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:

Whole-genome sequencing reveals that variants in the Interleukin 18 Receptor Accessory Protein 3′UTR protect against ALS

Abstract

The noncoding genome is substantially larger than the protein-coding genome but has been largely unexplored by genetic association studies. Here, we performed region-based rare variant association analysis of >25,000 variants in untranslated regions of 6,139 amyotrophic lateral sclerosis (ALS) whole genomes and the whole genomes of 70,403 non-ALS controls. We identified interleukin-18 receptor accessory protein (IL18RAP) 3′ untranslated region (3′UTR) variants as significantly enriched in non-ALS genomes and associated with a fivefold reduced risk of developing ALS, and this was replicated in an independent cohort. These variants in the IL18RAP 3′UTR reduce mRNA stability and the binding of double-stranded RNA (dsRNA)-binding proteins. Finally, the variants of the IL18RAP 3′UTR confer a survival advantage for motor neurons because they dampen neurotoxicity of human induced pluripotent stem cell (iPSC)-derived microglia bearing an ALS-associated expansion in C9orf72, and this depends on NF-κB signaling. This study reveals genetic variants that protect against ALS by reducing neuroinflammation and emphasizes the importance of noncoding genetic association studies.

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: Region-based rare variant association analysis reveals association of the IL18RAP 3′UTR with ALS.
Fig. 2: Odds of ALS are reduced with rare variants in the IL18RAP 3′UTR.
Fig. 3: IL18RAP 3′UTR variants correlate with attenuated IL-18–NF-κB signaling in human lymphoblastoid cells.
Fig. 4: IL18RAP 3′UTR variants destabilize IL18RAP mRNA in CRISPR-edited isogenic iPSC-derived microglia with a C9orf72 genetic background.
Fig. 5: Reduced association of dsRNA-binding proteins to the variant IL18RAP 3′UTR.
Fig. 6: Variant IL18RAP 3′UTR is protective in human microglia and in individuals with ALS.
Fig. 7: Variant IL18RAP 3′UTR dampens neurotoxic NF-κB signaling in human microglia.

Similar content being viewed by others

Data availability

Human genetics data are publically available from the sequencing consortia that control ethically appropriate usage of data, harmonization across studies and the safety of personal information donated by individuals that contributed their DNA for sequencing: the Project Mine ALS sequencing consortium, the NYGC ALS Consortium, the gnomAD and NHLBI’s TOPMed. Sequencing data are deposited at Gene Expression Omnibus under accession number GSE186757. All Other data used for this manuscript are available in the manuscript. The University of California Santa Cruz gene annotation93, miRBase v20 (ref. 57), RefSeq63, dbNSFP v2.0 (ref. 58) and ANNOVAR95 databases were used in this study. Source data are provided with this paper.

Code availability

Variant annotation scripts are available at GitHub at https://github.com/TsviyaOlender/Non-coding-Variants-in-ALS-genes-.

References

  1. Brown, R. H. & Al-Chalabi, A. Amyotrophic lateral sclerosis. N. Engl. J. Med. 377, 162–172 (2017).

    Article  CAS  PubMed  Google Scholar 

  2. Taylor, J. P., Brown, R. H. Jr. & Cleveland, D. W. Decoding ALS: from genes to mechanism. Nature 539, 197–206 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Renton, A. E., Chio, A. & Traynor, B. J. State of play in amyotrophic lateral sclerosis genetics. Nat. Neurosci. 17, 17–23 (2014).

    Article  CAS  PubMed  Google Scholar 

  4. Al-Chalabi, A., van den Berg, L. H. & Veldink, J. Gene discovery in amyotrophic lateral sclerosis: implications for clinical management. Nat. Rev. Neurol. 13, 96–104 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. van Rheenen, W. et al. Common and rare variant association analyses in amyotrophic lateral sclerosis identify 15 risk loci with distinct genetic architectures and neuron-specific biology. Nat. Genet. 53, 1636–1648 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  6. DeJesus-Hernandez, M. et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron 72, 245–256 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Renton, A. E. et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron 72, 257–268 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. La Spada, A. R. & Taylor, J. P. Repeat expansion disease: progress and puzzles in disease pathogenesis. Nat. Rev. Genet. 11, 247–258 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Cooper-Knock, J. et al. Rare variant burden analysis within enhancers identifies CAV1 as an ALS risk gene. Cell Rep. 33, 108456 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Povysil, G. et al. Rare-variant collapsing analyses for complex traits: guidelines and applications. Nat. Rev. Genet. 20, 747–759 (2019).

    Article  CAS  PubMed  Google Scholar 

  11. Cookson, W., Liang, L., Abecasis, G., Moffatt, M. & Lathrop, M. Mapping complex disease traits with global gene expression. Nat. Rev. Genet. 10, 184–194 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Knight, J. C. Regulatory polymorphisms underlying complex disease traits. J. Mol. Med. 83, 97–109 (2005).

    Article  CAS  PubMed  Google Scholar 

  13. An, J.Y., et al. Genome-wide de novo risk score implicates promoter variation in autism spectrum disorder. Science 362, eaat6576 (2018).

  14. Haramati, S. et al. miRNA malfunction causes spinal motor neuron disease. Proc. Natl Acad. Sci. USA 107, 13111–13116 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Emde, A. et al. Dysregulated miRNA biogenesis downstream of cellular stress and ALS-causing mutations: a new mechanism for ALS. EMBO J. 34, 2633–2651 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Eitan, C. & Hornstein, E. Vulnerability of microRNA biogenesis in FTD-ALS. Brain Res. 1647, 105–111 (2016).

  17. Campos-Melo, D., Droppelmann, C. A., He, Z., Volkening, K. & Strong, M. J. Altered microRNA expression profile in amyotrophic lateral sclerosis: a role in the regulation of NFL mRNA levels. Mol. Brain 6, 26 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Buratti, E. et al. Nuclear factor TDP-43 can affect selected microRNA levels. FEBS J. 277, 2268–2281 (2010).

    Article  CAS  PubMed  Google Scholar 

  19. Kawahara, Y. & Mieda-Sato, A. TDP-43 promotes microRNA biogenesis as a component of the Drosha and Dicer complexes. Proc. Natl Acad. Sci. USA 109, 3347–3352 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Morlando, M. et al. FUS stimulates microRNA biogenesis by facilitating co-transcriptional Drosha recruitment. EMBO J. 31, 4502–4510 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Hoye, M. L. et al. MicroRNA profiling reveals marker of motor neuron disease in ALS models. J. Neurosci. 37, 5574–5586 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Rotem, N. et al. ALS along the axons—expression of coding and noncoding RNA differs in axons of ALS models. Sci Rep. 7, 44500 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Butovsky, O. et al. Modulating inflammatory monocytes with a unique microRNA gene signature ameliorates murine ALS. J. Clin. Invest. 122, 3063–3087 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Figueroa-Romero, C. et al. Expression of microRNAs in human post-mortem amyotrophic lateral sclerosis spinal cords provides insight into disease mechanisms. Mol. Cell. Neurosci. 71, 34–45 (2016).

    Article  CAS  PubMed  Google Scholar 

  25. Williams, A. H. et al. MicroRNA-206 delays ALS progression and promotes regeneration of neuromuscular synapses in mice. Science 326, 1549–1554 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bartel, D. P. MicroRNAs: target recognition and regulatory functions. Cell 136, 215–233 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Mayr, C. Regulation by 3′-untranslated regions. Annu. Rev. Genet. 51, 171–194 (2017).

    Article  CAS  PubMed  Google Scholar 

  28. Lee, S., Abecasis, G. R., Boehnke, M. & Lin, X. Rare-variant association analysis: study designs and statistical tests. Am. J. Hum. Genet. 95, 5–23 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Alboni, S., Cervia, D., Sugama, S. & Conti, B. Interleukin 18 in the CNS. J. Neuroinflammation 7, 9 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Zhao, W. et al. TDP-43 activates microglia through NF-κB and NLRP3 inflammasome. Exp. Neurol. 273, 24–35 (2015).

    Article  CAS  PubMed  Google Scholar 

  31. Tsutsumi, N. et al. The structural basis for receptor recognition of human interleukin-18. Nat. Commun. 5, 5340 (2014).

    Article  CAS  PubMed  Google Scholar 

  32. Adachi, O. et al. Targeted disruption of the MyD88 gene results in loss of IL-1- and IL-18-mediated function. Immunity 9, 143–150 (1998).

    Article  CAS  PubMed  Google Scholar 

  33. Kato, Z. et al. The structure and binding mode of interleukin-18. Nat. Struct. Biol. 10, 966–971 (2003).

    Article  CAS  PubMed  Google Scholar 

  34. Matsumoto, S. et al. Interleukin-18 activates NF-κB in murine T helper type 1 cells. Biochem. Biophys. Res. Commun. 234, 454–457 (1997).

    Article  CAS  PubMed  Google Scholar 

  35. Robinson, D. et al. IGIF does not drive Th1 development but synergizes with IL-12 for interferon-γ production and activates IRAK and NFκB. Immunity 7, 571–581 (1997).

    Article  CAS  PubMed  Google Scholar 

  36. Kojima, H. et al. An essential role for NF-κB in IL-18-induced IFN-γ expression in KG-1 cells. J Immunol. 162, 5063–5069 (1999).

    CAS  PubMed  Google Scholar 

  37. Morel, J. C., Park, C. C., Kumar, P. & Koch, A. E. Interleukin-18 induces rheumatoid arthritis synovial fibroblast CXC chemokine production through NFκB activation. Lab. Invest. 81, 1371–1383 (2001).

    Article  CAS  PubMed  Google Scholar 

  38. Miyoshi, K., Obata, K., Kondo, T., Okamura, H. & Noguchi, K. Interleukin-18-mediated microglia/astrocyte interaction in the spinal cord enhances neuropathic pain processing after nerve injury. J. Neurosci. 28, 12775–12787 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Kadhim, H., Deltenre, P., Martin, J. J. & Sebire, G. In-situ expression of interleukin-18 and associated mediators in the human brain of sALS patients: hypothesis for a role for immune–inflammatory mechanisms. Med. Hypotheses 86, 14–17 (2016).

    Article  CAS  PubMed  Google Scholar 

  40. Johann, S. et al. NLRP3 inflammasome is expressed by astrocytes in the SOD1 mouse model of ALS and in human sporadic ALS patients. Glia 63, 2260–2273 (2015).

    Article  PubMed  Google Scholar 

  41. Italiani, P. et al. Evaluating the levels of interleukin-1 family cytokines in sporadic amyotrophic lateral sclerosis. J. Neuroinflammation 11, 94 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Huang, F. et al. Longitudinal biomarkers in amyotrophic lateral sclerosis. Ann. Clin. Transl Neurol. 7, 1103–1116 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lall, D. & Baloh, R. H. Microglia and C9orf72 in neuroinflammation and ALS and frontotemporal dementia. J. Clin. Invest. 127, 3250–3258 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Beers, D. R. & Appel, S. H. Immune dysregulation in amyotrophic lateral sclerosis: mechanisms and emerging therapies. Lancet Neurol. 18, 211–220 (2019).

    Article  CAS  PubMed  Google Scholar 

  45. Vahsen, B. F. et al. Non-neuronal cells in amyotrophic lateral sclerosis—from pathogenesis to biomarkers. Nat. Rev. Neurol. 17, 333–348 (2021).

    Article  PubMed  Google Scholar 

  46. McCauley, M. E. & Baloh, R. H. Inflammation in ALS/FTD pathogenesis. Acta Neuropathol. 137, 715–730 (2019).

    Article  CAS  PubMed  Google Scholar 

  47. Motataianu, A., Barcutean, L. & Balasa, R. Neuroimmunity in amyotrophic lateral sclerosis: focus on microglia. Amyotroph. Lateral Scler. Frontotemporal Degener. 21, 159–166 (2020).

    Article  CAS  PubMed  Google Scholar 

  48. Philips, T. & Robberecht, W. Neuroinflammation in amyotrophic lateral sclerosis: role of glial activation in motor neuron disease. Lancet Neurol. 10, 253–263 (2011).

    Article  CAS  PubMed  Google Scholar 

  49. Kaltschmidt, B. & Kaltschmidt, C. NF-κB in the nervous system. Cold Spring Harb. Perspect. Biol. 1, a001271 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Mattson, M. P. & Meffert, M. K. Roles for NF-κB in nerve cell survival, plasticity, and disease. Cell Death Differ. 13, 852–860 (2006).

    Article  CAS  PubMed  Google Scholar 

  51. Frakes, A. E. et al. Microglia induce motor neuron death via the classical NF-κB pathway in amyotrophic lateral sclerosis. Neuron 81, 1009–1023 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Uranishi, H. et al. Involvement of the pro-oncoprotein TLS (translocated in liposarcoma) in nuclear factor-κB p65-mediated transcription as a coactivator. J. Biol. Chem. 276, 13395–13401 (2001).

    Article  CAS  PubMed  Google Scholar 

  53. Swarup, V. et al. Deregulation of TDP-43 in amyotrophic lateral sclerosis triggers nuclear factor κB-mediated pathogenic pathways. J. Exp. Med. 208, 2429–2447 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Project MinE Consortium. Project MinE: study design and pilot analyses of a large-scale whole-genome sequencing study in amyotrophic lateral sclerosis. Eur. J. Hum. Genet. 26, 1537–1546 (2017).

  55. Lee, S. et al. Optimal unified approach for rare-variant association testing with application to small-sample case–control whole-exome sequencing studies. Am. J. Hum. Genet. 91, 224–237 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Dunckley, T. et al. Whole-genome analysis of sporadic amyotrophic lateral sclerosis. N. Engl. J. Med. 357, 775–788 (2007).

    Article  CAS  PubMed  Google Scholar 

  57. Griffiths-Jones, S., Grocock, R. J., van Dongen, S., Bateman, A. & Enright, A. J. miRBase: microRNA sequences, targets and gene nomenclature. Nucleic Acids Res. 34, D140–D144 (2006).

    Article  CAS  PubMed  Google Scholar 

  58. Liu, X., Jian, X. & Boerwinkle, E. dbNSFP v2.0: a database of human non-synonymous SNVs and their functional predictions and annotations. Hum. Mutat. 34, E2393–E2402 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Kenna, K. P. et al. NEK1 variants confer susceptibility to amyotrophic lateral sclerosis. Nat. Genet. 48, 1037–1042 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Rosen, D. R. et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 362, 59–62 (1993).

    Article  CAS  PubMed  Google Scholar 

  61. Chio, A. et al. Prevalence of SOD1 mutations in the Italian ALS population. Neurology 70, 533–537 (2008).

    Article  CAS  PubMed  Google Scholar 

  62. van der Spek, R. A. A. et al. The project MinE databrowser: bringing large-scale whole-genome sequencing in ALS to researchers and the public. Amyotroph. Lateral Scler. Frontotemporal Degener. 20, 432–440 (2019).

    Article  PubMed  Google Scholar 

  63. O’Leary, N. A. et al. Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation. Nucleic Acids Res. 44, D733–D745 (2016).

    Article  PubMed  Google Scholar 

  64. Smith, L. et al. Establishing the UK DNA Bank for motor neuron disease (MND). BMC Genet. 16, 84 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Shi, Y. et al. Haploinsufficiency leads to neurodegeneration in C9ORF72 ALS/FTD human induced motor neurons. Nat. Med. 24, 313–325 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Haukedal, H. & Freude, K. Implications of microglia in amyotrophic lateral sclerosis and frontotemporal dementia. J. Mol. Biol. 431, 1818–1829 (2019).

    Article  CAS  PubMed  Google Scholar 

  67. Haenseler, W. et al. A highly efficient human pluripotent stem cell microglia model displays a neuronal-co-culture-specific expression profile and inflammatory response. Stem Cell Reports 8, 1727–1742 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Peng, S. S., Chen, C. Y., Xu, N. & Shyu, A. B. RNA stabilization by the AU-rich element binding protein, HuR, an ELAV protein. EMBO J. 17, 3461–3470 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Fan, X. C. & Steitz, J. A. Overexpression of HuR, a nuclear-cytoplasmic shuttling protein, increases the in vivo stability of ARE-containing mRNAs. EMBO J. 17, 3448–3460 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Stellos, K. et al. Adenosine-to-inosine RNA editing controls cathepsin S expression in atherosclerosis by enabling HuR-mediated post-transcriptional regulation. Nat. Med. 22, 1140–1150 (2016).

    Article  CAS  PubMed  Google Scholar 

  71. Brennan, C. M. & Steitz, J. A. HuR and mRNA stability. Cell. Mol. Life Sci. 58, 266–277 (2001).

    Article  CAS  PubMed  Google Scholar 

  72. Garcia-Dominguez, D. J., Morello, D., Cisneros, E., Kontoyiannis, D. L. & Frade, J. M. Stabilization of Dll1 mRNA by Elavl1/HuR in neuroepithelial cells undergoing mitosis. Mol. Biol. Cell 22, 1227–1239 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Rothamel, K. et al. ELAVL1 primarily couples mRNA stability with the 3′ UTRs of interferon-stimulated genes. Cell Rep. 35, 109178 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Mukherjee, N. et al. Integrative regulatory mapping indicates that the RNA-binding protein HuR couples pre-mRNA processing and mRNA stability. Mol. Cell 43, 327–339 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Fernandopulle, M. S. et al. Transcription factor-mediated differentiation of human iPSCs into neurons. Curr. Protoc. Cell Biol. 79, e51 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Christian, F., Smith, E.L. & Carmody, R.J. The regulation of NF-κB subunits by phosphorylation. Cells 5, 12 (2016).

  77. Zhong, H., May, M. J., Jimi, E. & Ghosh, S. The phosphorylation status of nuclear NF-κB determines its association with CBP/p300 or HDAC-1. Mol. Cell 9, 625–636 (2002).

    Article  CAS  PubMed  Google Scholar 

  78. Zhong, H., Voll, R. E. & Ghosh, S. Phosphorylation of NF-κB p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with the coactivator CBP/p300. Mol. Cell 1, 661–671 (1998).

    Article  CAS  PubMed  Google Scholar 

  79. Oeckinghaus, A. & Ghosh, S. The NF-κB family of transcription factors and its regulation. Cold Spring Harb. Perspect. Biol. 1, a000034 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Waelchli, R. et al. Design and preparation of 2-benzamido-pyrimidines as inhibitors of IKK. Bioorg. Med. Chem. Lett. 16, 108–112 (2006).

    Article  CAS  PubMed  Google Scholar 

  81. Ayers, K. L. et al. A loss of function variant in CASP7 protects against Alzheimer’s disease in homozygous APOE ε4 allele carriers. BMC Genomics 17, 445 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Benitez, B. A. et al. Missense variant in TREML2 protects against Alzheimer’s disease. Neurobiol. Aging 35, 1510.e19-26 (2014).

    Article  PubMed  Google Scholar 

  83. Jonsson, T. et al. A mutation in APP protects against Alzheimer’s disease and age-related cognitive decline. Nature 488, 96–99 (2012).

    Article  CAS  PubMed  Google Scholar 

  84. Sims, R., et al. Rare coding variants in PLCG2, ABI3, and TREM2 implicate microglial-mediated innate immunity in Alzheimer’s disease. Nat. Genet. 49, 1373–1384 (2017).

  85. Landers, J. E. et al. Reduced expression of the kinesin-associated protein 3 (KIFAP3) gene increases survival in sporadic amyotrophic lateral sclerosis. Proc. Natl Acad. Sci. USA 106, 9004–9009 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Farhan, S. M. K. et al. Exome sequencing in amyotrophic lateral sclerosis implicates a novel gene, DNAJC7, encoding a heat-shock protein. Nat. Neurosci. 22, 1966–1974 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Lambrechts, D. et al. VEGF is a modifier of amyotrophic lateral sclerosis in mice and humans and protects motoneurons against ischemic death. Nat. Genet. 34, 383–394 (2003).

    Article  CAS  PubMed  Google Scholar 

  88. Reichenstein, I., et al. Human genetics and neuropathology suggest a link between miR-218 and amyotrophic lateral sclerosis pathophysiology. Sci Transl Med 11, eaav5264 (2019).

  89. Liao, Y., Wang, J., Jaehnig, E. J., Shi, Z. & Zhang, B. WebGestalt 2019: gene set analysis toolkit with revamped UIs and APIs. Nucleic Acids Res. 47, W199–W205 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Lorenz, R. et al. ViennaRNA package 2.0. Algorithms Mol. Biol. 6, 26 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  91. Zhan, X., Hu, Y., Li, B., Abecasis, G. R. & Liu, D. J. RVTESTS: an efficient and comprehensive tool for rare variant association analysis using sequence data. Bioinformatics 32, 1423–1426 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Raczy, C. et al. Isaac: ultra-fast whole-genome secondary analysis on Illumina sequencing platforms. Bioinformatics 29, 2041–2043 (2013).

    Article  CAS  PubMed  Google Scholar 

  93. Tyner, C. et al. The UCSC Genome Browser database: 2017 update. Nucleic Acids Res. 45, D626–D634 (2017).

    CAS  PubMed  Google Scholar 

  94. Danecek, P. et al. The variant call format and VCFtools. Bioinformatics 27, 2156–2158 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Wang, K., Li, M. & Hakonarson, H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 38, e164 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  96. Agarwal, V., Bell, G. W., Nam, J. W. & Bartel, D. P. Predicting effective microRNA target sites in mammalian mRNAs. eLife 4, e05005 (2015).

  97. Schildge, S., Bohrer, C., Beck, K. & Schachtrup, C. Isolation and culture of mouse cortical astrocytes. J. Vis. Exp. https://doi.org/10.3791/50079 (2013).

  98. Hsu, P. D. et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 31, 827–832 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Doench, J. G. et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR–Cas9. Nat. Biotechnol. 34, 184–191 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Xu, H. et al. Sequence determinants of improved CRISPR sgRNA design. Genome Res. 25, 1147–1157 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Chari, R., Mali, P., Moosburner, M. & Church, G. M. Unraveling CRISPR–Cas9 genome engineering parameters via a library-on-library approach. Nat. Methods 12, 823–826 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Erijman, A., Dantes, A., Bernheim, R., Shifman, J. M. & Peleg, Y. Transfer-PCR (TPCR): a highway for DNA cloning and protein engineering. J. Struct. Biol. 175, 171–177 (2011).

    Article  CAS  PubMed  Google Scholar 

  103. Peleg, Y. & Unger, T. Application of the restriction-free (RF) cloning for multicomponents assembly. Methods Mol. Biol. 1116, 73–87 (2014).

    Article  CAS  PubMed  Google Scholar 

  104. Keren-Shaul, H. et al. MARS-seq2.0: an experimental and analytical pipeline for indexed sorting combined with single-cell RNA sequencing. Nat. Protoc. 14, 1841–1862 (2019).

    Article  CAS  PubMed  Google Scholar 

  105. Jaitin, D. A. et al. Massively parallel single-cell RNA-seq for marker-free decomposition of tissues into cell types. Science 343, 776–779 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Kohen, R. et al. UTAP: user-friendly transcriptome analysis pipeline. BMC Bioinformatics 20, 154 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  107. Cox, J. & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 (2008).

    Article  CAS  PubMed  Google Scholar 

  108. Cox, J. & Mann, M. Quantitative, high-resolution proteomics for data-driven systems biology. Annu. Rev. Biochem. 80, 273–299 (2011).

    Article  CAS  PubMed  Google Scholar 

  109. Cox, J., Michalski, A. & Mann, M. Software lock mass by two-dimensional minimization of peptide mass errors. J. Am. Soc. Mass Spectrom. 22, 1373–1380 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Cox, J. et al. Accurate proteome-wide label-free quantification by delayed normalization and maximal peptide ratio extraction, termed MaxLFQ. Mol. Cell. Proteomics 13, 2513–2526 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Tyanova, S. et al. The Perseus computational platform for comprehensive analysis of (prote)omics data. Nat. Methods 13, 731–740 (2016).

    Article  CAS  PubMed  Google Scholar 

  112. Krach, F. et al. Transcriptome–pathology correlation identifies interplay between TDP-43 and the expression of its kinase CK1E in sporadic ALS. Acta Neuropathol. 136, 405–423 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Thompson, L. iMN (Exp 2)—ALS, SMA and control (unaffected) iMN cell lines differentiated from iPS cell lines using a long differentiation protocol—RNA-seq. LINCS (collection) http://identifiers.org/lincs.data/LDG-1338 (2017).

Download references

Acknowledgements

We gratefully acknowledge the contributions of all participants and the investigators who provided biological samples and data for the Project Mine ALS sequencing consortium, the NYGC ALS Consortium, the gnomAD and TOPMed of the NHLBI (https://www.nhlbiwgs.org/topmed-banner-authorship). We thank M. Ward (NINDS, NIH) for sharing human inducible i3LMN cells. Samples used in this research were in part obtained from the UK National DNA Bank for MND Research, funded by the MND Association and the Wellcome Trust. We acknowledge sample management undertaken by Biobanking Solutions funded by the Medical Research Council at the Centre for Integrated Genomic Medical Research, University of Manchester. We would like to thank the NINDS Biorepository at Coriell Institute for iPSC cell lines used in this study. We thank B. Oldak and J. Hanna for microglia differentiation protocols, N. Kozer and H. Barr for assistance with live-cell imaging, A. Savidor and Y. Levin for mass spectrometry and M. Shmueli, Y. Merbl and R. Rotkof for advice and protocols. We thank LSE for language and scientific editing. Some illustrations were created with BioRender. The Hornstein lab is supported by friends of S. Brenner. E.H. is Head of Andi and Larry Wolfe Center for Research on Neuroimmunology and Neuromodulation and incumbent of Ira & Gail Mondry Professorial chair. This work is funded by Legacy Heritage Fund (828/17), Bruno and Ilse Frick Foundation for Research on ALS, the RADALA Foundation for ALS research, Teva Pharmaceutical Industries., Ltd., as part of the Israeli National Network of Excellence in Neuroscience (NNE) and Minna-James-Heineman Stiftung through Minerva, the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ERC grant agreement number 617351, Israel Science Foundation (135/16, 3497/21); Target ALS 118945, the Minerva Foundation, with funding from the Federal German Ministry for Education and Research, the ALS-Therapy Alliance, AFM Telethon (20576 to E.H.), Motor Neuron Disease Association (UK), The Thierry Latran Foundation for ALS research, ERA-Net for Research Programmes on Rare Diseases (FP7), via the Israel Ministry of Health. A. Alfred Taubman through IsrALS, Yeda-Sela, Yeda-CEO, Israel Ministry of Trade and Industry, Y. Leon Benoziyo Institute for Molecular Medicine, Kekst Family Institute for Medical Genetics, David and Fela Shapell Family Center for Genetic Disorders Research, Crown Human Genome Center, Nathan, Shirley, Philip and Charlene Vener New Scientist Fund, Julius and Ray Charlestein Foundation, Fraida Foundation, Wolfson Family Charitable Trust, Adelis Foundation, Merck (UK), Maria Halphen, Estates of Fannie Sherr, Lola Asseof, Lilly Fulop, Andi and Larry Wolfe Center for Research on Neuroimmunology and Neuromodulation and Benoziyo center for Neurological diseases, Weizmann—Brazil Center for Research on Neurodegeneration at The Weizmann Institute of Science, Redhill Foundation—Sam and Jean Rothberg Charitable Trust, Edward and Janie Moravitz, the Israeli Council for Higher Education via the Weizmann Data Science Research Center and a research grant from the Estate of Tully and Michele Plesser and M. Judith Ruth Institute for Preclinical Brain Research. A.A.-C. received funding from Neurodegenerative Disease Research (JPND), Medical Research Council (MR/L501529/1, STRENGTH, MR/R024804/1, BRAIN-MEND), Economic and Social Research Council (ES/L008238/1, ALS-CarE), MND Association, National Institute for Health Research (NIHR) Biomedical Research Centre at South London and Maudsley NHS Foundation Trust and King’s College London. This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement number 772376, EScORIAL). The collaboration project is cofunded by the PPP Allowance made available by Health~Holland, Top Sector Life Sciences & Health to stimulate public–private partnerships. This study was supported by the ALS Foundation Netherlands. For P.V.D., Project MinE Belgium was supported by a grant from IWT (number 140935), the ALS Liga België, the National Lottery of Belgium and the KU Leuven Opening the Future Fund. P.V.D. holds a senior clinical investigatorship of FWO-Vlaanderen and is supported by E. von Behring Chair for Neuromuscular and Neurodegenerative Disorders, the ALS Liga België and the KU Leuven funds ‘Een Hart voor ALS’, ‘Laeversfonds voor ALS Onderzoek’ and the ‘Valéry Perrier Race against ALS Fund’. Several authors of this publication are members of the European Reference Network for Rare Neuromuscular Diseases. P.J.S. received funding from the Medical Research Council, MND Association, NIHR Senior Investigator Award, NIHR Sheffield Biomedical Research Centre and NIHR Sheffield Clinical Research Facility. P.M.A. received funding from the Knut and Alice Wallenberg Foundation, the Swedish Brain Foundation, the Swedish Science Council and the Ulla-Carin Lindquist Foundation. H.P.P. and sequencing activities at NYGC were supported by the ALS Association and The Tow Foundation. C.E. was supported by a scholarship from Teva Pharmaceutical Industries, Ltd., as part of the NNE. S.M.K.F. is supported by the ALS Canada Tim E. Noël Postdoctoral Fellowship. R.H.B.J. was funded by the ALS Association, ALS Finding a Cure, Angel Fund, ALS-One, Cellucci Fund and NIH grants (R01 NS104022, R01 NS073873 and NS111990-01 to R.H.B.J.). J.K.I. is a New York Stem Cell Foundation-Robertson Investigator. N.S.Y. was supported by the Israeli Council for Higher Education via the Weizmann Data Science Research Center, by a research grant from the Estate of Tully and Michele Plesser and by Maccabim Foundation. Work in the J.K.I. lab was supported by NIH grant R01NS097850, U.S. Department of Defense grant W81XWH-19-PRARP-CSRA and grants from the Tau Consortium, the New York Stem Cell Foundation, the ALS Association and the John Douglas French Alzheimer’s Foundation. R.L.McL. received funding from the Science Foundation Ireland (17/CDA/4737), and A.N.B. received funding from the Suna and Inan Kirac Foundation. J.E.L. received funding from the National Institute of Health/NINDS (R01 NS073873).

Author information

Authors and Affiliations

Authors

Consortia

Contributions

C.E. and A. Siany led the project. C.E. and A. Siany contributed to research conception, design and interpretations and wrote the manuscript with E.H. C.E., E.B., T.O., K.R.V.E., M.M., S.M.K.F., N.S.Y., J.C.-K., K.P.K., R.A.A.V.D.S., W.S., A.A.K., A.I., A. Shatunov, A.R.J., E.C., D.R., O.W., R.H.B.J., P.J.S., P.V.D., L.H.v.d.B., H.P., E.S., A.A.-C. and J.H.V. collected samples, were involved in the sequence analysis pipeline, phenotyping, variant calling, provided expertise or were involved in the genetic association analysis of rare noncoding variants in individuals with ALS. S.-T.H. and J.K.I. provided stem cells and initial data. S.B.-D., E.A., G.B. and H.M.-K. were involved in the design, generation and validation of CRISPR-edited IL18RAP isogenic iPSCs. H.M.-K. and Y.M.D. performed the pulldown experiments of IL18RAP 3′UTR RNA-associated proteins and analyzed the proteomic data. A. Siany and C.E. established human iPSC-derived microglia differentiation and culturing protocols, performed motor neuron survival experiments and interpreted data. A. Siany, N.R. and C.E. performed molecular biology studies in LCLs and U2OS cell lines, including reporter assays, qPCR and protein quantification by western blotting. E.Y. performed the bulk MARS-seq experiment. C.-H.Y., C.L. and S.L.M. provided expertise and processed mouse cortex samples for flow cytometry. Y.C., Y.E.-A., S.W. and D.P.S. helped perform research. E.H. conceived and supervised the study and wrote the manuscript with C.E. and A. Siany. All coauthors provided approval of the manuscript.

Corresponding author

Correspondence to Eran Hornstein.

Ethics declarations

Competing interests

J.K.I. is a cofounder of AcuraStem Incorporated. J.K.I. declares that he is bound by confidentiality agreements that prevent him from disclosing details of his financial interests in this work. J.H.V. and L.H.v.d.B. report to have sponsored research agreements with Biogen. E.H. is an inventor on pending patent family PCT/IL2016/050328 entitled ‘Methods of treating motor neuron diseases’. All other authors declare that they have no competing interests.

Peer review

Peer review information

Nature Neuroscience thanks Aaron Gitler, Donna Werling, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 Study design.

(a) Flow chart of approach for discovery of region-based rare-variants in non-coding genomic regions via association studies and (b) diagram depicting regions of interest comprising of 1,750 autosomal human pre-miRNA genes, 295 open reading frames encoding for proteins of interest, and 295 3′UTRs.

Extended Data Fig. 2 Region-based rare-variant association analyses.

(a-d) QQ (quantile-quantile) probability plot, of obtained and expected P-values (log scale) gained by region-based rare-variant association analysis of different genomic regions, comprised of (a) 295 candidate protein-coding regions listed in Supplementary Table 3. These ORFs encode for ALS-relevant proteins or proteins that are associated with miRNA biogenesis or activity. Variants were depicted if predicted to cause frameshifting, alternative splicing, abnormal stop codon or a deleterious non-synonymous amino acid substitution, in ≥ 3 of 7 independent dbNSFP prediction algorithms (genomic inflation λ = 0.96), (b) All known pre-miRNA genes in the human genome (genomic inflation λ = 1.31), (c) predicted networks, comprised of aggregated variants detected on a specific mature miRNA sequence and its cognate downstream 3’UTR targets (genomic inflation λ = 1.16), and (d) variants in 3′UTRs of the same 295 genes listed in Supplementary Table 3 (genomic inflation λ = 1.08). Data was obtained from 3,955 ALS cases and 1,819 controls (Project MinE). Features positioned on the diagonal line represent results obtained under the null hypothesis. Open-reading frames of 10 known ALS genes (blue). IL18RAP 3′UTR (red). P-values, calculated with SKAT-O.

Extended Data Fig. 3 3′UTR-based rare-variant association analysis, using different algorithms, and illustration of rare variants identified in the IL18RAP 3′UTR.

(a-d) QQ plot of obtained and expected P-values (log scale) gained by region-based rare-variant association analysis of genomic regions comprised of 295 3′UTRs listed in Supplementary Table 3, in the Project MinE cohort (3,955 ALS cases and 1,819 non-ALS controls). Features positioned on the diagonal line represent results obtained under the null hypothesis. IL18RAP 3′UTR (red) is the most significant 3’UTR associated with ALS using different algorithms: (a) Sequence Kernel Association Test, SKAT (genomic inflation λ = 1.02), (b) Combined Multivariate and Collapsing, CMC (genomic inflation λ = 1.34), (c) Variable Threshold with permutation analysis, VT (genomic inflation λ = 1.03). (d) IL18RAP 3′UTR also ranked as the top hit when aggregating variants abrogating or gaining miRNA recognition elements (MREs) in 3’UTRs (genomic inflation λ = 1.04). (e) Schematic of the IL18RAP transcript and 3′UTR (5′ to 3ʹ) showing the number of control (upper) or ALS (lower) samples in which variants (black arrow) were identified in the Project MinE discovery cohort (Supplementary Table 6).

Extended Data Fig. 4 Restricting rare-variant association analysis to the proximal part of 3’UTRs does not improve the association signal.

(a) Scatter plot with SKAT-O P-values (log scale) calculated for region-based rare-variant association analysis of the full 3’UTRs on the x-axis versus the 3’UTRs proximal quadrant on the y-axis, for the 295 3′UTRs listed in Supplementary Table 3, in the Project MinE cohort (3,955 ALS cases and 1,819 non-ALS controls) (Pearson correlation coefficient (r=0.61) and P-value ****<0.0001). The 45-degree diagonal line represents a perfect correlation of r=1. IL18RAP 3′UTR (red). (b) A Difference plot showing the difference between the two P-value measurements (3’UTRs proximal quadrant minus the full 3’UTRs, for the cohort of N=295 3’UTRs). The bias (difference between means) is only 0.03. Overall the P-values gained from the 3’UTRs proximal quadrant were comparable to that of the full 3’UTRs in the cohort of 295 3’UTRs. For box plot, the median is indicated by the central line, upper and lower quartiles are indicated by the box, and maximum/minimum values are indicated by the whiskers (Wilcoxon matched-pairs P-value > 0.05, Cohen’s d effect size = 0.1). Hence, the apparent spatial distribution of variants in IL18RAP 3’UTR seems to be a particular case, rather than part of a global pattern.

Extended Data Fig. 5 IL18RAP and p-NF-κB expression is elevated in lymphoblastoid cells from patients with the C9orf72 repeat expansion.

(a) IL18RAP mRNA expression (qPCR normalized to IPO8 mRNA levels) and (b) IL18RAP or (c) p-NF-κB protein expression (Western blots, normalized to Tubulin). Extracts from eight different human lymphoblastoid cell lines (listed in Supplementary Table 8): Four lines of healthy individuals (without ALS) carrying the canonical IL18RAP 3’UTR sequence (Control; Canonical IL18RAP 3’UTR, black) and four C9orf72 ALS patients carrying the canonical IL18RAP 3’UTR sequence (C9orf72; Canonical IL18RAP 3’UTR, red). (d) Representative blots processed with anti-IL18RAP, anti p-NF-κB and anti-Tubulin antibodies. Mann-Whitney test (A) or one-sided student’s t-test with Welch’s correction on log-transformed data (P = 0.056 for panel B; P = 0.0065 for panel C), was conducted based on the mean value of three independent passages for each of the eight human lymphoblastoid cell lines (Source Data Extended Data Fig. 5). Scatter dot plot with mean and SEM. **P<0.01.

Source data

Extended Data Fig. 6 IL18RAP 3’UTR variants attenuate IL-18 - NF-κB signaling in U2OS cells.

Diagram (a) and quantification (b) of NF-κB reporter assays in human U2OS cell line. To determine the ability of the IL18RAP variants V3 and V1 to induce NF-κB activity, U2OS cells were co-transfected with different IL18RAP coding region (CDS) and 3’UTR constructs (GFP, Canonical, V3, V1, n=9; 3CDS, n=4), along with an NF-κB activity reporter that drives luciferase (Luc2P) transcription via five copies of the NF-κB response element. NF-κB signaling was induced by adding human recombinant IL-18 to the medium. Variants V3 and V1 of the IL18RAP 3’UTR reduced NF-κB activity by ~10% and ~21%, respectively, relative to the WT IL18RAP 3’UTR. GFP vector and a dominant-negative coding mutant E210A-Y212A-Y214A CDS + WT 3’UTR (3CDS)31, served as controls. Luciferase expression was normalized to transfected U2OS cells that were not induced with human recombinant IL-18. One-way ANOVA followed by Dunnett’s multiple comparison test was performed on square root-transformed data. For box plots, the median is indicated by the central line, upper and lower quartiles are indicated by the box, and maximum/minimum values are indicated by the whiskers. * P<0.05; *** P<0.001. The experiment was repeated independently three times with similar results.

Extended Data Fig. 7 IL18RAP is mainly expressed on mouse microglia cells.

(a-c) Flow cytometry was used to characterize IL18RAP expression levels in dissociated wild-type mouse cortex cells. The expression of IL-18RAP (IL-18Rβ) was expressed as Mean Fluorescence Intensity (MFI) and % frequency after gating for the following cell types: immune cells (CD45hi), microglia (MG: CD45int CD11hi), neurons (CD45- CD11b- NeuN+), and astrocytes (CD45- CD11b- GFAP+). FACS analysis reveals that IL18RAP is mainly expressed on microglia cells. A scatter dot plot with mean and SEM values for the median fluorescence intensity (MFI) and percentage of IL18RAP+ cells is shown. One-way ANOVA followed by Tukey’s multiple comparison test. ** P<0.01, *** P<0.001, **** P<0.0001.

Extended Data Fig. 8 Evaluation of IL18RAP and IL-18 mRNA expression in motor neurons of patients with ALS.

(a-b) mRNA expression of IL18RAP (a) and IL-18 (b), as reads per kilobase million (RPKM), from NGS study of laser capture microdissection–enriched surviving motor neurons from lumbar spinal cords of patients with sALS with rostral onset and caudal progression (n = 12) and non-neurodegeneration controls (n = 9112; GSE76220). Two-sided Student’s t test with Welch’s correction on log-transformed data (P = 0.0138 for panel A; P = 0.0056 for panel B). (c) IL-18 mRNA expression, as log2-normalized counts, from NGS study of induced ALS motor neurons (n = 4 different donors in duplicates) or non-neurodegeneration controls (n=3 different donors in duplicates113; DESeq analysis, P = 0.0417). For box plots, the median is indicated by the central line, upper and lower quartiles are indicated by the box, and maximum/minimum values are indicated by the whiskers. *P < 0.05; **P < 0.01.

Extended Data Fig. 9 iPSC-derived microglia express the microglial-specific marker, TMEM119.

Immunofluorescence staining of TMEM119 (green) and DAPI (blue), in two different C9orf72 iPSC-derived progenitor microglia lines. Lens, ×20; scale bar, 100 μm.

Extended Data Fig. 10 Differentially bound RNA binding proteins to variant 3’UTR (V3) relative to canonical 3’UTR.

(a) Volcano plot of protein abundance associated with the canonical relative to variant (V3) IL18RAP 3’UTR (x-axis log2 scale), analyzed by MS. Y-axis depicts P-values (−log10 scale). Proteins significantly enriched in association with canonical/variant 3’UTR are colored (gray/orange). Features above the horizontal dashed line demarcate proteins with adjusted p < 0.05, in student’s t-test with FDR correction to multiple hypotheses. Vertical dashed lines are of 2 or ½ fold change (Supplementary Table 9). (b) Prediction of 3’UTR secondary structure by RNA Fold90, suggests a minor change to the structure of the sequence harboring a V3 variant (red), relative to the canonical 3’UTR (green).

Supplementary information

Reporting Summary

Supplementary Table

Supplementary Tables 1–18 and consortium member lists.

Supplementary Video 1

Motor neuron survival was significantly improved in the presence of microglia harboring variant IL18RAP 3′UTR relative to canonical IL18RAP 3′UTR.

Source data

Source Data Fig. 3

Source data for IL18RAP and p-NF-κB western blot studies in LCLs (Fig. 3d).

Source Data Fig. 4

Source data for IL18RAP western blot studies in isogenic microglia (Fig. 4b).

Source Data Fig. 6

Source data for motor neuron survival assays (Fig. 6b,c).

Source Data Fig. 7

Source data for p-NF-κB western blot studies in isogenic microglia following microglia activation (Fig. 7b).

Source Data Extended Data Fig. 5

Source data for IL18RAP and p-NF-κB western blot studies in control versus C9orf72 LCLs (Extended Data Fig. 5d).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Eitan, C., Siany, A., Barkan, E. et al. Whole-genome sequencing reveals that variants in the Interleukin 18 Receptor Accessory Protein 3′UTR protect against ALS. Nat Neurosci 25, 433–445 (2022). https://doi.org/10.1038/s41593-022-01040-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-022-01040-6

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