Scholar]
  • 13.Gangwal K., Close D., Enriquez C. A., Hill C. P., Lessnick S. L., Emergent properties of EWS/FLI regulation via GGAA microsatellites in Ewing’s sarcoma. Genes Cancer 1, 177–187 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Gangwal K., et al. , Microsatellites as EWS/FLI response elements in Ewing’s sarcoma. Proc. Natl. Acad. Sci. U.S.A. 105, 10149–10154 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Gorthi A., et al. , EWS-FLI1 increases transcription to cause R-loops and block BRCA1 repair in Ewing sarcoma. Nature 555, 387–391 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Yang L., Chansky H. A., Hickstein D. D., EWS.Fli-1 fusion protein interacts with hyperphosphorylated RNA polymerase II and interferes with serine-arginine protein-mediated RNA splicing. J. Biol. Chem. 275, 37612–37618 (2000). [DOI] [PubMed] [Google Scholar]
  • 17.Boone M. A., et al. , The FLI portion of EWS/FLI contributes a transcriptional regulatory function that is distinct and separable from its DNA-binding function in Ewing sarcoma. Oncogene 40, 4759–4769 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Jaishankar S., Zhang J., Roussel M. F., Baker S. J., Transforming activity of EWS/FLI is not strictly dependent upon DNA-binding activity. Oncogene 18, 5592–5597 (1999). [DOI] [PubMed] [Google Scholar]
  • 19.Weber S. C., Brangwynne C. P., Getting RNA and protein in phase. Cell 149, 1188–1191 (2012). [DOI] [PubMed] [Google Scholar]
  • 20.Boeynaems S., et al. , Protein phase separation: A new phase in cell biology. Trends Cell Biol. 28, 420–435 (2018), 10.1016/j.tcb.2018.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Mitrea D. M., Kriwacki R. W., Phase separation in biology; functional organization of a higher order. Cell Commun. Signal 14, 1 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Hnisz D., Shrinivas K., Young R. A., Chakraborty A. K., Sharp P. A., A phase separation model for transcriptional control. Cell 169, 13–23 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Boehning M., et al. , RNA polymerase II clustering through carboxy-terminal domain phase separation. Nat. Struct. Mol. Biol. 25, 833–840 (2018). [DOI] [PubMed] [Google Scholar]
  • 24.Liao S. E., Regev O., Splicing at the phase-separated nuclear speckle interface: A model. Nucleic Acids Res. 49, 636–645 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Strom A. R., et al. , Phase separation drives heterochromatin domain formation. Nature 547, 241–245 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Protter D. S., Parker R., Principles and properties of stress granules. Trends Cell Biol. 26, 668–679 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Kato M., et al. , Cell-free formation of RNA granules: Low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Chong S., et al. , Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 361, eaar2555 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Nosella M. L., et al. , O-linked-N-acetylglucosaminylation of the RNA-binding protein EWS N-terminal low complexity region reduces phase separation and enhances condensate dynamics. J. Am. Chem. Soc. 143, 11520–11534 (2021). [DOI] [PubMed] [Google Scholar]
  • 30.Spahn L., et al. , Homotypic and heterotypic interactions of EWS, FLI1 and their oncogenic fusion protein. Oncogene 22, 6819–6829 (2003). [DOI] [PubMed] [Google Scholar]
