Skip to main content

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

  • Article
  • Published:

Shell protein composition specified by the lncRNA NEAT1 domains dictates the formation of paraspeckles as distinct membraneless organelles

Abstract

Many membraneless organelles (MLOs) formed through phase separation play crucial roles in various cellular processes. Although these MLOs co-exist in cells, how they maintain their independence without coalescence or engulfment remains largely unknown. Here, we investigated the molecular mechanism by which paraspeckles with core–shell architecture scaffolded by NEAT1_2 long noncoding RNAs exist as distinct MLOs. We identified NEAT1 deletion mutants that assemble paraspeckles that are incorporated into nuclear speckles. Several paraspeckle proteins, including SFPQ, HNRNPF and BRG1, prevent this incorporation and thus contribute to the segregation of paraspeckles from nuclear speckles. Shell localization of these proteins in the paraspeckles, which is determined by NEAT1_2 long noncoding RNA domains, is required for this segregation process. Conversely, U2-related spliceosomal proteins are involved in internalizing the paraspeckles into nuclear speckles. This study shows that the paraspeckle shell composition dictates the independence of MLOs in the nucleus, providing insights into the importance of the shell in defining features and functions of MLOs.

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

Access options

Buy this article

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

Fig. 1: Mini-PSs are incorporated into NSs.
Fig. 2: A subset of PSPs are mislocalized in mini-PSs.
Fig. 3: Artificial tethering of SFPQ, HNRNPF and BRG1 to the shells induces segregation of mini-PS from NSs.
Fig. 4: The dimerization, oligomerization and prion-like domains of SFPQ are required for the mini-PS segregation from NSs.
Fig. 5: U2 snRNP-related proteins are major interacting proteins of the mini-NEAT1 RNA domain responsible for incorporation into NSs.
Fig. 6: Artificial shell tethering of U2 snRNP components induces the internalization of PSs into NSs.

Similar content being viewed by others

Data availability

The MS data were deposited in the Japan Proteome Standard repository (jPOST) under the ID JPST002268. Source data are provided with this paper. All other data supporting the findings of this study are available from the corresponding author on reasonable request.

References

  1. Banani, S. F., Lee, H. O., Hyman, A. A. & Rosen, M. K. Biomolecular condensates: organizers of cellular biochemistry. Nat. Rev. Mol. Cell Biol. 18, 285–298 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Shin, Y. & Brangwynne, C. P. Liquid phase condensation in cell physiology and disease. Science 357, eaaf4382 (2017). (1979).

    PubMed  Google Scholar 

  3. Alberti, S., Gladfelter, A. & Mittag, T. Considerations and challenges in studying liquid-liquid phase separation and biomolecular condensates. Cell 176, 419–434 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Sabari, B. R., Dall’Agnese, A. & Young, R. A. Biomolecular condensates in the nucleus. Trends Biochem. Sci. 45, 961–977 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Lyon, A. S., Peeples, W. B. & Rosen, M. K. A framework for understanding the functions of biomolecular condensates across scales. Nat. Rev. Mol. Cell Biol. 22, 215–235 (2021).

    CAS  PubMed  Google Scholar 

  6. Yamazaki, T. & Hirose, T. Control of condensates dictates nucleolar architecture. Science 373, 486–487 (2021).

    CAS  PubMed  Google Scholar 

  7. Roden, C. & Gladfelter, A. S. RNA contributions to the form and function of biomolecular condensates. Nat. Rev. Mol. Cell Biol. 22, 183–195 (2021).

    CAS  PubMed  Google Scholar 

  8. Van Treeck, B. & Parker, R. Emerging roles for intermolecular RNA-RNA interactions in RNP assemblies. Cell 174, 791–802 (2018).

    PubMed  PubMed Central  Google Scholar 

  9. Yamazaki, T., Nakagawa, S. & Hirose, T. Architectural RNAs for membraneless nuclear body formation. Cold Spring Harb. Symp. Quant. Biol. 84, 227–237 (2019).

    PubMed  Google Scholar 

  10. Chujo, T. & Hirose, T. Nuclear bodies built on architectural long noncoding RNAs: unifying principles of their construction and function. Mol. Cells 40, 889–896 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Chen, L. L. & Carmichael, G. G. Altered nuclear retention of mRNAs containing inverted repeats in human embryonic stem cells: functional role of a nuclear noncoding RNA. Mol. Cell 35, 467–478 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Clemson, C. M. et al. An architectural role for a nuclear noncoding RNA: NEAT1 RNA is essential for the structure of paraspeckles. Mol. Cell 33, 717–726 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Sasaki, Y. T. F., Ideue, T., Sano, M., Mituyama, T. & Hirose, T. MENε/β noncoding RNAs are essential for structural integrity of nuclear paraspeckles. Proc. Natl Acad. Sci. USA 106, 2525–2530 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Sunwoo, H. et al. Men ε/β nuclear-retained non-coding RNAs are up-regulated upon muscle differentiation and are essential components of paraspeckles. Genome Res. 19, 347–359 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Fox, A. H. et al. Paraspeckles: a novel nuclear domain. Curr. Biol. 12, 13–25 (2002).

    CAS  PubMed  Google Scholar 

  16. Visa, N., Puvion-Dutilleul, F., Bachellerie, J. P. & Puvion, E. Intranuclear distribution of U1 and U2 snRNAs visualized by high resolution in situ hybridization: Revelation of a novel compartment containing U1 but not U2 snRNA in HeLa cells. Eur. J. Cell Biol. 60, 308–321 (1993).

    CAS  PubMed  Google Scholar 

  17. Hirose, T. et al. NEAT1 long noncoding RNA regulates transcription via protein sequestration within subnuclear bodies. Mol. Biol. Cell 25, 169–183 (2014).

    PubMed  PubMed Central  Google Scholar 

  18. Imamura, K. et al. Long noncoding RNA NEAT1-dependent SFPQ relocation from promoter region to paraspeckle mediates IL8 expression upon immune stimuli. Mol. Cell 53, 393–406 (2014).

    CAS  PubMed  Google Scholar 

  19. Nakagawa, S. et al. The lncRNA Neat1 is required for corpus luteum formation and the establishment of pregnancy in a subpopulation of mice. Development 141, 4618–4627 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Adriaens, C. et al. P53 induces formation of NEAT1 lncRNA-containing paraspeckles that modulate replication stress response and chemosensitivity. Nat. Med. 22, 861–868 (2016).

    CAS  PubMed  Google Scholar 

  21. Mello, S. S. et al. Neat1 is a p53-inducible lincRNA essential for transformation suppression. Genes Dev. 31, 1095–1108 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Naganuma, T. et al. Alternative 3′-end processing of long noncoding RNA initiates construction of nuclear paraspeckles. EMBO J. 31, 4020–4034 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Yamazaki, T. et al. Functional domains of NEAT1 architectural lncRNA induce paraspeckle assembly through phase separation. Mol. Cell 70, 1038–1053.e7 (2018).

    CAS  PubMed  Google Scholar 

  24. Wilusz, J. E. et al. A triple helix stabilizes the 3′ ends of long noncoding RNAs that lack poly(A) tails. Genes Dev. 26, 2392–2407 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Fox, A. H., Bond, C. S. & Lamond, A. I. P54nrb forms a heterodimer with PSP1 that localizes to paraspeckles in an RNA-dependent manner. Mol. Biol. Cell 16, 5304–5315 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Mao, Y. S., Sunwoo, H., Zhang, B. & Spector, D. L. Direct visualization of the co-transcriptional assembly of a nuclear body by noncoding RNAs. Nat. Cell Biol. 13, 95–101 (2011).

    CAS  PubMed  Google Scholar 

  27. Kawaguchi, T. et al. SWI/SNF chromatin-remodeling complexes function in noncoding RNA-dependent assembly of nuclear bodies. Proc. Natl Acad. Sci. USA 112, 4304–4309 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Yamazaki, T. & Hirose, T. The building process of the functional paraspeckle with long non-coding RNAs. Front. Biosci. 7, 715 (2015).

    Google Scholar 

  29. Hennig, S. et al. Prion-like domains in RNA binding proteins are essential for building subnuclear paraspeckles. J. Cell Biol. 210, 529–539 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Modic, M. et al. Cross-regulation between TDP-43 and paraspeckles promotes pluripotency-differentiation transition. Mol. Cell 74, 951–965.e13 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Yamazaki, T. et al. Paraspeckles are constructed as block copolymer micelles. EMBO J. 40, e107270 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Souquere, S., Beauclair, G., Harper, F., Fox, A. & Pierron, G. Highly ordered spatial organization of the structural long noncoding NEAT1 RNAs within paraspeckle nuclear bodies. Mol. Biol. Cell 21, 4020–4027 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. West, J. A. et al. Structural, super-resolution microscopy analysis of paraspeckle nuclear body organization. J. Cell Biol. 214, 817–830 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Yamazaki, T., Yamamoto, T. & Hirose, T. Micellization: a new principle in the formation of biomolecular condensates. Front Mol. Biosci. 9, 974772 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Chujo, T. et al. Unusual semi‐extractability as a hallmark of nuclear body‐associated architectural noncoding RNAs. EMBO J. 36, 1447–1462 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Huang, J. et al. Crystal structure of a SFPQ/PSPC1 heterodimer provides insights into preferential heterodimerization of human DBHS family proteins. J. Biol. Chem. 293, 6593–6602 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Lee, M. et al. The structure of human SFPQ reveals a coiled-coil mediated polymer essential for functional aggregation in gene regulation. Nucleic Acids Res. 43, 3826–3840 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Dopie, J., Sweredoski, M. J., Moradian, A. & Belmont, A. S. Tyramide signal amplification mass spectrometry (TSA-MS) ratio identifies nuclear speckle proteins. J. Cell Biol. 219, e201910207 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Yap, K., Chung, T. H. & Makeyev, E. V. Hybridization-proximity labeling reveals spatially ordered interactions of nuclear RNA compartments. Mol. Cell 82, 463–478.e11 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Sanders, D. W. et al. Competing protein-RNA interaction networks control multiphase intracellular organization. Cell 181, 306–324.e28 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Kaida, D. et al. Spliceostatin A targets SF3b and inhibits both splicing and nuclear retention of pre-mRNA. Nat. Chem. Biol. 3, 576–583 (2007).

    CAS  PubMed  Google Scholar 

  42. Kotake, Y. et al. Splicing factor SF3b as a target of the antitumor natural product pladienolide. Nat. Chem. Biol. 3, 570–575 (2007).

    CAS  PubMed  Google Scholar 

  43. Iida, K., Hagiwara, M. & Takeuchi, A. Multilateral bioinformatics analyses reveal the function-oriented target specificities and recognition of the RNA-binding protein SFPQ. iScience 23, 101325 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Dias, A. P., Dufu, K., Lei, H. & Reed, R. A role for TREX components in the release of spliced mRNA from nuclear speckle domains. Nat. Commun. 1, 97 (2010).

    PubMed  Google Scholar 

  45. Brown, J. A., Valenstein, M. L., Yario, T. A., Tycowski, K. T. & Steitz, J. A. Formation of triple-helical structures by the 3′-end sequences of MALAT1 and MENβ noncoding RNAs. Proc. Natl Acad. Sci. USA 109, 19202–19207 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Watanabe, Y. & Yamamoto, M. S. pombe mei2+ encodes an RNA-binding protein essential for premeiotic DNA synthesis and meiosis I, which cooperates with a novel RNA species meiRNA. Cell 78, 487–498 (1994).

    CAS  PubMed  Google Scholar 

  47. Prasanth, K. V., Rajendra, T. K., Lal, A. K. & Lakhotia, S. C. Omega speckles – a novel class of nuclear speckles containing hnRNPs associated with noncoding hsr-omega RNA in Drosophila. J. Cell Sci. 113, 3485–3497 (2000).

    CAS  PubMed  Google Scholar 

  48. Biamonti, G. & Vourc’h, C. Nuclear stress bodies. Cold Spring Harb. Perspect. Biol. 2, a000695 (2010).

    PubMed  PubMed Central  Google Scholar 

  49. Audas, T. E., Jacob, M. D. & Lee, S. Immobilization of proteins in the nucleolus by ribosomal intergenic spacer noncoding RNA. Mol. Cell 45, 147–157 (2012).

    CAS  PubMed  Google Scholar 

  50. Gordon, P. M., Hamid, F., Makeyev, E. V. & Houart, C. A conserved role for the ALS-linked splicing factor SFPQ in repression of pathogenic cryptic last exons. Nat. Commun. 12, 1918 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Stagsted, L. V. W., O’leary, E. T., Ebbesen, K. K. & Hansen, T. B. The RNA-binding protein SFPQ preserves long-intron splicing and regulates circRNA biogenesis in mammals. eLife 10, e63088 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Gañez-Zapater, A. et al. The SWI/SNF subunit BRG1 affects alternative splicing by changing RNA binding factor interactions with nascent RNA. Mol. Genet. Genomics 297, 463–484 (2022).

    PubMed  PubMed Central  Google Scholar 

  53. Huelga, S. C. et al. Integrative genome-wide analysis reveals cooperative regulation of alternative splicing by hnRNP proteins. Cell Rep. 1, 167–178 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Courchaine, E. M. et al. DMA-tudor interaction modules control the specificity of in vivo condensates. Cell 184, 3612–3625.e17 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Schmidt, U. et al. Real-time imaging of cotranscriptional splicing reveals a kinetic model that reduces noise: implications for alternative splicing regulation. J. Cell Biol. 193, 819–829 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Brody, Y. et al. The in vivo kinetics of RNA polymerase II elongation during co-transcriptional splicing. PLoS Biol. 9, e1000573 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Shen, S. et al. rMATS: robust and flexible detection of differential alternative splicing from replicate RNA-Seq data. Proc. Natl Acad. Sci. USA 111, E5593–E5601 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Kawaguchi, T. & Hirose, T. Chromatin remodeling complexes in the assembly of long noncoding RNA-dependent nuclear bodies. Nucleus 6, 462–467 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Yamazaki, T. & Hirose, T. CRISPR-mediated mutagenesis of long noncoding RNAs. Methods Mol. Biol. 2254, 283–303 (2021).

    CAS  PubMed  Google Scholar 

  60. Yamazaki, T. et al. FUS-SMN protein interactions link the motor neuron diseases ALS and SMA. Cell Rep. 2, 799–806 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Ilık, İ. A. et al. SON and SRRM2 are essential for nuclear speckle formation. eLife 9, e60579 (2020).

    PubMed  PubMed Central  Google Scholar 

  62. Lee, K. A. W. & Green, M. R. Small-scale preparation of extracts from radiolabeled cells efficient in pre-mRNA splicing. Methods Enzymol. 181, 20–30 (1990).

    CAS  PubMed  Google Scholar 

  63. Perez, C. A. G. et al. Sense-overlapping lncRNA as a decoy of translational repressor protein for dimorphic gene expression. PLoS Genet. 17, e1009683 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Ideue, T., Hino, K., Kitao, S., Yokoi, T. & Hirose, T. Efficient oligonucleotide-mediated degradation of nuclear noncoding RNAs in mammalian cultured cells. RNA 15, 1578–1587 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Wang, Y. et al. Genome-wide screening of NEAT1 regulators reveals cross-regulation between paraspeckles and mitochondria. Nat. Cell Biol. 20, 1145–1158 (2018).

    CAS  PubMed  Google Scholar 

  66. Gao, K., Masuda, A., Matsuura, T. & Ohno, K. Human branch point consensus sequence is yUnAy. Nucleic Acids Res. 36, 2257–2267 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank M. Okamura, C. Fujikawa and A. Kubota for technical support and A. Marshall, A. H. Fox, C. S. Bond, B. Dyakov, A. C. Gingras, T. Yoda, M. Murakami and members of the Hirose laboratory for valuable discussions. This research was supported by grants from the Ministry of Education, Culture, Sports, Science, and Technology (MEXT) of Japan (to T. Yamazaki (19K06479, 19H5250, 21H00253, 22H02545), to S.A. (21H05736, 22H02225), to T. Yamamoto (20H05934, 21H00241, 21K03479), to T.H. (20H00448, 20H05377, 21H05276, 22K19293)), the Mochida Memorial Foundation for Medical and Pharmaceutical Research (T. Yamazaki), the Naito Foundation (T. Yamazaki), the Takeda Science Foundation (T. Yamazaki), JST CREST (JPMJCR20E6) to T.H. and an AMED grant (JP21ae0121049) to T.H. The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

H.T., T. Yamazaki and T.H. conceived and designed this study. H.T. and T. Yamazaki conducted most of the experiments. S.S. and G.P. performed EM analyses. S.A. and T.N. performed MS analyses. N.F. contributed to RNA-seq data analysis and manuscript editing. T. Yamamoto contributed to theoretical interpretations of the experimental data. H.Y. and S.N. contributed to the SRM analyses. H.T., T. Yamazaki and T.H. wrote the manuscript.

Corresponding authors

Correspondence to Tomohiro Yamazaki or Tetsuro Hirose.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Eugene Makeyev and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Mini-PSs are present within NSs, although NEAT1 transcription occurs outside of NSs.

a, Schematics of the NEAT1_2 configuration in the WT-PS and mini-PS. b, SRM images of PSs (RNA-FISH with NEAT1_5′ probe) and NSs (SRRM2 IF) in WT and mini-NEAT1 mutant cells in the absence of MG132. Scale bar, 500 nm. c, Graph showing the proportion of segregated or incorporated PSs in b. Data were collected from two independent experiments. WT: n = 167, mini-NEAT1: n = 210. d, SRM images of the PSs (RNA-FISH with NEAT1_5′ probe) and NSs (SRRM2 IF) in WT and another mini-NEAT1 mutant clone treated with MG132 (5 μM for 6 h). Scale bar, 10 μm. e, PSs and NSs were detected by smFISH with NEAT1_5′ probes (magenta) and SON IF (green) in the presence of MG132 (5 μM for 6 h). Nuclei were stained with DAPI. Line profiles of the PSs and NSs are shown (right). f, SRM images of PSs (smFISH with NEAT1_5′ probe) and NSs (SON IF) in WT and mini-NEAT1 cells treated with MG132 (5 μM for 6 h). Scale bar, 500 nm. g, PSs and NSs in WT and mini-NEAT1 cells were detected by RNA-FISH with NEAT1_5′ probe and MALAT1 probe (green and magenta, respectively) in the presence of MG132 (5 μM for 6 h). Nuclei were stained with DAPI. Line profiles of the PSs and NSs are shown (right). Scale bar, 10 μm. h, SRM images of PSs (RNA-FISH with NEAT1_5′ probe) and NSs (RNA FISH with MALAT1 probe) in WT and mini-NEAT1 cells treated with MG132 (5 μM for 6 h). Scale bar, 500 nm. i, EM observation of multiple, but unconnected, mini-PSs within an NS. NONO localization was detected as gold particles. Dotted magenta circles indicate the mini-PS position. Dotted black circles indicate the NS position. Scale bar, 500 nm. j, PSs and other nuclear bodies in WT and mini-NEAT1 cells treated with MG132 (5 μM for 6 h). PSs were visualized by NEAT1 RNA-FISH with NEAT1_5′ probe (green). Nuclear bodies (parentheses) were visualized by IF of marker proteins (magenta). Nuclei were stained with DAPI. Scale bar, 10 μm. Source numerical data are available in source data.

Source data

Extended Data Fig. 2 A subset of PSPs are similarly localized in WT-PSs and mini-PSs.

a, Immunoblotting to examine the expression levels of various PSPs (BRG1, SFPQ, PSPC1, NONO, HNRNPH1, HNRNPF, and TDP-43) in HAP1 WT and the NEAT1 deletion mutant cells in Fig. 2a. Cells were treated with MG132 (5 μM for 6 h). GAPDH was used as a control. b, Confocal observation of the PSs and various PSPs in HAP1 WT and mini-NEAT1 mutant cells treated with MG132 (5 μM for 6 h). PSs and PSPs were visualized by RNA-FISH using NEAT1_5′ probes (green) and IF (magenta). Nuclei were stained with DAPI. Scale bar, 10 μm. c, Quantification of fluorescence intensity ratio (PSPs/NEAT1) in the HAP1 WT (blue) and mini-NEAT1 (pink) cells as observed in b (all samples: n = 30). Each box plot shows the median (inside line), 25th-75th percentiles (box bottom to top), and minimum-maximum values (whisker bottom to top). Data were compared using Mann-Whitney U-test (two-sided). If the statistical test showed no significant difference (P > 0.05), it is not specified in the figure. Source numerical data and unprocessed blots are available in source data.

Source data

Extended Data Fig. 3 Reciprocal recruitment of SFPQ, HNRNPF, and BRG1 by their tethering.

a, Genotyping PCR at the mini-NEAT1 locus. Electrophoresis of the PCR products confirmed the 6×MS2 knock-in. The primers used are listed in Supplementary Table 2. b, Immunoblotting of Flag-tagged MCP-PSPs in the mini-NEAT1 and mini-NEAT1/6×MS2@8.2 kb cells (Fig. 3b and 3c). GAPDH is a control. Molecular weight markers (MW) are shown on the right. c, Immunoblotting of Flag-tagged MCP-PSPs in the mini-NEAT1, mini-NEAT1/6×MS2@8.2 kb, and mini-NEAT1/6×MS2@16.2 kb (Fig. 3b–d). GAPDH is a control. MW are shown on the right. d, Quantification of fluorescence intensity ratio (FLAG/NEAT1) in three mutant cells shown (n = 20). eg, Quantification of fluorescence intensity ratio (PSPs/NEAT1) with transfection of MCP-PSPs into mini-NEAT1/6×MS2@8.2 kb cells. (e) SFPQ recruitment: n = 30 (MCP-GFPNLS), n = 30 (MCP-HNRNPF), n = 25 (MCP-BRG1), n = 30 (MCP-NONO), n = 30 (MCP-PSPC1). (f) HNRNPF recruitment: n = 30 (MCP-GFPNLS), n = 30 (MCP-SFPQ), n = 30 (MCP-BRG1). (g) BRG1 recruitment: n = 30 (MCP-GFPNLS), n = 30 (MCP-SFPQ), n = 30 (MCP-HNRNPF). h, PSs and NSs with transfection of MCP-NONO or PSPC1 into mini-NEAT1/6×MS2@8.2 kb (upper) and mini-NEAT1 (lower) treated with MG132 (5 μM for 6 h). White boxes indicate the areas shown at a higher magnification. Scale bar, 10 μm. i, Recruitment of various PSPs to PSs in WT and mini-NEAT1, and ∆4–8 kb/16.6–22.6 kb, and ∆1–4 kb/16.6–22.6 kb mutants treated with MG132 (5 μM for 6 h). PSs and PSPs were visualized by RNA-FISH using NEAT1_5′ probes (green) and IF (magenta). Nuclei were stained with DAPI. Scale bar, 10 μm. j, Quantification of fluorescence intensity ratio (PSPs/NEAT1) in the WT and four mutants shown as observed in i (all samples: n = 30). For dg,j, each box plot shows the median (inside line), 25th-75th percentiles (box bottom to top), and minimum-maximum values (whisker bottom to top). Data were compared using Kruskal-Wallis ANOVA and post hoc Dunn’s multiple comparison test. The statistical test showing no significant difference (P > 0.05) is not specified in the figure. Source numerical data and unprocessed blots are available in source data.

