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:

Complex genetic signatures in immune cells underlie autoimmunity and inform therapy

An Author Correction to this article was published on 18 September 2020

This article has been updated

Abstract

We report on the influence of ~22 million variants on 731 immune cell traits in a cohort of 3,757 Sardinians. We detected 122 significant (P < 1.28 × 10−11) independent association signals for 459 cell traits at 70 loci (53 of them novel) identifying several molecules and mechanisms involved in cell regulation. Furthermore, 53 signals at 36 loci overlapped with previously reported disease-associated signals, predominantly for autoimmune disorders, highlighting intermediate phenotypes in pathogenesis. Collectively, our findings illustrate complex genetic regulation of immune cells with highly selective effects on autoimmune disease risk at the cell-subtype level. These results identify drug-targetable pathways informing the design of more specific treatments for autoimmune 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: Schematic representation of the main leukocyte subsets assessed by GWAS.
Fig. 2: Genetic associations of the immune traits assessed.
Fig. 3: Coincident genetic association between immune traits and diseases.
Fig. 4: Regional association plots in the CD40 region.
Fig. 5: Drug target prioritization (priority index, Pi) score of our drug target candidates.

Similar content being viewed by others

Data availability

Full GWAS summary statistics have been deposited in the GWAS Catalog with accession numbers from GCST0001391 (https://www.ebi.ac.uk/gwas/studies/GCST0001391) to GCST0002121 (https://www.ebi.ac.uk/gwas/studies/GCST0002121). The accession number for each trait is reported in Supplementary Table 1B.

Disease summary statistics used to identify coincident associations were obtained from the GWAS Catalog (https://www.ebi.ac.uk/gwas/home, version v1.0, Ensembl release version e93, date 2018-08-28) and from ImmunoBase (https://genetics.opentargets.org/immunobase, version 12-May-2016). Summary statistics for colocalization analyses were downloaded from the respective web pages: RA (http://plaza.umin.ac.jp/~yokada/datasource/files/GWASMetaResults/RA_GWASmeta_European_v2.txt.gz); T1D (https://datadryad.org/stash/dataset/doi:10.5061/dryad.ns8q3); MS (http://imsgc.net/publications/); SLE (http://insidegen.com/insidegen-LUPUS-data.html); allergy (https://genepi.qimr.edu.au/staff/manuelf/gwas_results/SHARE-without23andMe.LDSCORE-GC.SE-META.v0.gz); IBD, Crohn’s disease and ulcerative colitis (ftp://ftp.sanger.ac.uk/pub/project/humgen/summary_statistics/human/2016-11-07/). Molecular QTLs are from the LinDA QTL Catalog (http://linda.irgb.cnr.it/, version 20190109). Source data are provided with this paper.

Change history

  • 18 September 2020

    An amendment to this paper has been published and can be accessed via a link at the top of the paper.

References

  1. Orrù, V. et al. Genetic variants regulating immune cell levels in health and disease. Cell 155, 242–256 (2013).

    PubMed  PubMed Central  Google Scholar 

  2. Roederer, M. et al. The genetic architecture of the human immune system: a bioresource for autoimmunity and disease pathogenesis. Cell 161, 387–403 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Aguirre-Gamboa, R. et al. Differential effects of environmental and genetic factors on T and B cell immune traits. Cell Rep. 17, 2474–2487 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Patin, E. et al. Natural variation in the parameters of innate immune cells is preferentially driven by genetic factors. Nat. Immunol. 19, 302–314 (2018).

    CAS  PubMed  Google Scholar 

  5. Lagou, V. et al. Genetic architecture of adaptive immune system identifies key immune regulators. Cell Rep. 25, 798–810 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Sidore, C. et al. Genome sequencing elucidates Sardinian genetic architecture and augments association analyses for lipid and blood inflammatory markers. Nat. Genet. 47, 1272–1281 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Chang, Y. J., Zhao, X. Y. & Huang, X. J. Immune reconstitution after haploidentical hematopoietic stem cell transplantation. Biol. Blood Marrow Transplant. 20, 440–449 (2014).

    CAS  PubMed  Google Scholar 

  8. Ogonek, J. et al. Immune reconstitution after allogeneic hematopoietic stem cell transplantation. Front. Immunol. 17, 507 (2016).

    Google Scholar 

  9. Brodin, P. et al. Variation in the human immune system is largely driven by non-heritable influences. Cell 160, 37–47 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Benner, C. et al. FINEMAP: efficient variable selection using summary data from genome-wide association studies. Bioinformatics 32, 1493–1501 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Steri, M. et al. Overexpression of the cytokine BAFF and autoimmunity risk. N. Engl. J. Med. 376, 1615–1626 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Han, J. W. et al. Genome-wide association study in a Chinese Han population identifies nine new susceptibility loci for systemic lupus erythematosus. Nat. Genet. 41, 1234–1237 (2009).

    CAS  PubMed  Google Scholar 

  13. Javierre, B. M. et al. Lineage-specific genome architecture links enhancers and non-coding disease variants to target gene promoters. Cell 167, 1369–1384 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Westra, H. J. et al. Systematic identification of trans eQTLs as putative drivers of known disease associations. Nat. Genet. 45, 1238–1243 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Langefeld, C. D. et al. Transancestral mapping and genetic load in systemic lupus erythematosus. Nat. Commun. 8, 16021 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. International Multiple Sclerosis Genetics Consortium (IMSGC) et al. Analysis of immune-related loci identifies 48 new susceptibility variants for multiple sclerosis. Nat. Genet. 45, 1353–1360 (2013).

  17. Jostins, L. et al. Host–microbe interactions have shaped the genetic architecture of inflammatory bowel disease. Nature 491, 119–124 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Jiang, D. K. et al. Genetic variants in five novel loci including CFB and CD40 predispose to chronic hepatitis B. Hepatology 62, 118–128 (2015).

    CAS  PubMed  Google Scholar 

  19. Okada, Y. et al. Genetics of rheumatoid arthritis contributes to biology and drug discovery. Nature 506, 376–381 (2014).

    CAS  PubMed  Google Scholar 

  20. Lee, Y. C. et al. Two new susceptibility loci for Kawasaki disease identified through genome-wide association analysis. Nat. Genet. 44, 522–525 (2012).

    CAS  PubMed  Google Scholar 

  21. Ferreira, M. A. et al. Shared genetic origin of asthma, hay fever and eczema elucidates allergic disease biology. Nat. Genet. 49, 1752–1757 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Ellinghaus, D. et al. Analysis of five chronic inflammatory diseases identifies 27 new associations and highlights disease-specific patterns at shared loci. Nat. Genet. 48, 510–518 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Onengut-Gumuscu, S. et al. Fine mapping of type 1 diabetes susceptibility loci and evidence for colocalization of causal variants with lymphoid gene enhancers. Nat. Genet. 47, 381–386 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Hinks, A. et al. Dense genotyping of immune-related disease regions identifies 14 new susceptibility loci for juvenile idiopathic arthritis. Nat. Genet. 45, 664–669 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Liu, J. Z. et al. Dense genotyping of immune-related disease regions identifies nine new risk loci for primary sclerosing cholangitis. Nat. Genet. 45, 670–675 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Petukhova, L. et al. Genome-wide association study in alopecia areata implicates both innate and adaptive immunity. Nature 466, 113–117 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Betz, R. C. et al. Genome-wide meta-analysis in alopecia areata resolves HLA associations and reveals two new susceptibility loci. Nat. Commun. 6, 5966 (2015).

    CAS  PubMed  Google Scholar 

  28. International Multiple Sclerosis Genetics Consortium (IMSGC) et al. Genetic risk and a primary role for cell-mediated immune mechanisms in multiple sclerosis. Nature 476, 214–219 (2011).

  29. Jin, Y. et al. Variant of TYR and autoimmunity susceptibility loci in generalized vitiligo. N. Engl. J. Med. 362, 1686–1697 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Cooper, J. D. et al. Seven newly identified loci for autoimmune thyroid disease. Hum. Mol. Genet. 21, 5202–5208 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Trynka, G. et al. Dense genotyping identifies and localizes multiple common and rare variant association signals in celiac disease. Nat. Genet. 43, 1193–1201 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Eyre, S. et al. High-density genetic mapping identifies new susceptibility loci for rheumatoid arthritis. Nat. Genet. 44, 1336–1340 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Liu, J. Z. et al. Association analyses identify 38 susceptibility loci for inflammatory bowel disease and highlight shared genetic risk across populations. Nat. Genet. 47, 979–986 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Qiu, F. et al. A genome-wide association study identifies six novel risk loci for primary biliary cholangitis. Nat. Commun. 20, 14828 (2017).

    Google Scholar 

  35. Fortune, M. D. et al. Statistical colocalization of genetic risk variants for related autoimmune diseases in the context of common controls. Nat. Genet. 47, 839–846 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Jin, Y. et al. Genome-wide association studies of autoimmune vitiligo identify 23 new risk loci and highlight key pathways and regulatory variants. Nat. Genet. 48, 1418–1424 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Bradfield, J. P. et al. A genome-wide meta-analysis of six type 1 diabetes cohorts identifies multiple associated loci. PLoS Genet. 7, e1002293 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Morris, D. E. et al. Genome-wide association meta-analysis in Chinese and European individuals identifies ten new loci associated with systemic lupus erythematosus. Nat. Genet. 48, 940–946 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Roychoudhuri, R. et al. BACH2 represses effector programs to stabilize T(reg)-mediated immune homeostasis. Nature 498, 506–510 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Plenge, R. M., Scolnick, E. M. & Altshuler, D. Validating therapeutic targets through human genetics. Nat. Rev. Drug Discov. 12, 581–594 (2013).

    CAS  PubMed  Google Scholar 

  41. Floris, M., Olla, S., Schlessinger, D. & Cucca, F. Genetic-driven druggable target identification and validation. Trends Genet. 34, 558–570 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Fang, H. et al. A genetics-led approach defines the drug target landscape of 30 immune-related traits. Nat. Genet. 51, 1082–1091 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Cytlak, U. et al. Ikaros family zinc finger 1 regulates dendritic cell development and function in humans. Nat. Commun. 9, 1239 (2018).

    PubMed  PubMed Central  Google Scholar 

  44. Presto, J. K. & Werth, V. P. Cutaneous lupus erythematosus: current treatment options. Curr. Treat. Options Rheumatol. 2, 36–48 (2016).

    Google Scholar 

  45. Kim, J. M., Park, S. H., Kim, H. Y. & Kwok, S. K. A plasmacytoid dendritic cells-type I interferon axis is critically implicated in the pathogenesis of systemic lupus erythematosus. Int. J. Mol. Sci. 16, 14158–14170 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Dzionek, A. BDCA-2, a novel plasmacytoid dendritic cell-specific type II C-type lectin, mediates antigen capture and is a potent inhibitor of interferon alpha/beta induction. J. Exp. Med. 194, 1823–1834 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Blomberg, S., Eloranta, M. L., Magnusson, M., Alm, G. V. & Rönnblom, L. Expression of the markers BDCA-2 and BDCA-4 and production of interferon-alpha by plasmacytoid dendritic cells in systemic lupus erythematosus. Arthritis Rheumatol. 48, 2524–2532 (2003).

    CAS  Google Scholar 

  48. Cao, W. et al. BDCA2/Fc epsilon RI gamma complex signals through a novel BCR-like pathway in human plasmacytoid dendritic cells. PLoS Biol. 5, e248 (2007).

    PubMed  PubMed Central  Google Scholar 

  49. Gottlieb, P. A. Failure to preserve β-cell function with mycophenolate mofetil and daclizumab combined therapy in patients with new- onset type 1 diabetes. Diabetes Care 33, 826–832 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Du, F. H. et al. Next-generation anti-CD20 monoclonal antibodies in autoimmune disease treatment. Autoimmun. Highlights 8, 12 (2017).

    Google Scholar 

  51. Labrijn, A. F., Janmaat, M. L., Reichert, J. M. & Parren, P. W. H. I. Bispecific antibodies: a mechanistic review of the pipeline. Nat. Rev. Drug Discov. 18, 585–608 (2019).

    CAS  PubMed  Google Scholar 

  52. Gateva, V. et al. A large-scale replication study identifies TNIP1, PRDM1, JAZF1, UHRF1BP1 and IL10 as risk loci for systemic lupus erythematosus. Nat. Genet. 41, 1228–1233 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. McGovern, D. et al. Genome-wide association identifies multiple ulcerative colitis susceptibility loci. Nat. Genet. 42, 332–337 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Pala, M. et al. Population- and individual-specific regulatory variation in Sardinia. Nat. Genet. 49, 700–707 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. de Lange, K. M. et al. Genome-wide association study implicates immune activation of multiple integrin genes in inflammatory bowel disease. Nat. Genet. 49, 256–261 (2017).

    PubMed  PubMed Central  Google Scholar 

  56. Pruim, R. J. et al. LocusZoom: regional visualization of genome-wide association scan results. Bioinformatics 26, 2336–2337 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Pilia, G. et al. Heritability of cardiovascular and personality traits in 6,148 Sardinians. PLoS Genet. 2, e132 (2006).

    PubMed  PubMed Central  Google Scholar 

  58. Das, S. et al. Next-generation genotype imputation service and methods. Nat. Genet. 48, 1284–1287 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Pistis, G. et al. Rare variant genotype imputation with thousands of study-specific whole-genome sequences: implications for cost-effective study designs. Eur. J. Hum. Genet. 23, 975–983 (2015).

    PubMed  Google Scholar 

  60. 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).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Ferreira, M. A. et al. Quantitative trait loci for CD4:CD8 lymphocyte ratio are associated with risk of type 1 diabetes and HIV-1 immune control. Am. J. Hum. Genet. 86, 88–92 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Bentham, J. et al. Genetic association analyses implicate aberrant regulation of innate and adaptive immunity genes in the pathogenesis of systemic lupus erythematosus. Nat. Genet. 47, 1457–1464 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Censing, J. C. et al. Childhood adiposity and risk of type 1 diabetes: A Mendelian randomization study. PLoS Med. 14, e1002362 (2017).

    Google Scholar 

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

    PubMed  PubMed Central  Google Scholar 

  65. Palmer, T. M. et al. Instrumental variable estimation of causal risk ratios and causal odds ratios in Mendelian randomization analyses. Am. J. Epidemiol. 173, 1392–1403 (2011).

    PubMed  Google Scholar 

  66. Hemani, G. et al. The MR-Base platform supports systematic causal inference across the human phenome. eLife 7, e34408 (2018).

    PubMed  PubMed Central  Google Scholar 

  67. Teumer, A. Common methods for performing Mendelian randomization. Front. Cardiovasc. Med. 5, 51 (2018).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank all of the volunteers who generously participated in this study; S. Naitza and J. Todd for helpful suggestions; the Consortium ‘Sa Corona Arrubia della Marmilla’ for making available equipment and scientific instruments; and the IMSGC and WTCCC2 consortia for access to the summary statistics from the key disease GWAS. We acknowledge support by grants (2011/R/13 and 2015/R/09, to F.C.) from the Italian Foundation for Multiple Sclerosis; contracts (HHSN271201600005C, to F.C.) from the Intramural Research Program of the National Institute on Aging, National Institutes of Health (NIH); a grant (FaReBio2011 ‘Farmaci e Reti Biotecnologiche di Qualità’, to F.C.) from the Italian Ministry of Economy and Finance; a grant (633964, to F.C.) from the Horizon 2020 Research and Innovation Program of the European Union; grants (U1301.2015/AI.1157. BE Prat. 2015-1651, to F.C. and U965.2014/AI.847.MGB Prat. 2014.0597, to M.S.) from Fondazione di Sardegna; grants (‘Centro per la Ricerca di Nuovi Farmaci per Malattie Rare, Trascurate e della Povertà’ and ‘Progetto Collezione di Composti Chimici ed Attività di Screening’ to F.C.) from Ministero dell’Istruzione, dell’Università e della Ricerca.

Author information

Authors and Affiliations

Authors

Contributions

F.C. conceived and supervised the study. F.C., V.O., M.S. and C.S. drafted the manuscript; M. Pala, S. Olla and M.F. contributed to the writing of specific sections; D.S., M.G., M. Devoto, M.Z. and G.R.A. revised the manuscript; V.O., E.F. and F.C. designed flow cytometric panels; M. Dei., S.L., V.S. and F.V. measured immunophenotypes; V.O., V.S. and E.F. analyzed raw immunophenotype data; A.M., S.L., M. Dei, M.L., M.G.P., M. Pitzalis, F.D. and A.L. isolated DNA; M.Z. and A.M. performed array genotyping; M.S., C.S., Michele Marongiu and G.S. carried out genetic analysis; M. Pala, S. Onano and Mara Marongiu performed RNA data analyses; M.F. and S. Olla carried out bioinformatics and drug target analyses; Michele Marongiu managed the SardiNIA project database; S.S. shared genetic data on MS; F.C., G.R.A. and D.S. provided funds and reagents. All authors read the paper and contributed to its final form.

Corresponding author

Correspondence to Francesco Cucca.

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.

Extended data

Extended Data Fig. 1 Flow cytometry gating strategy of TBNK, regulatory T cell, maturation stages of T cell, and dendritic cell antibody panel.

TBNK panel. a, Lymphocytes (violet) and granulocytes (blue). b, CD14+ monocytes (light blue). c, HLA DR++CD14+ monocytes. d, CD3+ (T cells, purple) and CD3- (green) lymphocytes. e, B and NK cells are CD19+ and CD16/CD56+, respectively. f, HLA DR+ NK cells. g, T cells are divided based on CD4 and CD8 expression. h, TCR-ϒδ+ T cells. i, NKT are CD3+ and CD16/CD56+. j, HLA DR+ T cells. HLA DR+CD4 and HLA DR+CD8 subsets are obtained by intersecting HLA DR+ T cell with CD4+ and CD8br lymphocytes. Regulatory T cell panel. a, CD4+ (blue) and CD8+ (violet) lymphocytes. b, CD4+ Tregs (green) are CD25high CD127low. c, Resting (CD45RA+CD25++, pink), activated (CD45RA-CD25+++, orange) and secreting (CD45RA-CD25++, purple) CD4+ Tregs. D-E) CD25hiCD4+ lymphocytes are divided based on CD45RA expression. F) CD4-CD8- T cells (DN, black) are divided in CD28+ and CD28-. G-H-I-J-K) CD39 expression on Treg subsets, CD4 and CD8br T cells, respectively. L) CD8br cells division based on CD45RA vs CD28 expression. M-N) CD25++CD28-CD8br and CD127-CD28-CD8br identification. Maturation stages of T cell panel. a, CD4+ (blue), CD8br (violet) and CD4-CD8- (black) T cells are analyzed for CD45RA vs CCR7 (plots B-C-D, respectively) identifying naïve (CCR7+CD45RA+), central memory (CM, CCR7+CD45RA-), effector memory (EM, CCR7-CD45RA-), and terminally differentiated (TD, CCR7-CD45RA+) subsets. Dendritic cell antibody panel. a, Monocytes (pink). b, c, DCs are Lineage (Lin) negative and HLA DR+. d, Myeloid (green) and plasmacytoid (violet) DCs are CD11c+ and CD123+, respectively. e, f, CD86 and CD62L expression on cDC. G-H-I) CD11c, CD62L and HLA DR expression on monocytes.

Extended Data Fig. 2 Flow cytometry gating strategy of B cell, monocyte, myeloid cell antibody panel.

B cell panel. a-b, Lymphocyte (red). c, B lymphocytes (violet) are CD19+. d, IgD+ B cells. B cells classification based on e) CD24 vs CD38; f) CD27 vs IgD; g) IgD vs CD38; h) IgD vs CD24; i) CD24+CD27+ memory B cells. j, Plasma blasts/plasma cells (PB/PC) are CD20-CD38hi B cells. Monocyte panel. A-B-C) Monocyte (blue). D) Monocytes division into CD14+CD16- (classical), CD14-CD16+ (non-classical) and CD14+CD16+ (intermediate). Myeloid panel. A-B-C) Lympho-monocytes (red). d, Viable and myeloid-enriched cells (green) are obtained excluding lymphoid cells, which are lineage1 (Lin1) positive, and dead cells, which are 7-aminoactinomycin-D (7AAD) positive. e, Hematopoietic stem cells (HSC). f, CD14 vs CD66b expression and g) CD33 vs HLA DR expression on myeloid-enriched cells. The intersection of CD33dim/br HLA DRdim/- cells in g) with CD14+ monocytes (orange) in f) results in monocytic myeloid-derived dendritic cells (Mo MDSC). h, The deletion of CD14+ monocytes (orange) from cells in g) discriminates five subsets using CD33 vs HLA DR markers. i, CD66b+ cells were excluded from the CD33dim HLA DR- cells (blue) and j) the resulting CD33dimHLA DR-CD66- population was further divided into basophils and immature MDSC (Im MDSC) based on CD45 and CD11b expression. k, CD33br HLA DR+ cells (black) division into CD14 dim and CD14-. l, CD11b expression on CD33dim HLA DR+ cells (purple). Intersection of CD33dim HLA DR- in h) with CD66b++ cells in f) corresponds to granulocytic myeloid-derived dendritic cells (Gr MDSC).

Extended Data Fig. 3 Phenotypic correlation among expression level of surface markers.

Heatmap of phenotypic correlations for MFI pairs calculated using the Spearman coefficient. Dendrograms represent the clustering: short branches indicate strong phenotypic correlation between traits, whereas long branches weak correlation. Color gradations represent the correlation strength, with red indicating direct correlation (from 0 to +1) and blue inverse correlation (from 0 to -1).

Source data

Extended Data Fig. 4 Genetic correlation among expression level of surface markers.

Heatmap of genetic correlations for MFIs pairs calculated as previously described1. The description of the figure is as for Extended Data Fig. 3.

Source data

Extended Data Fig. 5 Phenotypic correlation among cell levels.

Heatmap of phenotypic correlations for cell counts and T/B and CD4/CD8br ratios, calculated using the Spearman coefficient. The description of the figure is as for Extended Data Fig. 3.

Source data

Extended Data Fig. 6 Genetic correlation among absolute cell counts.

Heatmap of genetic correlations for cell counts and T/B and CD4/CD8br ratios, calculated as previously described1. The description of the figure is as for Extended Data Fig. 3.

Source data

Extended Data Fig. 7 Drug target prioritization (Priority index, Pi) score of our drug targets candidates segmented by gene categories.

It is shown the distributions of the Pi-rating (computed in Fang et al., 42) of our candidate genes (colored boxplots) segmented for different gene categories (“All” genes, “eGenes”, “Seed” genes and cell surface genes) with the relative background distributions (grey boxplots, that is that consider all the genes belonging to the respective category). eGenes means that the gene has an eQTL colocalizing with the disease; seed gene means that the genes have a genetic link to the disease (by eQTL, gene proximity or chromatin conformation) as defined in Fang et al., 42. The boxplot inside the violin plot reports a white circle indicating the median value, with the box limits indicating the upper and the lower quartiles. The whisker at the upper side of the box extends to the minimum between the interquartile range (IQR) x 1.5 and the overall maximum value of the data. The whisker at the bottom side of the box extends to the maximum between IQR x 1.5 and the overall minimum value of the data.

Source data

Supplementary information

Supplementary Information

Supplementary table list, results and Fig. 1.

Reporting Summary

Supplementary Tables

Supplementary Tables 1–8.

Source data

Source Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Orrù, V., Steri, M., Sidore, C. et al. Complex genetic signatures in immune cells underlie autoimmunity and inform therapy. Nat Genet 52, 1036–1045 (2020). https://doi.org/10.1038/s41588-020-0684-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-020-0684-4

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