  • 31.Ahmed N. S., et al. , Fusion protein EWS-FLI1 is incorporated into a protein granule in cells. RNA 27, 920–932 (2021), 10.1261/rna.078827.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Bertolotti A., et al. , 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 hTAFII68, and subunits of TFIID and RNA Polymerase II complexes. Mol. Cell Biol. 18, 1489–1497 (1998). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Johnson C. N., et al. , Insights into molecular diversity within the FUS/EWS/TAF15 protein family: Unraveling phase separation of the N-Terminal low-complexity domain from RNA-Binding Protein EWS. J. Am. Chem. Soc. 146, 8071–8085 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Seong B. K. A., et al. , TRIM8 modulates the EWS/FLI oncoprotein to promote survival in Ewing sarcoma. Cancer Cell 39, 1262–1278.e7 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Burke K. A., Janke A. M., Rhine C. L., Fawzi N. L., Residue-by-residue view of in vitro FUS granules that bind the C-terminal domain of RNA Polymerase II. Mol. Cell 60, 231–241 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Qamar S., et al. , FUS Phase separation is modulated by a molecular chaperone and methylation of arginine cation-pi interactions. Cell 173, 720–734.e15 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Martin E. W., et al. , Valence and patterning of aromatic residues determine the phase behavior of prion-like domains. Science 367, 694–699 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kedage V., et al. , An interaction with Ewing’s sarcoma breakpoint protein EWS defines a specific oncogenic mechanism of ETS factors rearranged in prostate cancer. Cell Rep. 17, 1289–1301 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Chong S., et al. , Tuning levels of low-complexity domain interactions to modulate endogenous oncogenic transcription. Mol. Cell 82, 2084–2097.e5 (2022), 10.1016/j.molcel.2022.04.007. [DOI] [PubMed] [Google Scholar]
  • 40.Boija A., et al. , Transcription factors activate genes through the phase-separation capacity of their activation domains. Cell 175, 1842–1855.e16 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Patel A., et al. , A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell 162, 1066–1077 (2015). [DOI] [PubMed] [Google Scholar]
  • 42.Emmanouilidis L., et al. , A solid beta-sheet structure is formed at the surface of FUS droplets during aging. Nat. Chem. Biol. 20, 1044–1052 (2024), 10.1038/s41589-024-01573-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Murray D. T., et al. , Structure of FUS protein fibrils and its relevance to self-assembly and phase separation of low-complexity domains. Cell 171, 615–627.e16 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Guo L., et al. , Nuclear-import receptors reverse aberrant phase transitions of RNA-binding proteins with prion-like domains. Cell 173, 677–692.e20 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Couthouis J., et al. , Evaluating the role of the FUS/TLS-related gene EWSR1 in amyotrophic lateral sclerosis. Hum. Mol. Genet. 21, 2899–2911 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Karim F. D., et al. , The ETS-domain: A new DNA-binding motif that recognizes a purine-rich core DNA sequence. Genes Dev. 4, 1451–1453 (1990). [DOI] [PubMed] [Google Scholar]
  • 47.Guillon N., et al. , The oncogenic EWS-FLI1 protein binds in vivo GGAA microsatellite sequences with potential transcriptional activation function. PLoS One 4, e4932 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Hou C., Mandal A., Rohr J., Tsodikov O. V., Allosteric interference in oncogenic FLI1 and ERG transactions by mithramycins. Structure 29, 404–412.e4 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Hou C., Tsodikov O. V., Structural basis for dimerization and DNA binding of transcription factor FLI1. Biochemistry 54, 7365–7374 (2015). [DOI] [PubMed] [Google Scholar]
  • 50.Kodandapani R., et al. , A new pattern for helix-turn-helix recognition revealed by the PU.1 ETS-domain-DNA complex. Nature 380, 456–460 (1996). [DOI] [PubMed] [Google Scholar]