Source data

Extended Data Fig. 4 Effects of SFPQ deletion or mutations on interactions between SFPQ and the other DBHS family proteins and PS assembly.

a, Graph predicting PLD and IDR of SFPQ using PLAAC (upper) and PONDR (lower), respectively. b, Amino acid sequence of SFPQ. PLD is highlighted in red. c,d, Enrichment or depletion patterns of individual amino acids in SFPQ PLD detected by COMPOSITION PROFILER (refer to http://www.cprofiler.org). n = 100,000 bootstrap iterations. Data were compared using two-sample t-test (two-sided). e, Immunoblotting of Flag-tagged MCP-SFPQ WT and mutants in the mini-NEAT1 and mini-NEAT1/6×MS2@8.2 kb cells (Fig. 4b, c). GAPDH is a control. MW are shown on the right. f, PSs (NEAT1_5′), the NSs (SON), and SFPQ in mini-NEAT1/6×MS2@8.2 kb cells expressing MCP-SFPQ WT or ∆RRM2/NOPS and in HAP1 WT and mini-NEAT1 cells treated with MG132 (5 μM for 6 h). g, Quantification of fluorescence intensity ratio (SFPQ/NEAT1) as observed in f (n = 30). h, Proportion of the cells with PSs segregated from NSs in f. mini-NEAT1/6×MS2@8.2 kb: n = 58 (MCP-SFPQ WT), 53 (∆RRM2/NOPS). i, Co-IP of MCP-SFPQ WT and mutant proteins with HA-SFPQ. Immunoblotting was performed to detect HA-tagged SFPQ proteins in the co-IP samples. GAPDH is a negative control. j, Quantification of the data shown in i. Values (HA/FLAG) are mean ± SD of three independent experiments. Data were compared using Kruskal-Wallis test (P = 0.03). k, PS formation with transfection of MCP-SFPQ WT and mutant proteins into m13–16.6 kb/6 × MS2BS or m13–16.6 kb cells with MG132 treatment (5 μM for 6 h). The rescued PSs were visualized by NEAT1 RNA FISH (green) and IF of Flag-tagged MCP-SFPQ WT and mutants (magenta). l, PS sizes observed in k. m13-16.6 kb/6×MS2BS: n = 78 (WT), 54 (ΔRRM2/NOPS), 65 (ΔCC), 62 (ΔPLD), 78 (Q to G), 66 (P to A partial); m13-16.6k: n = 58 (WT), 64 (ΔRRM2/NOPS), 53 (ΔCC), 62 (ΔPLD), 60 (Q to G), 56 (P to A partial). g,l, Each box plot shows the median (inside line), 25th-75th percentiles (box bottom to top), and minimum-maximum values (whisker bottom to top). Data were compared using Kruskal-Wallis ANOVA and post hoc Dunn’s multiple comparison test. Scale bar, 10 μm (f,k). Source numerical data and unprocessed blots are available in source data.

Source data

Extended Data Fig. 5 U2 snRNP components are included in both NS and PS proteomes.

a, Confocal observation of PSs, BRG1, and SFPQ in HAP1 WT, mini-NEAT, and m9.8–16.6 kb mutant cells treated with MG132 (5 μM for 6 h). PSs were visualized by NEAT1 RNA-FISH with NEAT1_5′ probe (white). BRG1 (green) and SFPQ (magenta) were visualized by IF. Nuclei were stained with DAPI. Scale bar, 10 μm. b, Quantification of fluorescence intensity ratio (PSPs/NEAT1) observed in a (all samples: n = 30). Each box plot shows the median (inside line), 25th-75th percentiles (box bottom to top), and minimum-maximum values (whisker bottom to top). Data were compared using Kruskal-Wallis ANOVA and post hoc Dunn’s multiple comparison test. If the statistical test showed no significant difference (P > 0.05), it is not specified in the figure. c,d, Venn diagram showing the overlap between 8–9.2 kb and 9.2–9.8 kb of RNA pulldown (c), or between NEAT1 HyPro-MS39 and SC35 TSA-MS38 (d). e, Top 10 GO:CC enrichment for proteins shared between NEAT1 HyPro-MS39 and SC35 TSA-MS38. Source numerical data are available in source data.

Source data

Extended Data Fig. 6 SF3A and SF3B complex components bind to the 5′ terminal shell region of NEAT1, and the NEAT1_2 sequence contains multiple potential branch point sequences.

a, Genotyping PCR data are as in Extended data Fig. 3. b, SRM maximum projection images of PSs and NSs with transfection of MCP proteins into NEAT1/6×MS2@1.4 kb (left), NEAT1/6×MS2 @14 kb (middle), and WT (right) with MG132 treatment (5 μM for 6 h). White boxes indicate the areas shown at a higher magnification. Scale bar, 5 μm. c, Proportion of the cells with PSs segregated from NSs in b. NEAT1/6×MS2@1.4 kb @1.4 kb; n = 60 (SF3A3-MCP), 59 (MCP-PTBP1), NEAT1/6×MS2@1.4 kb @14 kb; n = 61 (SF3A3-MCP), 59 (MCP-PTBP1), WT; n = 62 (SF3A3-MCP), 64 (MCP-PTBP1). Data were compared using Fisher’s exact test (two-sided) and Bonferroni correction. d, Immunoblotting of Flag-tagged MCP proteins used in Fig. 6b, c and Extended Data Fig. 6b, c. GAPDH is a control. MW are shown on the right. e, SRM images of localization of SF3B1 within PSs in WT and mini-NEAT1 cells treated with MG132 (5 μM for 6 h) (upper). PSs were visualized by RNA FISH with NEAT1_5′ probe (green), and SF3B1 IF (magenta). Scale bar, 500 nm. Graph showing average fluorescence intensity of SF3B1 in the shell or core of PSs (lower) (n = 31). Each box plot shows the median (inside line), 25th-75th percentiles (box bottom to top), and minimum-maximum values (whisker bottom to top). Data were compared using Kruskal-Wallis ANOVA and post hoc Dunn’s multiple comparison test. f, U2 snRNA expression levels of the U2 ASO-transfected mini-NEAT1 mutant cells treated with MG132 (5 μM for 6 h) by RT-qPCR in Fig. 6f, g. Data are presented as mean ± SD (n = 3). Data were compared using one-way ANOVA and post hoc Tukey’s multiple comparison test. g, eCLIP data of SF3B4 (HepG2 and K562), SF3B1 (K562), SF3A3 (HepG2), U2AF2 (HepG2 and K562), SFPQ (HepG2), and NONO (K562) and HITS-CLIP data of HNRNPF (293 T)53 mapped on NEAT1_2. The number of SF1 consensus motifs (YNYURAY; Y = U or C, R = A or G, N = any nucleotide) present in the NEAT1_2 sequence is shown66. Source numerical data and unprocessed blots are available in source data.

Source data

Extended Data Fig. 7 Shell protein composition, but not core protein composition, of the WT-PS or mini-PS dictates their localization to NS.

A summary of the shell or core tethering in the WT-PS (upper) and mini-PS (lower) is depicted.

Supplementary information

Reporting Summary

Peer Review File

Supplementary Tables

Supplementary Tables 1–7.

Supplementary Video 1

3D visualization of PSs and NSs in HAP1 WT cells in the presence of MG132 (5 μM for 6 h), corresponding to Fig. 1e. PSs were visualized by NEAT1 RNA FISH with NEAT1_5′ probe (green), and NSs were visualized by SRRM2 IF (magenta).

Supplementary Video 2

3D visualization of PSs and NSs in mini-NEAT1 mutant cells in the presence of MG132 (5 μM for 6 h), corresponding to Fig. 1e. PSs were visualized by NEAT1 RNA FISH with NEAT1_5′ probe (green), and NSs were visualized by SRRM2 IF (magenta).

Source data

Figs. 1–6 and Extended Data Figs. 1–6

Statistical source data.

Figs. 2, 4 and Extended Data Figs. 2, 4, 6

Unprocessed western blots and/or gels.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Takakuwa, H., Yamazaki, T., Souquere, S. et al. Shell protein composition specified by the lncRNA NEAT1 domains dictates the formation of paraspeckles as distinct membraneless organelles. Nat Cell Biol 25, 1664–1675 (2023). https://doi.org/10.1038/s41556-023-01254-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-023-01254-1

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing