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:

Functional characterization of Alzheimer’s disease genetic variants in microglia

Abstract

Candidate cis-regulatory elements (cCREs) in microglia demonstrate the most substantial enrichment for Alzheimer’s disease (AD) heritability compared to other brain cell types. However, whether and how these genome-wide association studies (GWAS) variants contribute to AD remain elusive. Here we prioritize 308 previously unreported AD risk variants at 181 cCREs by integrating genetic information with microglia-specific 3D epigenome annotation. We further establish the link between functional variants and target genes by single-cell CRISPRi screening in microglia. In addition, we show that AD variants exhibit allelic imbalance on target gene expression. In particular, rs7922621 is the effective variant in controlling TSPAN14 expression among other nominated variants in the same cCRE and exerts multiple physiological effects including reduced cell surface ADAM10 and altered soluble TREM2 (sTREM2) shedding. Our work represents a systematic approach to prioritize and characterize AD-associated variants and provides a roadmap for advancing genetic association to experimentally validated cell-type-specific phenotypes and mechanisms.

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: Characterization of hPSC-derived microglia-like cells.
Fig. 2: Chromatin accessibility and interaction dynamics influence IFNβ-stimulated transcriptional changes in microglia.
Fig. 3: Fine-mapping of AD risk loci with microglia 3D epigenome annotation.
Fig. 4: Characterization of fine-mapped AD risk loci with pooled CRISPRi and targeted scRNA-seq in hPSC-derived microglia-like cells.
Fig. 5: Linking prioritized AD variants to phenotypes by allelic analyses and cellular functional assays.

Similar content being viewed by others

Data availability

All datasets used in this study (pcHi-C, ATAC–seq, RNA-seq, scRNA-seq and HypR-seq) are available under Gene Expression Omnibus (GEO) accession number GSE173316. Data can be visualized on the WashU Epigenome Browser using the following session bundle ID: b62e2e70-f64c-11ec-9287-3f211c43000c. Source data are provided with this paper.

Code availability

A copy of the custom code used for data analysis in the CRISRPi/HyPR-seq screen is released on Zenodo68 (https://doi.org/10.5281/zenodo.8206584).

References

  1. Nott, A. et al. Brain cell type-specific enhancer-promoter interactome maps and disease-risk association. Science 366, 1134–1139 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Neuner, S. M., Tcw, J. & Goate, A. M. Genetic architecture of Alzheimer’s disease. Neurobiol. Dis. 143, 104976 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Song, M. et al. Mapping cis-regulatory chromatin contacts in neural cells links neuropsychiatric disorder risk variants to target genes. Nat. Genet. 51, 1252–1262 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Song, M. et al. Cell-type-specific 3D epigenomes in the developing human cortex. Nature 587, 644–649 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Martin, P. et al. Capture Hi-C reveals novel candidate genes and complex long-range interactions with related autoimmune risk loci. Nat. Commun. 6, 10069 (2015).

    CAS  PubMed  Google Scholar 

  6. Martin, J. S. et al. HUGIn: Hi-C unifying genomic interrogator. Bioinformatics 33, 3793–3795 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Ousman, S. S. & Kubes, P. Immune surveillance in the central nervous system. Nat. Neurosci. 15, 1096–1101 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Gjoneska, E. et al. Conserved epigenomic signals in mice and humans reveal immune basis of Alzheimer’s disease. Nature 518, 365–369 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Schwartzentruber, J. et al. Genome-wide meta-analysis, fine-mapping and integrative prioritization implicate new Alzheimer’s disease risk genes. Nat. Genet. 53, 392–402 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Bellenguez, C. et al. New insights into the genetic etiology of Alzheimer’s disease and related dementias. Nat. Genet. 54, 412–436 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Douvaras, P. et al. Directed differentiation of human pluripotent stem cells to microglia. Stem Cell Rep. 8, 1516–1524 (2017).

    CAS  Google Scholar 

  12. Gosselin, D. et al. An environment-dependent transcriptional network specifies human microglia identity. Science 356, eaal3222 (2017).

    PubMed  PubMed Central  Google Scholar 

  13. Abud, E. M. et al. iPSC-derived human microglia-like cells to study neurological diseases. Neuron 94, 278–293 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Heinz, S. et al. Transcription elongation can affect genome 3D structure. Cell 174, 1522–1536 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Carlin, A. F. et al. Deconvolution of pro- and antiviral genomic responses in Zika virus-infected and bystander macrophages. Proc. Natl Acad. Sci. USA 115, E9172–E9181 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Levine, K.S. et al. Virus exposure and neurodegenerative disease risk across national biobanks. Neuron 111, 1086–1093 (2023).

    CAS  PubMed  Google Scholar 

  17. Sayed, F. A. et al. AD-linked R47H-TREM2 mutation induces disease-enhancing microglial states via AKT hyperactivation. Sci. Transl. Med. 13, eabe3947 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Olah, M. et al. Single cell RNA sequencing of human microglia uncovers a subset associated with Alzheimer’s disease. Nat. Commun. 11, 6129 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Mathys, H. et al. Temporal tracking of microglia activation in neurodegeneration at single-cell resolution. Cell Rep. 21, 366–380 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Dolan, M. J. et al. Exposure of iPSC-derived human microglia to brain substrates enables the generation and manipulation of diverse transcriptional states in vitro. Nat. Immunol. 24, 1382–1390 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Bottcher, C. et al. Human microglia regional heterogeneity and phenotypes determined by multiplexed single-cell mass cytometry. Nat. Neurosci. 22, 78–90 (2019).

    PubMed  Google Scholar 

  22. Zhou, Y. et al. Human and mouse single-nucleus transcriptomics reveal TREM2-dependent and TREM2-independent cellular responses in Alzheimer’s disease. Nat. Med. 26, 131–142 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Mostafavi, S. et al. A molecular network of the aging human brain provides insights into the pathology and cognitive decline of Alzheimer’s disease. Nat. Neurosci. 21, 811–819 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Morabito, S. et al. Single-nucleus chromatin accessibility and transcriptomic characterization of Alzheimer’s disease. Nat. Genet. 53, 1143–1155 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Kosoy, R. et al. Genetics of the human microglia regulome refines Alzheimer’s disease risk loci. Nat. Genet. 54, 1145–1154 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Ray, J. et al. Chromatin conformation remains stable upon extensive transcriptional changes driven by heat shock. Proc. Natl Acad. Sci. USA 116, 19431–19439 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Jin, F. et al. A high-resolution map of the three-dimensional chromatin interactome in human cells. Nature 503, 290–294 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Kiani, K., Sanford, E. M., Goyal, Y. & Raj, A. Changes in chromatin accessibility are not concordant with transcriptional changes for single-factor perturbations. Mol. Syst. Biol. 18, e10979 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Calderon, D. et al. Landscape of stimulation-responsive chromatin across diverse human immune cells. Nat. Genet. 51, 1494–1505 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Ghavi-Helm, Y. et al. Enhancer loops appear stable during development and are associated with paused polymerase. Nature 512, 96–100 (2014).

    CAS  PubMed  Google Scholar 

  31. Proitsi, P. et al. Alzheimer’s disease susceptibility variants in the MS4A6A gene are associated with altered levels of MS4A6A expression in blood. Neurobiol. Aging 35, 279–290 (2014).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Kichaev, G. et al. Integrating functional data to prioritize causal variants in statistical fine-mapping studies. PLoS Genet. 10, e1004722 (2014).

    PubMed  PubMed Central  Google Scholar 

  34. Pickrell, J. K. Joint analysis of functional genomic data and genome-wide association studies of 18 human traits. Am. J. Hum. Genet. 94, 559–573 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Mathys, H. et al. Single-cell transcriptomic analysis of Alzheimer’s disease. Nature 570, 332–337 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Marshall, J. L. et al. HyPR-seq: single-cell quantification of chosen RNAs via hybridization and sequencing of DNA probes. Proc. Natl Acad. Sci. USA 117, 33404–33413 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Fernandes, S. et al. SHIP1 deficiency in inflammatory bowel disease is associated with severe Crohn’s disease and peripheral T cell reduction. Front. Immunol. 9, 1100 (2018).

    PubMed  PubMed Central  Google Scholar 

  38. Fu, Q. et al. SHIP1 inhibits cell growth, migration, and invasion in nonsmall cell lung cancer through the PI3K/AKT pathway. Oncol. Rep. 41, 2337–2350 (2019).

    CAS  PubMed  Google Scholar 

  39. Hamaoui, D. & Subtil, A. ATG16L1 functions in cell homeostasis beyond autophagy. FEBS J. 289, 1779–1800 (2022).

    CAS  PubMed  Google Scholar 

  40. Dixon, J. R. et al. Chromatin architecture reorganization during stem cell differentiation. Nature 518, 331–336 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Anzalone, A. V. et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 576, 149–157 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Noy, P. J. et al. TspanC8 tetraspanins and A disintegrin and metalloprotease 10 (ADAM10) interact via their extracellular regions: evidence for distinct binding mechanisms for different TspanC8 proteins. J. Biol. Chem. 291, 3145–3157 (2016).

    CAS  PubMed  Google Scholar 

  43. Kleinberger, G. et al. TREM2 mutations implicated in neurodegeneration impair cell surface transport and phagocytosis. Sci. Transl. Med. 6, 243ra286 (2014).

    Google Scholar 

  44. Zhong, L. et al. Soluble TREM2 ameliorates pathological phenotypes by modulating microglial functions in an Alzheimer’s disease model. Nat. Commun. 10, 1365 (2019).

    PubMed  PubMed Central  Google Scholar 

  45. Ewers, M. et al. Increased soluble TREM2 in cerebrospinal fluid is associated with reduced cognitive and clinical decline in Alzheimer’s disease. Sci.Transl. Med. 11, eaav6221 (2019).

    PubMed  PubMed Central  Google Scholar 

  46. Franzmeier, N. et al. Higher CSF sTREM2 attenuates ApoE4-related risk for cognitive decline and neurodegeneration. Mol. Neurodegener. 15, 57 (2020).

    PubMed  PubMed Central  Google Scholar 

  47. Hu, W. T. et al. Higher CSF sTNFR1-related proteins associate with better prognosis in very early Alzheimer’s disease. Nat. Commun. 12, 4001 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Haining, E. J. et al. The TspanC8 subgroup of tetraspanins interacts with A disintegrin and metalloprotease 10 (ADAM10) and regulates its maturation and cell surface expression. J. Biol. Chem. 287, 39753–39765 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Jansen, I. E. et al. Genome-wide meta-analysis identifies new loci and functional pathways influencing Alzheimer’s disease risk. Nat. Genet. 51, 404–413 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Kunkle, B. W. et al. Genetic meta-analysis of diagnosed Alzheimer’s disease identifies new risk loci and implicates Aβ, tau, immunity and lipid processing. Nat. Genet. 51, 414–430 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Choi, S.W. & O’Reilly, P.F. PRSice-2: Polygenic Risk Score software for biobank-scale data. Gigascience 8, giz082 (2019).

    PubMed  PubMed Central  Google Scholar 

  52. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Korsunsky, I. et al. Fast, sensitive and accurate integration of single-cell data with Harmony. Nat. Methods 16, 1289–1296 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  56. Corces, M. R. et al. An improved ATAC–seq protocol reduces background and enables interrogation of frozen tissues. Nat. Methods 14, 959–962 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Rosen, J. D. et al. HPRep: quantifying reproducibility in HiChIP and PLAC-seq datasets. Curr. Issues Mol. Biol. 43, 1156–1170 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    PubMed  PubMed Central  Google Scholar 

  59. Ross-Innes, C. S. et al. Differential oestrogen receptor binding is associated with clinical outcome in breast cancer. Nature 481, 389–393 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Cairns, J., Orchard, W. R., Malysheva, V. & Spivakov, M. Chicdiff: a computational pipeline for detecting differential chromosomal interactions in Capture Hi-C data. Bioinformatics 35, 4764–4766 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Yu, G., Wang, L. G., Han, Y. & He, Q. Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS 16, 284–287 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Jiang, M. Z. et al. Canonical correlation analysis for multi-omics: application to cross-cohort analysis. PLoS Genet. 19, e1010517 (2023).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Huang, L. et al. TOP-LD: a tool to explore linkage disequilibrium with TOPMed whole-genome sequence data. Am. J. Hum. Genet 109, 1175–1181 (2022).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Sun, Q. et al. From GWAS variant to function: a study of approximately 148,000 variants for blood cell traits. HGG Adv. 3, 100063 (2022).

    CAS  PubMed  Google Scholar 

  65. Young, A. M. H. et al. A map of transcriptional heterogeneity and regulatory variation in human microglia. Nat. Genet. 53, 861–868 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Van Buren, E. et al. TWO-SIGMA: a novel two-component single cell model-based association method for single-cell RNA-seq data. Genet. Epidemiol. 45, 142–153 (2021).

    CAS  PubMed  Google Scholar 

  67. Zhang, B. et al. Integrated systems approach identifies genetic nodes and networks in late-onset Alzheimer’s disease. Cell 153, 707–720 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Yang, X. et al. Functional characterization of Alzheimer’s disease genetic variants in microglia. Zenodo. https://doi.org/10.5281/zenodo.8206584 (2023).

  69. Perez, A. R. et al. GuideScan software for improved single and paired CRISPR guide RNA design. Nat. Biotechnol. 35, 347–349 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Hsu, J. Y. et al. PrimeDesign software for rapid and simplified design of prime editing guide RNAs. Nat. Commun. 12, 1034 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Van de Geijn, B., McVicker, G., Gilad, Y. & Pritchard, J. K. WASP: allele-specific software for robust molecular quantitative trait locus discovery. Nat. Methods 12, 1061–1063 (2015).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank B. Ren (University of California, San Diego) for sharing genome-wide pcHi-C probes, J. Engreitz (Stanford University) for sharing HyPR-seq gRNA detection barcodes and R. Corces (Gladstone Institutes) for his valuable advice on ATAC–seq analysis. This work was supported by the National Institutes of Health (NIH; grants R01AG057497 and RF1AG079557 to Y.S. and L.G.; R56AG079271 and R01AG079271 to Y.S., L.G. and Y.L.; R01EY027789 and UM1HG009402 to Y.S.; R01AG072758 and R01AG054214 to L.G.; U01HG011720, P50HD103573, R01MH125236 and R01MH123724 to Y.L.), the Hillblom Foundation and the American Federation for Aging Research New Investigator Award in AD (to Y.S). This work was made possible in part by NIH grants P30DK063720, and S101S10OD021822-01 to the UCSF Parnassus Flow Cytometry Core. Sequencing was performed at the UCSF CAT, supported by UCSF PBBR, RRP IMIA, and NIH (grant 1S10OD028511-01).

Author information

Authors and Affiliations

Authors

Contributions

Y.S., L.G. and Y.L. conceived and supervised the study. X.Y., H.Y., I.R.J., X.Z., C.C., W.L., M.Y.W., X.R., X.C., M.S., M.R. and C.S.Y.C. performed experiments under the supervision of Y.S., L.G. and R.P. X.Y., J.W., W. L. B.L., H.L, C.C. N.E., E.V.B., I.R.J. and M.J. performed computational analysis under the supervision of Y.S. and Y.L. Y.S. and D.L. performed scRNA-seq with the supervision of C.J.Y. X.Y., Y.L. and Y.S. analyzed and interpreted the data, and prepared the manuscript with input from all other authors.

Corresponding authors

Correspondence to Yun Li or Yin Shen.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Peer review

Peer review information

Peer reviewer reports are available.

Additional information

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

Extended data

Extended Data Fig. 1 Characterization of the hPSC-derived microglia-like cells.

a, PRS of each donor are shown with respect to PRS for individuals of matched continental ancestry from the 1000 Genomes Project (1000G). For WTC11, we used 1000G East Asian (EAS), and H1 European (EUR). The red dashed lines represent PRS for WTC11 and H1. b, The yield of IBA-1 or TMEM119 positive microglia is represented by the number of immunostaining positive cells divided by the total number of cells. Six coverslips from 3 independent differentiations were used for statistics. Boxplots indicate the median and interquartile range. Whiskers mark the 5th and 95th percentiles. c, Marker gene expressions are displayed in the UMAP of scRNA-seq. Percentage of positively expressed cells are calculated by pct.exp in the Seurat package. d, Representative contour plots depicting FACS gating strategy. Cells were separated from debris of various sizes based on the forward scatter area (FSC-A) and side scatter area (SSC-A). Two singlet gates were applied using the width and height metrics of the side scatter (SSC-H versus SSC-W) and forward scatter (FSC-H versus FSC-W). Latex beads-FITC signals are shown for all singlets.

Source data

Extended Data Fig. 2 Transcriptome analysis of hPSC-derived microglia-like cells with other cell types.

a, RNA-seq replicates were hierarchically clustered according to gene expression distances using DESeq2 (left). PCA plot displaying all samples (right). b, PCA plots of RNA-seq comparisons between hPSC-microglia differentiated with multiple protocols11,13 and primary microglia in vitro and in vivo12 and other cell types3. c, Heatmap showing scRNA-seq analysis of cell type-specific and shared IFNß stimulation responsive genes in microglia and peripheral myeloid cells. d, Examples of genes highly expressed (top 5) or lowly expressed (bottom 3) in microglia compared to peripheral myeloid cells. e, Examples of microglia-specific IFNß responsive genes. f, Top enriched GO terms of microglia specific IFNß responsive genes. Enriched GO terms are ranked by the percentage of total microglia-specific genes in the given GO term. The counts of enriched genes and adjusted P value for multiple comparisons were reported. Expanded lists of enriched GO terms are available in Supplementary Table 3.

Extended Data Fig. 3 Enrichment analysis of IFNβ responsive genes compared to DAM feature genes.

a, Barplots show the q values and enrichment scores of GSEA results for IFNβ responsive genes, in comparison with published datasets17,18,23,24,25,26 (dash line, q = 0.05). IFNβ responsive genes are highly enriched in multiple clusters of DAM by Olah et al.18, including the C4 cluster, representing cells with activated IFN signaling, the C7 cluster, representing cells expressing DAM genes, as well as C5 and C6, representing cells expressing genes related to anti-inflammatory responses. The C2 cluster, representing homeostatic microglia which are more likely derived from the temporal neocortex of younger temporal lobe epilepsy patients compared to the other homeostatic population, is also enriched for IFNβ responsive genes. 4 clusters are not enriched for IFN responsive genes, including C1, a homeostatic population shared by all brain regions in all donors, C3, cells with enriched expression of cellular stress genes, C8, cells enriched for respiratory electron transport, and C9 enriched with genes of cell cycle. In addition, IFNβ responsive genes are enriched in microglia samples associated with AD from 5 additional studies, including microglia samples in the human-MG4 cluster and the mouse-MG4 cluster, which are most enriched with DAM genes among all clusters from Sayed et al.17, the MG0 cluster (highly represented in AD microglia) compared to the MG1 cluster (control microglia) from Zhou et al.23, and AD DAM DEGs from Mostafavi, et al.24, Kosoy et al.26, and Morabito et al.25. b, UMAP plot visualizing integration of 3,038 WTC11-microglia scRNA-seq with 4,126 primary microglia snRNA-seq from Morabito et al. Cells are colored by sample origins. c, UMAP plot visualizing joint clustering splitted by donor condition (AD/control) or treatment (IFNβ stimulation/control). d, Cell proportions of each cluster splitted by donor condition or treatment. e, Cell proportion fold change in AD vs control or IFNβ stimulation vs control for 3 major clusters using monte-carlo/permutation test. Data are shown as mean ± s.d. (n_permutations = 1000).

Source data

Extended Data Fig. 4 Integrative analysis of chromatin accessibility, chromatin interactions, and gene expression.

a, Heatmap with pairwise correlations and hierarchical clustering of read densities at the set of unified open chromatin peaks for ATAC-seq datasets (left panel). PCA plot of ATAC-seq comparisons between hPSC-derived microglia-like cells, primary microglia12 and other cell types3 (right panel). b, Upset plot showing overlapping peaks of ATAC-seq datasets in WTC11 (hPSC), excitatory neuron, macrophage, microglia ex vivo, microglia in vitro, microglia derived from WTC11 (control and IFNβ stimulated) and microglia derived from H1 (control and IFNβ stimulated). c, Heatmap of Jaccard Index for pairwise overlap among the 9 ATAC-seq datasets in (b). Two-sided chi-squared tests on pairwise overlapping all led to P values less than 2.2e-16. d, Motif enrichment analysis for 93 TSS overlapping DARs in response to IFNß treatment. P values from HOMER and corresponding TF expression levels are shown. e, Heat map with pairwise similarity based on reproducibility analysis for pcHi-C replicates using HPRep (left). Heatmap of the Jaccard index for comparison of chromatin interaction profile in primary microglia, neuron, oligodendrocyte1 and hPSC derived microglia (right). f, Upset plot showing most of the IFNß stimulation responsive genes (total 3,811 including 1,460 down-regulated and 2,351 up-regulated genes) are not associated with DARs or DCRs. g, Pairwise canonical variable (CV) plots for all samples: (left) RNA-seq CV1 vs. ATAC-seq CV1; (middle) RNA-seq CV1 vs pcHi-C CV1; (right) pcHi-C CV1 vs ATAC-seq CV1. h, Volcano plot showing DEGs upon IFNß stimulation in microglia with a cutoff of adjusted P < 0.05 and absolute log2(fold change) > 0.5. MS4A6A gene is labeled.

Extended Data Fig. 5 Summarized results of CRISPRi and scRNA-seq analysis of cCREs with prioritized AD variants.

a–d, In all examples, tested cCREs are highlighted with orange or brown boxes. gRNAs targeting cCREs with AD variants are shown as red vertical lines. Genes expressed in microglia and exhibiting expression changes upon perturbation are shown with red labels. Distributions of relative gene expression levels are shown in violin plots where circles mark the median, and the black bars mark the upper and lower quantiles. Each dot represents one single cell. Number of cells are indicated in Supplementary Table 7d. P values are calculated by comparing gene expression between cells infected with control gRNAs and cells infected with gRNAc targeting cCREs using two-sided two-sample t-test and adjusted by Benjamini-Hochberg FDR multiple testing correction. Adjusted P values (FDR) are labeled. (a) TREM2 locus, including TREM2, NFYA and OARD1. (b) RIN3 locus, including RIN3, CPSF2, LGMN and NDUFB1. (c) BIN1 locus, including BIN1, IWS1, MAP3K2 and ERCC3. (d) PICALM locus, including PICALM, EDD and TMEM126A. Notably at the TREM2 locus, microglia receiving both TSS gRNA2 and cCRE1 showed enhanced downregulation of TREM2 compared to cells with TSS gRNA2 alone. e, Gene expression levels after CRIPSRi targeting cCREs at BIN1 and RIN3 loci with 2 gRNAs in WTC11-derived microglia-like cells. P values calculated using two-sided two-sample t-test (n = 3). Boxplots indicate the median and interquartile range. Whiskers mark the 5th and 95th percentiles.

Source data

Extended Data Fig. 6 Functional validation of AD risk cCREs under control and IFNβ stimulated conditions.

a, CRIPSRi validation on cCREs at INPP5D, BIN1, RIN3 and TREM2 loci in WTC11 microglia-like cells treated with IFNβ. P values calculated using two-sided two-sample t-test. Three independent replicates per condition and two sgRNAs per replicate were used for each experiment. Boxplots indicate the median and interquartile range. Whiskers mark the 5th and 95th percentiles. b, Scatter plot showing the fold change of cCRE perturbation in control or IFNβ treated condition. The Pearson correlation coefficient and its P value are reported. Linear regression line (black) with 95% confident interval (gray shade) are plotted. c, Genome browser snapshot showing the INPP5D locus containing a cCRE with prioritized AD variants and gRNAs for perturbation in single cell analysis. Genes expressed in microglia at this locus (red labels) are analyzed. Green boxes highlight the cCRE and promoters of neighboring genes. d, Down-regulation of INPP5D, GIGYF2, ATG16L1, and EIF4E2 by perturbing the cCRE region are confirmed by bulk CRISPRi followed by RT-qPCRs. P values are calculated with two-sided two sample t-test. Three independent replicates per condition and two sgRNAs per replicate were used for each experiment. Boxplots indicate the median and interquartile range. Whiskers mark the 5th and 95th percentiles.

Source data

Extended Data Fig. 7 Phenotypic analysis of hPSC-derived microglia-like cells under synergistic inhibition of TREM2 enhancer and promoter.

a, FACS analysis of proliferation for WTC11- and H1-derived microglia-like cells perturbed with synergistic inhibition of TREM2 enhancer and promoter in both control and IFNβ stimulated conditions. b, Representative contour plots of Ki-67 FITC FACS gating strategy. Cells were separated from debris based on the forward scatter area and side scatter area. Two singlet gates were applied using the width and height metrics of the side scatter and forward scatter. Ki-67 FITC signals are shown for all singlets. c, Negative control population using microglia not stained with Ki-67 antibody. d, FACS analysis of phagocytosis capacity for WTC11- and H1-derived microglia-like cells perturbed with synergistic inhibition of TREM2 enhancer and promoter in both control and IFNβ stimulated conditions. e, Representative contour plots of Latex beads-FITC FACS gating strategy. Cells were separated from debris based on the forward scatter area and side scatter area. Two singlet gates were applied using the width and height metrics of the side scatter and forward scatter. Latex beads-FITC signals are shown for all singlets. f, Negative control population using microglia not incubated with Latex beads.

Extended Data Fig. 8 Validation of prioritized AD variants by allelic analysis.

a, The total expression levels of TSPAN14 are elevated in microglia-like cells derived from H1 with prime editing rs7922621 (A/C to C/C), but not in microglia with prime editing at rs7910643 (A/G to G/G). P values are calculated with two-sided paired t-test (dash line indicating the pairing within each differentiation batch, n = 5). Each dot represents one biological replicate. b, Representative results from sanger sequencing display WTC11 rs7922621 wildtype (A/A) and KI clones (A/C). c. The ratios of allelic expression of TSPAN14 decrease in microglia-like cells derived from KI clones (rs7922621, A/C) compared to those derived from wildtype clones rs7922621 (A/A). P values calculated using two-sided two-sample t-test (n = 6). Boxplots indicate the median and interquartile range. Whiskers mark the 5th and 95th percentiles. d, Representative contour plots of ADAM10 FACS gating strategy. Cells were separated from debris based on the forward scatter area and side scatter area. Two singlet gates were applied using the width and height metrics of the side scatter and forward scatter. Live cells are selected based on SYTOX Blue signal and ADAM10 signals are shown for all live singlets. e, Violin plot of log10(ADAM10 intensity) in flow cytometry for WT controls, rs7922621 (G/G) edited cells, rs7910643 (C/C) edited cells across all replicates. P values are calculated using one-sided two-sample Wilcoxon test on ADAM10 levels for all cells compared to WT control within each replicate (n = 5). The horizontal line indicates the median. f, RT-qPCR experiments show that editing rs7922621 from A/C to C/C does not affect the expression levels of TSPAN15, TSPAN17, or TSPAN33 in microglia. P values are calculated with two-sided paired t-test (dash line indicating the pairing within each differentiation batch, n = 5). Each dot represents one biological replicate.

Source data

Extended Data Fig. 9 Phenotypic analysis of rs7922621 prime-edited WTC11- and H1-derived microglia-like cells.

FACS analysis of a, proliferation and b, phagocytosis capacity of WTC11 and H1 wild type and rs7922621 prime-edited clones derived microglia-like cells in both control and IFNβ stimulated conditions.

Supplementary information

Source data

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 3

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. 8

Statistical source data.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, X., Wen, J., Yang, H. et al. Functional characterization of Alzheimer’s disease genetic variants in microglia. Nat Genet 55, 1735–1744 (2023). https://doi.org/10.1038/s41588-023-01506-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-023-01506-8

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research