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:

A novel protein domain in an ancestral splicing factor drove the evolution of neural microexons

Abstract

The mechanisms by which entire programmes of gene regulation emerged during evolution are poorly understood. Neuronal microexons represent the most conserved class of alternative splicing in vertebrates, and are critical for proper brain development and function. Here, we discover neural microexon programmes in non-vertebrate species and trace their origin to bilaterian ancestors through the emergence of a previously uncharacterized ‘enhancer of microexons’ (eMIC) protein domain. The eMIC domain originated as an alternative, neural-enriched splice isoform of the pan-eukaryotic Srrm2/SRm300 splicing factor gene, and subsequently became fixed in the vertebrate and neuronal-specific splicing regulator Srrm4/nSR100 and its paralogue Srrm3. Remarkably, the eMIC domain is necessary and sufficient for microexon splicing, and functions by interacting with the earliest components required for exon recognition. The emergence of a novel domain with restricted expression in the nervous system thus resulted in the evolution of splicing programmes that qualitatively expanded the neuronal molecular complexity in bilaterians.

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: Neural microexon programmes in bilaterian animals.
Fig. 2: Evolution of the Srrm2/3/4 locus in metazoans.
Fig. 3: The eMIC domain is key for neural microexon splicing.
Fig. 4: The eMIC domain interacts with early spliceosomal factors to promote microexon inclusion.

Similar content being viewed by others

Code availability

All of the software used to analyse the data is publicly available and listed in the Reporting Summary. VastDB files to run vast-tools are available to download for each species (https://github.com/vastgroup/vast-tools), as indicated in the Methods. Custom codes to generate orthologous gene clusters and figure plots are available upon request.

Data availability

Raw RNA-Seq data were submitted to the Sequence Read Archive (SRP149913). Mass spectrometry data were submitted to the ProteomeXchange Consortium: AP-MS through MassIVE (https://massive.ucsd.edu; accession codes: MSV000082361 and PXD009779) and RNP mass spectrometry via PRIDE59 (PXD010034). All other RNA-Seq datasets used in the study are listed in Supplementary Dataset 3.

References

  1. Irimia, M. et al. A highly conserved program of neuronal microexons is misregulated in autistic brains. Cell 159, 1511–1523 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Li, Y. I., Sanchez-Pulido, L., Haerty, W. & Ponting, C. P. RBFOX and PTBP1 proteins regulate the alternative splicing of micro-exons in human brain transcripts. Genome Res. 25, 1–13 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Quesnel-Vallières, M., Irimia, M., Cordes, S. P. & Blencowe, B. J. Essential roles for the splicing regulator nSR100/SRRM4 during nervous system development. Genes Dev. 29, 746–759 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Nakano, Y. et al. A mutation in the Srrm4 gene causes alternative splicing defects and deafness in the Bronx waltzer mouse. PLoS Genet. 8, e1002966 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Quesnel-Vallieres, M. et al. Misregulation of an activity-dependent splicing network as a common mechanism underlying autism spectrum disorders. Mol. Cell 64, 1023–1034 (2016).

    Article  CAS  PubMed  Google Scholar 

  6. Calarco, J. A. et al. Regulation of vertebrate nervous system alternative splicing and development by an SR-related protein. Cell 138, 898–910 (2009).

    Article  CAS  PubMed  Google Scholar 

  7. Raj, B. et al. Global regulatory mechanism underlying the activation of an exon network required for neurogenesis. Mol. Cell 56, 90–103 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Parras, A. et al. Autism-like phenotype and risk gene mRNA deadenylation by CPEB4 mis-splicing. Nature 560, 441–446 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Gonatopoulos-Pournatzis, T. et al. Genome-wide CRISPR–Cas9 interrogation of splicing networks reveals a mechanism for recognition of autism-misregulated neuronal microexons. Mol. Cell 72, 510–524 (2018).