  • 51.Regan M. C., et al. , Structural and dynamic studies of the transcription factor ERG reveal DNA binding is allosterically autoinhibited. Proc. Natl. Acad. Sci. U.S.A. 110, 13374–13379 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Cooper C. D., Newman J. A., Aitkenhead H., Allerston C. K., Gileadi O., Structures of the ETS protein DNA-binding domains of transcription factors Etv1, Etv4, Etv5, and Fev: Determinants of DNA binding and redox regulation by disulfide bond formation. J. Biol. Chem. 290, 13692–13709 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Liang H., et al. , The secondary structure of the ETS domain of human Fli-1 resembles that of the helix-turn-helix DNA-binding motif of the Escherichia coli catabolite gene activator protein. Proc. Natl. Acad. Sci. U.S.A. 91, 11655–11659 (1994). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Laudet V., Hanni C., Stehelin D., Duterque-Coquillaud M., Molecular phylogeny of the ETS gene family. Oncogene 18, 1351–1359 (1999). [DOI] [PubMed] [Google Scholar]
  • 55.Mao X., Miesfeldt S., Yang H., Leiden J. M., Thompson C. B., The FLI-1 and chimeric EWS-FLI-1 oncoproteins display similar DNA binding specificities. J. Biol. Chem. 269, 18216–18222 (1994). [PubMed] [Google Scholar]
  • 56.Johnson C. N., Xu X., Holloway S. P., Libich D. S., The 1H, 15N and 13C resonance assignments of the low-complexity domain from the oncogenic fusion protein EWS-FLI1. Biomol. NMR Assign. 16, 67–73 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Jumper J., et al. , Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Li H., et al. , Ewing sarcoma gene EWS is essential for meiosis and B lymphocyte development. J. Clin. Invest. 117, 1314–1323 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Paronetto M. P., Minana B., Valcarcel J., The Ewing sarcoma protein regulates DNA damage-induced alternative splicing. Mol. Cell 43, 353–368 (2011). [DOI] [PubMed] [Google Scholar]
  • 60.Knoop L. L., Baker S. J., The splicing factor U1C represses EWS/FLI-mediated transactivation. J. Biol. Chem. 275, 24865–24871 (2000). [DOI] [PubMed] [Google Scholar]
  • 61.Embree L. J., Azuma M., Hickstein D. D., Ewing sarcoma fusion protein EWSR1/FLI1 interacts with EWSR1 leading to mitotic defects in zebrafish embryos and human cell lines. Cancer Res. 69, 4363–4371 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Hallier M., Lerga A., Barnache S., Tavitian A., Moreau-Gachelin F., The transcription factor Spi-1/PU.1 interacts with the potential splicing factor TLS. J. Biol. Chem. 273, 4838–4842 (1998). [DOI] [PubMed] [Google Scholar]
  • 63.Delva L., et al. , Multiple functional domains of the oncoproteins Spi-1/PU.1 and TLS are involved in their opposite splicing effects in erythroleukemic cells. Oncogene 23, 4389–4399 (2004). [DOI] [PubMed] [Google Scholar]
  • 64.Kato M., McKnight S. L., Cross-beta polymerization of low complexity sequence domains. Cold Spring Harb. Perspect. Biol. 9, a023598 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Hughes M. P., et al. , Atomic structures of low-complexity protein segments reveal kinked beta sheets that assemble networks. Science 359, 698–701 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Lee M., Ghosh U., Thurber K. R., Kato M., Tycko R., Molecular structure and interactions within amyloid-like fibrils formed by a low-complexity protein sequence from FUS. Nat. Commun. 11, 5735 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Sawaya M. R., Hughes M. P., Rodriguez J. A., Riek R., Eisenberg D. S., The expanding amyloid family: Structure, stability, function, and pathogenesis. Cell 184, 4857–4873 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Ng K. P., et al. , Multiple aromatic side chains within a disordered structure are critical for transcription and transforming activity of EWS family oncoproteins. Proc. Natl. Acad. Sci. U.S.A. 104, 479–484 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Janke A. M., et al. , Lysines in the RNA polymerase II C-terminal domain contribute to TAF15 fibril recruitment. Biochemistry 57, 2549–2563 (2017), 10.1021/acs.biochem.7b00310. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Murthy A. C., et al. , Molecular interactions contributing to FUS SYGQ LC-RGG phase separation and co-partitioning with RNA polymerase II heptads. Nat. Struct. Mol. Biol. 28, 923–935 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Stsiapura V. I., et al. , Thioflavin T as a molecular rotor: Fluorescent properties of thioflavin T in solvents with different viscosity. J. Phys. Chem. B 112, 15893–15902 (2008). [DOI] [PubMed] [Google Scholar]
  • 72.Ryan J. J., Sprunger M. L., Holthaus K., Shorter J., Jackrel M. E., Engineered protein disaggregases mitigate toxicity of aberrant prion-like fusion proteins underlying sarcoma. J. Biol. Chem. 294, 11286–11296 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Dörner K., et al. , Tag with caution—How protein tagging influences the formation of condensates. bioRxiv [Preprint] (2024). 10.1101/2024.10.04.616694 (Accessed 20 March 2025). [DOI]
  • 74.Selvanathan S. P., et al. , Oncogenic fusion protein EWS-FLI1 is a network hub that regulates alternative splicing. Proc. Natl. Acad. Sci. U.S.A. 112, E1307–E1316 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Selig E. S., et al. , pAG8Ha-EWS-FLI1_Type2. Addgene. https://www.addgene.org/188042. Deposited 25 January 2023.
  • 76.Selig E. S., et al. , pAG8Ha-EWS. Addgene. https://www.addgene.org/188044. Deposited 25 January 2023.
  • 77.Selig E. S., et al. , pAG8Ha-EWS_RRM(359-447). Addgene. https://www.addgene.org/188045. Deposited 25 January 2023.
  • 78.Selig E. S., et al. , pAG8Ha-EWS_RRM-RGG(359-513). Addgene. https://www.addgene.org/188046. Deposited 25 January 2023.
  • 79.Selig E. S., et al. , pAG8Ha-FLI1_DBD(276-399_F362A). Addgene. https://www.addgene.org/188047. Deposited 25 January 2023.
  • 80.Selig E. S., et al. , pAG8Ha-FLI1_DBD(276-399_C299S_F362A_S390C). Addgene. https://www.addgene.org/188048. Deposited 25 January 2023.
  • 81.Selig E. S., et al. , pAG8Ha-FLI1_DBD_R2L2. Addgene. https://www.addgene.org/188049. Deposited 25 January 2023.
  • 82.Selig E. S., et al. , pAG8Ha-FLI1_DBD_deltaAlpha4. Addgene. https://www.addgene.org/188050. Deposited 25 January 2023.
  • 83.Selig E. S., et al. , pAG8Ha-ERG_DBD(306-429_F392A). Addgene. https://www.addgene.org/188051. Deposited 25 January 2023.
  • 84.Selig E. S., et al. , pAG8Ha-ETV1_DBD(332-458). Addgene. https://www.addgene.org/188052. Deposited 25 January 2023.
  • 85.Selig E. S., et al. , pAG8Ha-PU.1_DBD(168-270). Addgene. https://www.addgene.org/188053. Deposited 25 January 2023.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Appendix 01 (PDF)

Movie S1.

Dynamics of condensates formed by 25 μM EWSLCD in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 30 μm.

Download video file (9MB, mp4)
Movie S2.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM FLI1DBD in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 30 μm.

Download video file (8.6MB, mp4)
Movie S3.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM EWSRRM in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (9.4MB, mp4)
Movie S4.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM EWSRRMRGG2 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (8.9MB, mp4)
Movie S5.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM PU.1DBD in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10x magnification. Video speed is increased 6×. Scale bar = 30 μm.

Download video file (7.4MB, mp4)
Movie S6.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM EWSRGG3 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (10.6MB, mp4)
Movie S7.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM poly-L-lysine 30 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (8.9MB, mp4)
Movie S8.

Dynamics of condensates formed by 25 μM EWSLCD + 25 μM poly-L-lysine 50 in 100 mM NaCl, 20 mM sodium phosphate buffer pH 7.4. Images were collected at 10× magnification. Video speed is increased 6×. Scale bar = 20 μm.

Download video file (9MB, mp4)

Data Availability Statement

Expression plasmids data have been deposited in Addgene (188042188053) (7585).


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES