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:

RNA editing underlies genetic risk of common inflammatory diseases

Abstract

A major challenge in human genetics is to identify the molecular mechanisms of trait-associated and disease-associated variants. To achieve this, quantitative trait locus (QTL) mapping of genetic variants with intermediate molecular phenotypes such as gene expression and splicing have been widely adopted1,2. However, despite successes, the molecular basis for a considerable fraction of trait-associated and disease-associated variants remains unclear3,4. Here we show that ADAR-mediated adenosine-to-inosine RNA editing, a post-transcriptional event vital for suppressing cellular double-stranded RNA (dsRNA)-mediated innate immune interferon responses5,6,7,8,9,10,11, is an important potential mechanism underlying genetic variants associated with common inflammatory diseases. We identified and characterized 30,319 cis-RNA editing QTLs (edQTLs) across 49 human tissues. These edQTLs were significantly enriched in genome-wide association study signals for autoimmune and immune-mediated diseases. Colocalization analysis of edQTLs with disease risk loci further pinpointed key, putatively immunogenic dsRNAs formed by expected inverted repeat Alu elements as well as unexpected, highly over-represented cis-natural antisense transcripts. Furthermore, inflammatory disease risk variants, in aggregate, were associated with reduced editing of nearby dsRNAs and induced interferon responses in inflammatory diseases. This unique directional effect agrees with the established mechanism that lack of RNA editing by ADAR1 leads to the specific activation of the dsRNA sensor MDA5 and subsequent interferon responses and inflammation7,8,9. Our findings implicate cellular dsRNA editing and sensing as a previously underappreciated mechanism of common inflammatory diseases.

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: RNA edQTL map in human tissues.
Fig. 2: Contribution of edQTLs to complex diseases and traits.
Fig. 3: Identification, characterization and validation of disease-relevant, putatively immunogenic dsRNAs.
Fig. 4: Risk variants of inflammatory diseases collectively associated with reduced nearby dsRNA editing levels and induction of interferon response.

Similar content being viewed by others

Data availability

GTEx V8 release editing-level data and edQTL call sets are available on the GTEx Portal: https://gtexportal.org. Details of the GWAS summary statistics used for colocalization are provided in Supplementary Table 1. icSHAPE data were obtained from the RASP database: http://rasp.zhanglab.net/. CLIP data were obtained from the POSTAR3 database: http://postar.ncrnalab.org/.

Code availability

A full pipeline to preprocess the GWAS and edQTL data, prioritize relevant loci, run colocalization tests and generate the associated plots is publicly available at https://github.com/mikegloudemans/rna-editing-coloc and https://github.com/vargasliqin/GTEx_edQTL.

References

  1. Albert, F. W. & Kruglyak, L. The role of regulatory variation in complex traits and disease. Nat. Rev. Genet. 16, 197–212 (2015).

    Article  CAS  PubMed  Google Scholar 

  2. GTEx Consortium. The GTEx Consortium atlas of genetic regulatory effects across human tissues. Science 369, 1318–1330 (2020).

    Article  Google Scholar 

  3. Chun, S. et al. Limited statistical evidence for shared genetic effects of eQTLs and autoimmune-disease-associated loci in three major immune-cell types. Nat. Genet. 49, 600–605 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Umans, B. D., Battle, A. & Gilad, Y. Where are the disease-associated eQTLs? Trends Genet. 37, 109–124 (2021).

    Article  CAS  PubMed  Google Scholar 

  5. Nishikura, K. Functions and regulation of RNA editing by ADAR deaminases. Annu. Rev. Biochem. 79, 321–349 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Rice, G. I. et al. Mutations in ADAR1 cause Aicardi–Goutieres syndrome associated with a type I interferon signature. Nat. Genet. 44, 1243–1248 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Mannion, N. M. et al. The RNA-editing enzyme ADAR1 controls innate immune responses to RNA. Cell Rep. 9, 1482–1494 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Liddicoat, B. J. et al. RNA editing by ADAR1 prevents MDA5 sensing of endogenous dsRNA as nonself. Science 349, 1115–1120 (2015).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Pestal, K. et al. Isoforms of RNA-editing enzyme ADAR1 independently control nucleic acid sensor MDA5-driven autoimmunity and multi-organ development. Immunity 43, 933–944 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Eisenberg, E. & Levanon, E. Y. A-to-I RNA editing—immune protector and transcriptome diversifier. Nat. Rev. Genet. 19, 473–490 (2018).