    Article  CAS  PubMed  Google Scholar 

  10. Putnam, N. et al. The amphioxus genome and the evolution of the chordate karyotype. Nature 453, 1064–1071 (2008).

    Article  CAS  PubMed  Google Scholar 

  11. Blencowe, B. J., Issner, R., Nickerson, J. A. & Sharp, P. A. A coactivator of pre-mRNA splicing. Genes Dev. 12, 996–1009 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Eldridge, A. G., Li, Y., Sharp, P. A. & Blencowe, B. J. The SRm160/300 splicing coactivator is required for exon-enhancer function. Proc. Natl Acad. Sci. USA 96, 6125–6130 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Khanna, M. et al. A systematic characterization of Cwc21, the yeast ortholog of the human spliceosomal protein SRm300. RNA 15, 2174–2185 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Grainger, R. J., Barrass, J. D., Jacquier, A., Rain, J. C. & Beggs, J. D. Physical and genetic interactions of yeast Cwc21p, an ortholog of human SRm300/SRRM2, suggest a role at the catalytic center of the spliceosome. RNA 15, 2161–2173 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Blencowe, B. J. et al. The SRm160/300 splicing coactivator subunits. RNA 6, 111–120 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Nakanishi, N., Renfer, E., Technau, U. & Rentzsch, F. Nervous systems of the sea anemone Nematostella vectensis are generated by ectoderm and endoderm and shaped by distinct mechanisms. Development 139, 347–357 (2012).

    Article  CAS  PubMed  Google Scholar 

  17. Wongpalee, S. P. et al. Large-scale remodeling of a repressed exon ribonucleoprotein to an exon definition complex active for splicing. eLife 5, e19743 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  18. McKeown, A. N. et al. Evolution of DNA specificity in a transcription factor family produced a new gene regulatory module. Cell 159, 58–68 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Logeman, B. L., Wood, L. K., Lee, J. & Thiele, D. J. Gene duplication and neo-functionalization in the evolutionary and functional divergence of the metazoan copper transporters Ctr1 and Ctr2. J. Biol. Chem. 292, 11531–11546 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Arnegard, M. E., Zwickl, D. J., Lu, Y. & Zakon, H. H. Old gene duplication facilitates origin and diversification of an innovative communication system—twice. Proc. Natl Acad. Sci. USA 107, 22172–22177 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Wang, J. et al. LSD1n is an H4K20 demethylase regulating memory formation via transcriptional elongation control. Nat. Neurosci. 18, 1256–1264 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Matsushita, M., Yamamoto, R., Mitsui, K. & Kanazawa, H. Altered motor activity of alternative splice variants of the mammalian kinesin-3 protein KIF1B. Traffic 10, 1647–1654 (2009).

    Article  CAS  PubMed  Google Scholar 

  23. Tapial, J. et al. An atlas of alternative splicing profiles and functional associations reveals new regulatory programs and genes that simultaneously express multiple major isoforms. Genome Res. 27, 1759–1768 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Braunschweig, U. et al. Widespread intron retention in mammals functionally tunes transcriptomes. Genome Res. 24, 1774–1786 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Giampietro, C. et al. The alternative splicing factor Nova2 regulates vascular development and lumen formation. Nat. Commun. 6, 8479 (2015).

    Article  CAS  PubMed  Google Scholar 

  26. Gueroussov, S. et al. An alternative splicing event amplifies evolutionary differences between vertebrates. Science 349, 868–873 (2015).

    Article  CAS  PubMed  Google Scholar 

  27. Han, H. et al. MBNL proteins repress ES-cell-specific alternative splicing and reprogramming. Nature 498, 241–245 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Solana, J. et al. Conserved functional antagonism of CELF and MBNL proteins controls stem cell-specific alternative splicing in planarians. eLife 5, e16797 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Burguera, D. et al. Evolutionary recruitment of flexible Esrp-dependent splicing programs into diverse embryonic morphogenetic processes. Nat. Commun. 8, 1799 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Altenhoff, A. M., Gil, M., Gonnet, G. H. & Dessimoz, C. Inferring hierarchical orthologous groups from orthologous gene pairs. PLoS ONE 8, e53786 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Alexeyenko, A., Tamas, I., Liu, G. & Sonnhammer, E. L. Automatic clustering of orthologs and inparalogs shared by multiple proteomes. Bioinformatics 22, e9–e15 (2006).

    Article  CAS  PubMed  Google Scholar 

  32. Singh, P. P., Arora, J. & Isambert, H. Identification of ohnolog genes originating from whole genome duplication in early vertebrates, based on synteny comparison across multiple genomes. PLoS Comput. Biol. 11, e1004394 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Katoh, K., Kuma, K., Toh, H. & Miyata, T. MAFFT version 5: improvement in accuracy of multiple sequence alignment. Nucleic Acids Res. 33, 511–518 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Irimia, M. & Roy, S. W. Spliceosomal introns as tools for genomic and evolutionary analysis. Nucleic Acids Res. 36, 1703–1712 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Yeo, G. W. & Burge, C. B. Maximum entropy modeling of short sequence motifs with applications to RNA splicing signals. J. Comput. Biol. 11, 377–394 (2004).

    Article  CAS  PubMed  Google Scholar 

  36. Corvelo, A., Hallegger, M., Smith, C. W. & Eyras, E. Genome-wide association between branch point properties and alternative splicing. PLoS Comput. Biol. 6, e1001016 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Gohr, A. & Irimia, M. Matt: Unix tools for alternative splicing analysis. Bioinformatics 35, 130–132 (2019).

    Article  PubMed  Google Scholar 

  38. Gonzalez, M. et al. Generation of stable Drosophila cell lines using multicistronic vectors. Sci. Rep. 1, 75 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Huelsenbeck, J. P. & Ronquist, F. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17, 754–755 (2001).

    Article  CAS  PubMed  Google Scholar 

  41. Müller, T. & Vingron, M. Modeling amino acid replacement. J. Comput. Biol. 7, 761–776 (2000).

    Article  PubMed  Google Scholar 

  42. Darriba, D., Taboada, G. L., Doallo, R. & Posada, D. ProtTest 3: fast selection of best-fit models of protein evolution. Bioinformatics 27, 1164–1165 (2011).

    Article  CAS  PubMed  Google Scholar 

  43. Birney, E., Clamp, M. & Durbin, R. Genewise and genomewise. Genome Res. 14, 988–995 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Yachdav, G. et al. MSAViewer: interactive JavaScript visualization of multiple sequence alignments. Bioinformatics 32, 3501–3503 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Wheeler, T. J., Clements, J. & Finn, R. D. Skylign: a tool for creating informative, interactive logos representing sequence alignments and profile hidden Markov models. BMC Bioinformatics 15, 7 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    Article  CAS  PubMed  Google Scholar 

  47. Sakamoto, H., Inoue, K., Higuchi, I., Ono, Y. & Shimura, Y. Control of Drosophila sex-lethal pre-mRNA splicing by its own female-specific product. Nucleic Acids Res. 20, 5533–5540 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Fritzenwanker, J. H. & Technau, U. Induction of gametogenesis in the basal cnidarian Nematostella vectensis (Anthozoa). Dev. Genes Evol. 212, 99–103 (2002).

    Article  PubMed  Google Scholar 

  49. Lambert, J. P., Tucholska, M., Go, C., Knight, J. D. & Gingras, A. C. Proximity biotinylation and affinity purification are complementary approaches for the interactome mapping of chromatin-associated protein complexes. J. Proteomics 118, 81–94 (2015).

    Article  CAS  PubMed  Google Scholar 

  50. Youn, J. Y. et al. High-density proximity mapping reveals the subcellular organization of mRNA-associated granules and bodies. Mol. Cell 69, 517–532 (2018).

    Article  CAS  PubMed  Google Scholar 

  51. Liu, G. et al. Data independent acquisition analysis in ProHits 4.0. J. Proteomics 149, 64–68 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Deutsch, E. W. et al. A guided tour of the trans-proteomic pipeline. Proteomics 10, 1150–1159 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Shteynberg, D. et al. iProphet: multi-level integrative analysis of shotgun proteomic data improves peptide and protein identification rates and error estimates. Mol. Cell. Proteomics 10, M111.007690 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Teo, G. et al. SAINTexpress: improvements and additional features in Significance Analysis of INTeractome software. J. Proteomics 100, 37–43 (2014).

    Article  CAS  PubMed  Google Scholar 

  55. Mackereth, C. D. et al. Multi-domain conformational selection underlies pre-mRNA splicing regulation by U2AF. Nature 475, 408–411 (2011).

    Article  CAS  PubMed  Google Scholar 

  56. Blencowe, B. J. & Lamond, A. I. Purification and depletion of RNP particles by antisense affinity chromatography. Methods Mol. Biol. 118, 275–287 (1999).

    CAS  PubMed  Google Scholar 

  57. Barabino, S. M., Blencowe, B. J., Ryder, U., Sproat, B. S. & Lamond, A. I. Targeted snRNP depletion reveals an additional role for mammalian U1 snRNP in spliceosome assembly. Cell 63, 293–302 (1990).

    Article  CAS  PubMed  Google Scholar 

  58. Dignam, J. D., Lebovitz, R. M. & Roeder, R. G. Accurate transcription initiation by RNA polymerase II in a soluble extract from isolated mammalian nuclei. Nucleic Acids Res. 11, 1475–1489 (1983).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Vizcaino, J. A. et al. ProteomeXchange provides globally coordinated proteomics data submission and dissemination. Nat. Biotechnol. 32, 223–226 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank B. Lehner and S. W. Roy for critical reading of the manuscript, M. Akam and K. Siggens for providing access to Strigamia specimens, M. Sattler and P. Zou for pETM11 clones, S. Taylor for providing the HeLa Flp-In cell line, B. Bergum (Flow Cytometry Core Facility at the University of Bergen) for assistance with Nematostella fluorescence-activated cell sorting, and the CRG Genomics Unit. Animal silhouettes were obtained from PhyloPic. This work has been funded by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (ERC-StG-LS2-637591 to M.Ir. and ERC-AdvG-670146 to J.V.), the Spanish Ministry of Economy and Competitiveness (BFU2014-55076-P and BFU2017-89201-P to M.Ir., BFU2014-005153 to J.V., and the ‘Centro de Excelencia Severo Ochoa 2013–2017’ (SEV-2012-0208)), AGAUR, Fundación Botín (to J.V.) and the Canadian Institutes of Health Research (to B.J.B. and A.-C.G.). RNP mass spectrometric analyses were performed at the CRG/UPF Proteomics Unit (part of ProteoRed-PRB3, supported by PE I+D+i 2013–2016 (PT17/0019) of the ISCIII and ERDF) by ‘Programa CERCA Generalitat de Catalunya’ and ‘Secretaria d’Universitats i Recerca del Departament d’Economia i Coneixement’ (2017SGR595). A.T.-M. held an FPI-SO fellowship, and Y.M. a Marie Skłodowska-Curie individual fellowship.

Author information

Authors and Affiliations

Authors

Contributions

A.T-M. performed the bioinformatic analyses and molecular biology and cell culture experiments. S.B. performed the cell culture, spliceosome assembly and RNP mass spectrometry experiments under the supervision of J.V. Y.M. performed the bioinformatics analyses. J.R. performed and analysed the AP-MS experiments under the supervision of A-C.G. and B.J.B. M.Ig. performed the Nematostella experiments under the supervision of F.R. J.P. performed the Strigamia dissections and molecular biology work. D.O. performed the co-immunoprecipitation experiments under the supervision of B.J.B. T.G., I.A., F.G. and M.S. performed the Drosophila and molecular biology experiments. M.Ir. conceived and supervised the study, provided resources and performed the bioinformatic analyses. A.T-M., S.B., M.Ir. and B.J.B. wrote the manuscript with input from all authors.

Corresponding author

Correspondence to Manuel Irimia.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Supplementary information

Supplementary Information

Supplementary Figures 1–7

Reporting Summary

Supplementary Data 1

Conserved microexons across phyla

Supplementary Data 2

Mass spectrometry results

Supplementary Data 3

RNA-seq data used in this study

Supplementary Data 4

Orthologous genes in the six bilaterian species studied

Supplementary Data 5

Srrm protein sequences used for comparative analyses

Supplementary Data 6

List of primer sequences used in this study

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Torres-Méndez, A., Bonnal, S., Marquez, Y. et al. A novel protein domain in an ancestral splicing factor drove the evolution of neural microexons. Nat Ecol Evol 3, 691–701 (2019). https://doi.org/10.1038/s41559-019-0813-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41559-019-0813-6

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