    Article  CAS  PubMed  Google Scholar 

  11. Samuel, C. E. Adenosine deaminase acting on RNA (ADAR1), a suppressor of double-stranded RNA-triggered innate immune responses. J. Biol. Chem. 294, 1710–1720 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Li, Y. I. et al. RNA splicing is a primary link between genetic variation and disease. Science 352, 600–604 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  13. Jonkhout, N. et al. The RNA modification landscape in human disease. RNA 23, 1754–1769 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Walkley, C. R. & Li, J. B. Rewriting the transcriptome: adenosine-to-inosine RNA editing by ADARs. Genome Biol. 18, 205 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Ramaswami, G. & Li, J. B. Identification of human RNA editing sites: a historical perspective. Methods 107, 42–47 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Ramaswami, G. et al. Accurate identification of human Alu and non-Alu RNA editing sites. Nat. Methods 9, 579–581 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Tan, M. H. et al. Dynamic landscape and regulation of RNA editing in mammals. Nature 550, 249–254 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  18. Porath, H. T., Knisbacher, B. A., Eisenberg, E. & Levanon, E. Y. Massive A-to-I RNA editing is common across the metazoa and correlates with dsRNA abundance. Genome Biol. 18, 185 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Chalk, A. M., Taylor, S., Heraud-Farlow, J. E. & Walkley, C. R. The majority of A-to-I RNA editing is not required for mammalian homeostasis. Genome Biol. 20, 268 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Rice, G. I. et al. Gain-of-function mutations in IFIH1 cause a spectrum of human disease phenotypes associated with upregulated type I interferon signaling. Nat. Genet. 46, 503–509 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Nejentsev, S., Walker, N., Riches, D., Egholm, M. & Todd, J. A. Rare variants of IFIH1, a gene implicated in antiviral responses, protect against type 1 diabetes. Science 324, 387–389 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Emdin, C. A. et al. Analysis of predicted loss-of-function variants in UK Biobank identifies variants protective for disease. Nat. Commun. 9, 1613 (2018).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  23. Li, Y. et al. Carriers of rare missense variants in IFIH1 are protected from psoriasis. J. Invest. Dermatol. 130, 2768–2772 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Huang, H. et al. Fine-mapping inflammatory bowel disease loci to single-variant resolution. Nature 547, 173–178 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. Jin, Y., Andersen, G. H. L., Santorico, S. A. & Spritz, R. A. Multiple functional variants of IFIH1, a gene involved in triggering innate immune responses, protect against vitiligo. J. Invest. Dermatol. 137, 522–524 (2017).

    Article  CAS  PubMed  Google Scholar 

  26. Sun, B. B. et al. Genetic associations of protein-coding variants in human disease. Nature 603, 95–102 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Shallev, L. et al. Decreased A-to-I RNA editing as a source of keratinocytes' dsRNA in psoriasis. RNA 24, 828–840 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Vlachogiannis, N. I. et al. Increased adenosine-to-inosine RNA editing in rheumatoid arthritis. J. Autoimmun. 106, 102329 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Roth, S. H. et al. Increased RNA editing may provide a source for autoantigens in systemic lupus erythematosus. Cell Rep. 23, 50–57 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Tossberg, J. T., Heinrich, R. M., Farley, V. M., Crooke, P. S. 3rd & Aune, T. M. Adenosine-to-inosine RNA editing of Alu double-stranded (ds)RNAs is markedly decreased in multiple sclerosis and unedited Alu dsRNAs are potent activators of proinflammatory transcriptional responses. J. Immunol. 205, 2606–2617 (2020).

    Article  CAS  PubMed  Google Scholar 

  31. Belkadi, A. et al. Identification of genetic variants controlling RNA editing and their effect on RNA structure stabilization. Eur. J. Hum. Genet. 28, 1753–1762 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Ruan, H. et al. GPEdit: the genetic and pharmacogenomic landscape of A-to-I RNA editing in cancers. Nucleic Acids Res. 50, D1231–D1237 (2022).

    Article  CAS  PubMed  Google Scholar 

  33. Breen, M. S. et al. Global landscape and genetic regulation of RNA editing in cortical samples from individuals with schizophrenia. Nat. Neurosci. 22, 1402–1412 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Park, E., Jiang, Y., Hao, L., Hui, J. & Xing, Y. Genetic variation and microRNA targeting of A-to-I RNA editing fine tune human tissue transcriptomes. Genome Biol. 22, 77 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Franzen, O. et al. Global analysis of A-to-I RNA editing reveals association with common disease variants. PeerJ 6, e4466 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Park, E. et al. Population and allelic variation of A-to-I RNA editing in human transcriptomes. Genome Biol. 18, 143 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Wijetunga, N. A. et al. The meta-epigenomic structure of purified human stem cell populations is defined at cis-regulatory sequences. Nat. Commun. 5, 5195 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  38. Baeza-Centurion, P., Minana, B., Schmiedel, J. M., Valcarcel, J. & Lehner, B. Combinatorial genetics reveals a scaling law for the effects of mutations on splicing. Cell 176, 549–563.e23 (2019).

    Article  CAS  PubMed  Google Scholar 

  39. Ongen, H., Buil, A., Brown, A. A., Dermitzakis, E. T. & Delaneau, O. Fast and efficient QTL mapper for thousands of molecular phenotypes. Bioinformatics 32, 1479–1485 (2016).

    Article  CAS  PubMed  Google Scholar 

  40. Urbut, S. M., Wang, G., Carbonetto, P. & Stephens, M. Flexible statistical methods for estimating and testing effects in genomic studies with multiple conditions. Nat. Genet. 51, 187–195 (2019).

    Article  CAS  PubMed  Google Scholar 

  41. Matthews, M. M. et al. Structures of human ADAR2 bound to dsRNA reveal base-flipping mechanism and basis for site selectivity. Nat. Struct. Mol. Biol. 23, 426–433 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Hormozdiari, F. et al. Leveraging molecular quantitative trait loci to understand the genetic architecture of diseases and complex traits. Nat. Genet. 50, 1041–1047 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Finucane, H. K. et al. Partitioning heritability by functional annotation using genome-wide association summary statistics. Nat. Genet. 47, 1228–1235 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Navab, M., Anantharamaiah, G. M., Reddy, S. T., Van Lenten, B. J. & Fogelman, A. M. HDL as a biomarker, potential therapeutic target, and therapy. Diabetes 58, 2711–2717 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Zeisbrich, M. et al. Hypermetabolic macrophages in rheumatoid arthritis and coronary artery disease due to glycogen synthase kinase 3b inactivation. Ann. Rheum. Dis. 77, 1053–1062 (2018).

    Article  CAS  PubMed  Google Scholar 

  46. Misra, M. K., Damotte, V. & Hollenbach, J. A. The immunogenetics of neurological disease. Immunology 153, 399–414 (2018).

    Article  CAS  PubMed  Google Scholar 

  47. Suciu, C. F. et al. Oxidized low density lipoproteins: the bridge between atherosclerosis and autoimmunity. Possible implications in accelerated atherosclerosis and for immune intervention in autoimmune rheumatic disorders. Autoimmun. Rev. 17, 366–375 (2018).

    Article  CAS  PubMed  Google Scholar 

  48. Yao, D. W., O'Connor, L. J., Price, A. L. & Gusev, A. Quantifying genetic effects on disease mediated by assayed gene expression levels. Nat. Genet. 52, 626–633 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Sayaman, R. W. et al. Germline genetic contribution to the immune landscape of cancer. Immunity 54, 367–386.e8 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Bazak, L. et al. A-to-I RNA editing occurs at over a hundred million genomic sites, located in a majority of human genes. Genome Res. 24, 365–376 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Faghihi, M. A. & Wahlestedt, C. Regulatory roles of natural antisense transcripts. Nat. Rev. Mol. Cell Biol. 10, 637–643 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Dias Junior, A. G., Sampaio, N. G. & Rehwinkel, J. A balancing act: MDA5 in antiviral immunity and autoinflammation. Trends Microbiol. 27, 75–85 (2019).

    Article  CAS  PubMed  Google Scholar 

  53. Wu, B. et al. Structural basis for dsRNA recognition, filament formation, and antiviral signal activation by MDA5. Cell 152, 276–289 (2013).

    Article  CAS  PubMed  Google Scholar 

  54. Li, P., Zhou, X., Xu, K. & Zhang, Q. C. RASP: an atlas of transcriptome-wide RNA secondary structure probing data. Nucleic Acids Res. 49, D183–D191 (2021).

    Article  CAS  PubMed  Google Scholar 

  55. Zhao, W. et al. POSTAR3: an updated platform for exploring post-transcriptional regulation coordinated by RNA-binding proteins. Nucleic Acids Res. 50, D287–D294 (2022).

    Article  CAS  PubMed  Google Scholar 

  56. Reshef, Y. A. et al. Detecting genome-wide directional effects of transcription factor binding on polygenic disease risk. Nat. Genet. 50, 1483–1493 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Patterson, J. B. & Samuel, C. E. Expression and regulation by interferon of a double-stranded-RNA-specific adenosine deaminase from human cells: evidence for two forms of the deaminase. Mol. Cell. Biol. 15, 5376–5388 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Heraud-Farlow, J. E. & Walkley, C. R. The role of RNA editing by ADAR1 in prevention of innate immune sensing of self-RNA. J. Mol. Med. 94, 1095–1102 (2016).

    Article  CAS  PubMed  Google Scholar 

  59. Guo, Y. et al. CD40L-dependent pathway is active at various stages of rheumatoid arthritis disease progression. J. Immunol. 198, 4490–4501 (2017).

    Article  CAS  PubMed  Google Scholar 

  60. Elkjaer, M. L. et al. Molecular signature of different lesion types in the brain white matter of patients with progressive multiple sclerosis. Acta Neuropathol. Commun. 7, 205 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Tokuyama, M. et al. ERVmap analysis reveals genome-wide transcription of human endogenous retroviruses. Proc. Natl Acad. Sci. USA 115, 12565–12572 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Numata, K. et al. Comparative analysis of cis-encoded antisense RNAs in eukaryotes. Gene 392, 134–141 (2007).

    Article  CAS  PubMed  Google Scholar 

  63. Ramaswami, G. & Li, J. B. RADAR: a rigorously annotated database of A-to-I RNA editing. Nucleic Acids Res. 42, D109–D113 (2014).

    Article  CAS  PubMed  Google Scholar 

  64. Zhang, R. et al. Quantifying RNA allelic ratios by microfluidic multiplex PCR and sequencing. Nat. Methods 11, 51–54 (2014).

    Article  CAS  PubMed  Google Scholar 

  65. Ramaswami, G. et al. Genetic mapping uncovers cis-regulatory landscape of RNA editing. Nat. Commun. 6, 8194 (2015).

    Article  ADS  PubMed  Google Scholar 

  66. Tran, S. S. et al. Widespread RNA editing dysregulation in brains from autistic individuals. Nat. Neurosci. 22, 25–36 (2019).

    Article  CAS  PubMed  Google Scholar 

  67. Storey, J. D. & Tibshirani, R. Statistical significance for genomewide studies. Proc. Natl Acad. Sci. USA 100, 9440–9445 (2003).

    Article  ADS  MathSciNet  CAS  PubMed  PubMed Central  MATH  Google Scholar 

  68. Frankish, A. et al. GENCODE reference annotation for the human and mouse genomes. Nucleic Acids Res. 47, D766–D773 (2019).

    Article  CAS  PubMed  Google Scholar 

  69. Bahn, J. H. et al. Genomic analysis of ADAR1 binding and its involvement in multiple RNA processing pathways. Nat. Commun. 6, 6355 (2015).

    Article  ADS  CAS  PubMed  Google Scholar 

  70. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  71. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).

    Article  CAS  PubMed  Google Scholar 

  72. Konkel, M. K. et al. Sequence analysis and characterization of active human Alu subfamilies based on the 1000 Genomes Pilot Project. Genome Biol. Evol. 7, 2608–2622 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Eggington, J. M., Greene, T. & Bass, B. L. Predicting sites of ADAR editing in double-stranded RNA. Nat. Commun. 2, 319 (2011).

    Article  ADS  PubMed  Google Scholar 

  74. Wagih, O. ggseqlogo: A versatile R package for drawing sequence logos. Bioinformatics 33, 3645–3647 (2017).

    Article  CAS  PubMed  Google Scholar 

  75. Tian, S. & Das, R. RNA structure through multidimensional chemical mapping. Q. Rev. Biophys. 49, e7 (2016).

    Article  PubMed  Google Scholar 

  76. Danaee, P. et al. bpRNA: large-scale automated annotation and analysis of RNA secondary structure. Nucleic Acids Res. 46, 5381–5394 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Purcell, S. et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Pers, T. H., Timshel, P. & Hirschhorn, J. N. SNPsnap: a web-based tool for identification and annotation of matched SNPs. Bioinformatics 31, 418–420 (2015).

    Article  CAS  PubMed  Google Scholar 

  79. Gazal, S. et al. Linkage disequilibrium-dependent architecture of human complex traits shows action of negative selection. Nat. Genet. 49, 1421–1427 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Rice, G. I. et al. Assessment of type I interferon signaling in pediatric inflammatory disease. J. Clin. Immunol. 37, 123–132 (2017).

    Article  CAS  PubMed  Google Scholar 

  81. Giambartolomei, C. et al. Bayesian test for colocalisation between pairs of genetic association studies using summary statistics. PLoS Genet. 10, e1004383 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Balbin, O. A. et al. The landscape of antisense gene expression in human cancers. Genome Res. 25, 1068–1079 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  84. Zhang, F., Lu, Y., Yan, S., Xing, Q. & Tian, W. SPRINT: an SNP-free toolkit for identifying RNA editing sites. Bioinformatics 33, 3538–3548 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We acknowledge the GTEx consortium for making the data available for this study; members of the Li and Montgomery laboratories for insightful discussions and suggestions; C. Walkley, J. Engreitz, A. Marderstein and O. de Geode for critical reading of the manuscript; and X. Liu, W. Tsui and V. Sochat for technical support. This work was funded by US National Institutes of Health grants R01 GM102484, R01 GM124215, R01 MH115080, R35 GM144100, R01 AG066490, R01 MH125244, U01 HG009431 and U01 HG007593. Q.L. was partially funded by the American Heart Association Postdoctoral Fellowship (17IFUNP33820059). M.J.G. was funded by NLM training grant T15 LM 007033 and a Stanford Graduate Fellowship.

Author information

Authors and Affiliations

Authors

Contributions

Q.L., M.J.G., S.B.M. and J.B.L. conceptualized the study. Q.L., M.J.G., J.M.G., B.F., G.R., Y.I.L., J.-B.M., S.B.M. and J.B.L. performed the study. Q.L., M.J.G., J.M.G., B.F. and F.A. undertook the formal analysis. Q.L. and J.B.L. wrote the original draft of the manuscript. Q.L., M.J.G., J.M.G., T.S., G.R., Y.I.L., J.K.P., S.B.M. and J.B.L. reviewed and edited the manuscript. J.B.L. acquired funding. S.B.M. and J.B.L. supervised the study.

Corresponding author

Correspondence to Jin Billy Li.

Ethics declarations

Competing interests

S.B.M. is a consultant for BioMarin, MyOme and Tenaya Therapeutics. J.B.L. is a co-founder of AIRNA Bio and a consultant for Risen Pharma. J.B.L. and Q.L. are named inventors of a provisional patent filed by Stanford University (serial no. 63/473,678), describing a method related to this work. The other authors declare no competing interests.

Peer review

Peer review information

Nature thanks Kaur Alasoo 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 figures and tables

Extended Data Fig. 1 Quantification of RNA editing levels in GTEx data.

a, Number of editing sites used for cis-edQTL mapping in each tissue. Editing sites mapped uniquely by ≥20 reads in ≥60 samples of each tissue type were considered. Both GTEx V6p and V8 results are plotted here for comparison. b, Comparison of RNA editing levels quantified using exact read count versus estimated read count for cis-NAT editing sites (Methods). To maximize the number of tested cis-NAT editing sites, we collectively analyzed strand-specific RNA-seq data of 5 human tissues (spleen, lung, thyroid, colon and skin) by pooling the raw reads together. In total, n = 806 editing sites of 47 cis-NAT dsRNAs were used for comparison (≥20 reads coverage). c, Percentage of editing variance across individuals explained by top 10 principal components (PCs) of editing level. d, Association between ADAR1 expression level with PC 1 of editing level in n = 175 Brain – cerebellar hemisphere samples. e, Density plot of editing level measurements and editing level variance between individuals (n = 175; normalized standard deviation) of 123,707 sites in Brain – cerebellar hemisphere tissue. Editing sites detected in ≥60 individuals were used for this analysis.

Extended Data Fig.2 Identification of RNA editing QTLs across GTEx tissues.

a, Fraction of editing sites (left) and edited genes (right) found with edQTLs across n = 49 GTEx tissues. The LCLs data was obtained from previous edQTL study33. b, Fractions of edQTLs shared between multiple editing sites. c, Fractions of edGenes having multiple independent edQTLs. Each of the tissues (n = 49) was tested and plotted individually in b and c. d, Exemplary locus of TRMT9B showing two independent edQTLs each regulating multiple editing sites. The two edQTLs are represented by their lead SNPs rs13268982 and rs34995506, respectively. Genome browser view of the TRMT9B 3’ UTR that consists of four IRAlus (shown in the 1st track from top), with editing sites (2nd track from top), SNP location (black bars) and variant-editing association effect sizes for each editing site (3rd and 4th tracks) shown from top to bottom. Effect sizes are colored by direction of effects: red for positive effect and blue for negative effect. e, Comparison of the number of edSites identified in this study and in Park et al., 2021. For each tissue type, edSites are compared between two studies by evaluating (1) whether they are tested for genetic association, and (2) whether the associations are significant under different p-value cut-offs (p < 1e-3 on the left, p < 1e-5 on the right). f, Sharing of edQTLs across tissues by sign (same direction of effects). Euclidean distance matrix across tissues was calculated for hierarchical clustering using Ward's method. g, Fraction of edQTLs shared by the number of tissues according to sign. g’, Fraction of edQTLs shared by the number of tissues according to magnitude. Effects with more than 2-fold change of sizes from one another are considered as of different magnitudes. h, Number of tissue-specific edQTLs effects found in 20 representative tissues. Representative tissues were picked by two criteria: 1) large sample size (≥150); 2) least number of shared editing sites with other tissue. For example, a high number of editing sites are shared between three arteries tissues so we choose Artery-Coronary tissue as the representative one since it has the largest sample size.

Extended Data Fig. 3 Comparison between edQTLs, eQTLs and sQTLs in GTEx tissues.

a, Sharing of edQTLs with eQTLs and sQTLs. We estimated the fraction of shared QTLs according to Storey’s π1 (Methods). SNPs with matching numbers and allele frequencies with the edQTLs were randomly sampled genome-wide in each tissue to be used as the control set. b, Distance between the best edQTL and best eQTL for genes with both types of QTL, using 1,478 genes in whole blood tissue as an example. c, Enrichment of functional elements underlying edQTLs, eQTLs and sQTLs (left to right). We used a Bayesian hierarchical model to identify putatively causal variants driving a QTL from a set of variants associated with the locus, and quantified the enrichment of strongly associated variants in functional elements. For edQTLs, we additionally identified the functional elements underlying edQTL SNPs that are >800 bp away from the edSites (grey dots). We used chromHMM annotations and gene-level annotations (e.g. intronic variant, splice region variants, UTR variants, etc.) from snpEff. Splice region variants include variants located within the region of splice site (1-3 bases of the exon side or 3-8 bases of the intron side) and branch point. Structured RNA are defined as regions with icSHAPE score ≥0.7 in RNA structure mapping data in vivo from multiple cell lines obtained from http://rasp.zhanglab.net/. ADAR1 and other RBP binding sites are defined using CLIP peaks obtain from http://postar.ncrnalab.org/. Error bars represent 95% confidence intervals. edQTLs n = 30,319; eQTLs n = 24,740; sQTLs n = 14,424.

Extended Data Fig. 4 Characterization of edQTLs’ effects on RNA sequences and secondary structures.

a, Q-Q plot of edQTLs (n = 30,319) annotated with distance from the associated editing sites. b, Summarized edQTL effect sizes of alternative alleles by 4 types of RNA ribonucleotides at each position from –50 to +50 nt relative to the editing site. Stronger effects were observed for ±2 and ±1 positions and illustrated to show the “AUAGG” motif centered at the edited “A”. c, Further breakdown of the effects of 12 different nucleotide changes (reference allele -> alternative allele) caused by SNPs located at ±2 and ±1 positions. Reference alleles (y-axis) and alternative alleles (x-axis) were accounted for by their transcribed ribonucleotides on the RNA. d, An exemplary edSite association showing negative effects of C-to-U change at +2 position of editing site. Editing site at chr1: 184761188 and SNP at chr1: 184761186 are presented here (hg38 coordinates). Predicted RNA secondary structures for reference allele (G on DNA and C on RNA) and alternative allele (A on DNA and U on RNA) are shown on the left, with editing level measurements of different alternative allele dosages on the right. Biologically independent sample size n = 138. e, Schematic diagram showing RNA secondary structural features defined by bpRNA66. f, Frequency of edQTL SNPs found in different structural features (n = 142) compared to non-QTL SNPs (n = 4,080) in the same predicted structure. Statistical significance was calculated using two-sided Mann–Whitney U test. Box plots show interquartile ranges and median, with whiskers extending to minima and maxima.

Extended Data Fig. 5 RNA editing QTLs are enriched in immune-related diseases and immune traits.

a, QQ-plots of IBD, MS, RA and CAD GWAS with QTL annotations. The expression levels of eGenes and sGenes were matched to the levels of edGenes within ±15% of deviation (see Methods). b, Enrichment of heritability for 42 human traits and diseases mediated by edQTLs, sQTLs and eQTLs. Autoimmune diseases are shown in bold. Enrichment is measured as the regression coefficient of MESC (Methods). Error bars represent standard deviation with mean as center. Meta-analysis of RNA editing QTL mediated heritability is shown in c and heritability enrichment in c’, for 4 exemplary autoimmune and immune-related diseases. Tissue groups are defined the same way as in Fig. 2c. Tissues and tissues groups shown in c’ are the same as in c. Error bars represent standard deviation with mean as center. Sample sizes of studies are in Supplementary Table 1.

Extended Data Fig.6 Fraction of SNP heritability (h2g) mediated by edQTL in immune-related trait GWAS.

We performed MESC analysis on 33 GWAS data (n = 9,138 individuals) obtained from Sayaman et al.42. in the same way as in Fig. 2C. We filtered out 11 immune traits with low total SNP heritability (h2g < 0.01) from the final result. The remaining 22 traits are categorized and colored by immune function annotations according to Sayaman et al.42. Dashed line indicates 0.1 of mediated SNP heritability (h2g)42.

Extended Data Fig. 7 Formation of cis-NAT dsRNAs in vitro with MDA5 proteins and in human cells.

For (a) TNFRSF14:TNFRSF14-AS1, (b) CTSA:PLTP and (c) HBP1:COG cis-NAT dsRNAs, negative stain electron microscopy (EM) images are shown for MDA5 proteins incubated with cis-NAT dsRNAs (left), sense strand ssRNA control (middle) and antisense strand ssRNA control (right), respectively. All filamentation experiments were repeated independently for 3 times for each cis-NAT pair with similar results. d, In vitro editing status of dsRNAs formed by TNFRSF14:TNFRSF14-AS1 cis-NAT. Editing sites in both sense (TNFRSF14, top) and antisense (TNFRSF14-AS1, bottom) strands are shown as red bars separately for the two strands. Editing information was determined by Sanger sequencing. e, Genome browser snapshot of the CTSA:PLTP cis-NAT locus. Annotations of gene structure, Alu repeats, exonic editing sites, RNA structure mapping signal (icSHAPE scores in + and – strands) and ADAR1 CLIP signal are shown from top to bottom. f, Correlation of icSHAPE scores between sense and antisense strands for cis-NAT dsRNAs (n = 38). icSHAPE data was obtained from the RASP database: http://rasp.zhanglab.net/. ADAR1 CLIP data of human U87 cell lines was obtained from the POSTAR3 database: http://postar.ncrnalab.org/.

Extended DataFig. 8 Overexpression of CTSA:PLTP cis-NAT in human cells.

a, Schematic diagram showing plasmid construction for bidirectional transcription of CTSA:PLTP cis-NAT. Expression is driven by two opposing EF-1alpha promoters. Sense strand (CTSA 3’UTR) and antisense strand (PLTP 3’UTR) of the cis-NAT are downstream of mClover3 and mRuby3, respectively, with 207 bp overlapping sequences to mimic cis-NAT dsRNA formation. b, Cartoon showing vector design expressing cis-NAT, its scrambled formation and ssRNA control. c, RT-PCR data for validating the transcription of sense (CTSA) and antisense (PLTP) strands of CTSA:PLTP cis-NAT under different treatments. Amplicons (spanning mClover3-CTSA) are expected only when CTSA is expressed, and amplicons (spanning mRuby3-PLTP) are expected only when PLTP is expressed. Experiments were repeated independently for 2 times with similar results. d, Real-time PCR measurement of overexpression of CTSA:PLTP cis-NAT relative to mClover3 ssRNA control. Overexpression of sense (CTSA) and antisense (PLTP) strands was measured by the co-expressed mClover3 and mRuby3, respectively. n = 1 independent experiment. e, Editing level of CTSA:PLTP cis-NAT in HEK293 cells with WT ADAR1. Editing level was measured by Sanger sequencing.

Extended Data Fig. 9 Directional effects of risk variants associated with complex traits and diseases on RNA editing levels.

a, Traits and diseases with overall negative effects on RNA editing (increased risk associated with reduced editing level). b, Traits and diseases with overall positive or unclear direction of effects on RNA editing. c, Directional effect analysis of four exemplary diseases for edQTLs (top row), eQTLs (middle row) and control edQTLs with randomly flipped signs (bottom row). In all panels, we plotted the genome-wide −log10(P) against estimated effect size in GTEx tissues (n = 49), all estimated using SLDP regression. Significant tissues are colored in red (Bonferroni corrected multiple test p-value < 1x10−3). Larger circles denote lower p-values.

Extended Data Fig. 10 Risk genetic variants are associated with reduced RNA editing levels and induced IFN score in inflammatory diseases.

RNA-seq data from patient-derived samples of three diseases (multiple sclerosis, lupus and CAD) were used to calculate the overall editing levels of dsRNAs associated with risk vs protective alleles and the IFN scores. We used allele-specific editing analysis to determine editing levels associated with risk vs protective alleles (a, b and c). The IFN scores were calculated by summing up the expression of representative ISGs for each disease (Methods), and grouped in four different bins. Samples are then grouped by their editing level differences (protective alleles – risk alleles, divided by quartiles) to show IFN scores plotted for each group and compared between groups of the same disease type (a’, b’ and c’ with matching sample size as a, b and c, respectively). ***, P < 0.001; **, P < 0.01; one-way ANOVA test. Box plots in a’, b’ and c’ show interquartile ranges and median, with whiskers extending to minima and maxima.

Supplementary information

Reporting Summary

Peer Review File

Supplementary Table 1

Information of GWAS summary statistics used in this study.

Supplementary Table 2

Estimated heritability mediated by RNA editing in complex traits and diseases.

Supplementary Table 3

Overlap between edQTLs and GWAS catalog and colocalization results of edQTLs in GWAS of 24 traits and diseases.

Supplementary Table 4

Information of putatively immunogenic dsRNA loci formed by cis-NATs.

Supplementary Table 5

Allele-specific editing (ASED) levels measured in immune-related disease samples.

Supplementary Table 6

List of primer sequences used in this work.

Rights and permissions

Springer Nature or its licensor 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

Li, Q., Gloudemans, M.J., Geisinger, J.M. et al. RNA editing underlies genetic risk of common inflammatory diseases. Nature 608, 569–577 (2022). https://doi.org/10.1038/s41586-022-05052-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05052-x

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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