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.

  • Resource
  • Published:

A single-cell atlas of entorhinal cortex from individuals with Alzheimer’s disease reveals cell-type-specific gene expression regulation

Abstract

There is currently little information available about how individual cell types contribute to Alzheimer’s disease. Here we applied single-nucleus RNA sequencing to entorhinal cortex samples from control and Alzheimer’s disease brains (n = 6 per group), yielding a total of 13,214 high-quality nuclei. We detail cell-type-specific gene expression patterns, unveiling how transcriptional changes in specific cell subpopulations are associated with Alzheimer’s disease. We report that the Alzheimer’s disease risk gene APOE is specifically repressed in Alzheimer’s disease oligodendrocyte progenitor cells and astrocyte subpopulations and upregulated in an Alzheimer’s disease-specific microglial subopulation. Integrating transcription factor regulatory modules with Alzheimer’s disease risk loci revealed drivers of cell-type-specific state transitions towards Alzheimer’s disease. For example, transcription factor EB, a master regulator of lysosomal function, regulates multiple disease genes in a specific Alzheimer’s disease astrocyte subpopulation. These results provide insights into the coordinated control of Alzheimer’s disease risk genes and their cell-type-specific contribution to disease susceptibility. These results are available at http://adsn.ddnetbio.com.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Single-nuclei sequencing of human entorhinal cortex recapitulates cell-type-specific marker genes and cell-type-specific changes in Alzheimer’s disease.
Fig. 2: Single-nuclei sequencing of human entorhinal cortex uncovers high cellular heterogeneity with each cell type.
Fig. 3: Single-nuclei sequencing of human Alzheimer’s disease and control entorhinal cortex reveals homeostatic, Alzheimer’s disease-specific, and shared ontological cell subclusters.
Fig. 4: Alzheimer’s disease genes identified by GWAS show specific gene expression patterns across cell types and cell-type subclusters.
Fig. 5: GRN analysis predicts transcription factors for conversion of control to Alzheimer’s disease subcluster signatures.

Similar content being viewed by others

Data availability

All single-cell RNA sequencing data are available from the Gene Expression Omnibus (GEO) under the accession number GSE138852. Data can also be visualized via the interactive web application at adsn.ddnetbio.com. Single-cell gene expression data and metadata can also be downloaded directly via adsn.ddnetbio.com.

Code availability

Code is available from the authors by reasonable request.

References

  1. Huang, K.-L. et al. A common haplotype lowers PU.1 expression in myeloid cells and delays onset of Alzheimer’s disease. Nat. Neurosci. 20, 1052–1061 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Lambert, J. C. et al. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer’s disease. Nat. Genet. 45, 1452–1458 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. 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 

  4. Füger, P. et al. Microglia turnover with aging and in an Alzheimer’s model via long-term in vivo single-cell imaging. Nat. Neurosci. 20, 1371–1376 (2017).

    PubMed  Google Scholar 

  5. Whitehouse, P. J. et al. Alzheimer’s disease and senile dementia: loss of neurons in the basal forebrain. Science 215, 1237–1239 (1982).

    CAS  PubMed  Google Scholar 

  6. Zhong, S. et al. A single-cell RNA-seq survey of the developmental landscape of the human prefrontal cortex. Nature 555, 524–528 (2018).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Wilhelmsson, U. et al. Injury leads to the appearance of cells with characteristics of both microglia and astrocytes in mouse and human brain. Cereb. Cortex 27, 3360–3377 (2017).

    PubMed  Google Scholar 

  9. Zeisel, A. et al. Brain structure. Cell types in the mouse cortex and hippocampus revealed by single-cell RNA-seq. Science 347, 1138–1142 (2015).

    CAS  PubMed  Google Scholar 

  10. Darmanis, S. et al. A survey of human brain transcriptome diversity at the single cell level. Proc. Natl Acad. Sci. USA 112, 7285–7290 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Miceli, F., et al. KCNQ3-Related Disorders. in Gene Reviews (ed. Adam, M. P.) https://www.ncbi.nlm.nih.gov/books/NBK201978/ (Univ. Washington, 2014).

  12. Ebermann, I. et al. GPR98 mutations cause Usher syndrome type 2 in males. J. Med. Genet. 46, 277–280 (2009).

    CAS  PubMed  Google Scholar 

  13. Grubman, A., Choo, X. Y., Chew, G., Ouyang, J. F. & Sun, G. Mouse and human microglial phenotypes in Alzheimer’s disease are controlled by amyloid plaque phagocytosis through Hif1α. Preprint at biorXiv https://www.biorxiv.org/content/10.1101/639054v1 (2019).

  14. Veereshwarayya, V., Kumar, P., Rosen, K. M., Mestril, R. & Querfurth, H. W. Differential effects of mitochondrial heat shock protein 60 and related molecular chaperones to prevent intracellular β-amyloid-induced inhibition of complex IV and limit apoptosis. J. Biol. Chem. 281, 29468–29478 (2006).

    CAS  PubMed  Google Scholar 

  15. De Strooper, B. & Karran, E. The cellular phase of Alzheimer’s disease. Cell 164, 603–615 (2016).

    PubMed  Google Scholar 

  16. Xie, H. et al. Rapid cell death is preceded by amyloid plaque-mediated oxidative stress. Proc. Natl Acad. Sci. USA 110, 7904–7909 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Liddelow, S. A. et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature 541, 481–487 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Krasemann, S. et al. The TREM2-APOE pathway drives the transcriptional phenotype of dysfunctional microglia in neurodegenerative diseases. Immunity 47, 566–581.e9 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Keren-Shaul, H. et al. A unique microglia type associated with restricting development of Alzheimer’s disease. Cell 169, 1276–1290.e17 (2017).

    CAS  PubMed  Google Scholar 

  20. De Rossi, P. et al. Predominant expression of Alzheimer’s disease-associated BIN1 in mature oligodendrocytes and localization to white matter tracts. Mol. Neurodegener. 11, 59 (2016).

    PubMed  PubMed Central  Google Scholar 

  21. Savvaki, M. et al. The expression of TAG-1 in glial cells is sufficient for the formation of the juxtaparanodal complex and the phenotypic rescue of tag-1 homozygous mutants in the CNS. J. Neurosci. 30, 13943–13954 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Bartzokis, G. Age-related myelin breakdown: a developmental model of cognitive decline and Alzheimer’s disease. Neurobiol. Aging 25, 5–18 (2004).

    CAS  PubMed  Google Scholar 

  23. Behrendt, G. et al. Dynamic changes in myelin aberrations and oligodendrocyte generation in chronic amyloidosis in mice and men. Glia 61, 273–286 (2013).

    PubMed  Google Scholar 

  24. Lake, B. B. et al. Neuronal subtypes and diversity revealed by single-nucleus RNA sequencing of the human brain. Science 352, 1586–1590 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. John Lin, C.-C. et al. Identification of diverse astrocyte populations and their malignant analogs. Nat. Neurosci. 20, 396–405 (2017).

    CAS  PubMed  Google Scholar 

  26. Mi, S. et al. LINGO-1 negatively regulates myelination by oligodendrocytes. Nat. Neurosci. 8, 745–751 (2005).

    CAS  PubMed  Google Scholar 

  27. Santoro, M. et al. Expression profile of long non-coding RNAs in serum of patients with multiple sclerosis. J. Mol. Neurosci. 59, 18–23 (2016).

    CAS  PubMed  Google Scholar 

  28. Fallin, M. D. et al. Bipolar I disorder and schizophrenia: a 440–single-nucleotide polymorphism screen of 64 candidate genes among Ashkenazi Jewish case-parent trios. Am. J. Hum. Genet. 77, 918–936 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Hall, C. N., Klein-Flügge, M. C., Howarth, C. & Attwell, D. Oxidative phosphorylation, not glycolysis, powers presynaptic and postsynaptic mechanisms underlying brain information processing. J. Neurosci. 32, 8940–8951 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Romito-DiGiacomo, R. R., Menegay, H., Cicero, S. A. & Herrup, K. Effects of Alzheimer’s disease on different cortical layers: the role of intrinsic differences in Abeta susceptibility. J. Neurosci. 27, 8496–8504 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Bis, J. C. et al. Whole exome sequencing study identifies novel rare and common Alzheimer’s-Associated variants involved in immune response and transcriptional regulation. Mol. Psychiatry https://doi.org/10.1038/s41380-018-0112-7 (2018).

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

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Kajiho, H. et al. Characterization of RIN3 as a guanine nucleotide exchange factor for the Rab5 subfamily GTPase Rab31. J. Biol. Chem. 286, 24364–24373 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. He, Z.-Y., Li, L., Wang, Y.-Z., Liu, X. & Yuan, L.-Y. Associations between thromboxane A synthase 1 gene polymorphisms and the risk of ischemic stroke in a Chinese Han population.Neural Regen. Res. 13, 463 (2018).

    PubMed  PubMed Central  Google Scholar 

  35. Diener, H. C. et al. European Stroke Prevention Study 2. Dipyridamole and acetylsalicylic acid in the secondary prevention of stroke. J. Neurol. Sci. 143, 1–13 (1996).

    CAS  PubMed  Google Scholar 

  36. Paris, D. et al. Inhibition of Alzheimer’s beta-amyloid induced vasoactivity and proinflammatory response in microglia by a cGMP-dependent mechanism. Exp. Neurol. 157, 211–221 (1999).

    CAS  PubMed  Google Scholar 

  37. Alzheimer’s Association. 2018 Alzheimer’s disease facts and figures. Alzheimers Dement. 14, 367–429 (2018).

    Google Scholar 

  38. Raj, T. et al. Integrative transcriptome analyses of the aging brain implicate altered splicing in Alzheimer’s disease susceptibility. Nat. Genet. 50, 1584–1592 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Tan, M. G. et al. Genome wide profiling of altered gene expression in the neocortex of Alzheimer’s disease. J. Neurosci. Res. 88, 1157–1169 (2010).

    CAS  PubMed  Google Scholar 

  40. Blalock, E. M., Buechel, H. M., Popovic, J., Geddes, J. W. & Landfield, P. W. Microarray analyses of laser-captured hippocampus reveal distinct gray and white matter signatures associated with incipient Alzheimer’s disease. J. Chem. Neuroanat. 42, 118–126 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Lin, Y.-T. et al. APOE4 causes widespread molecular and cellular alterations associated with Alzheimer’s disease phenotypes in human iPSC-derived brain cell types. Neuron 98, 1294 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Bartzokis, G. et al. Apolipoprotein E genotype and age-related myelin breakdown in healthy individuals: implications for cognitive decline and dementia. Arch. Gen. Psychiatry 63, 63–72 (2006).

    CAS  PubMed  Google Scholar 

  43. Chapuis, J. et al. Increased expression of BIN1 mediates Alzheimer genetic risk by modulating tau pathology. Mol. Psychiatry 18, 1225–1234 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Yu, L. et al. Association of brain DNA methylation in SORL1, ABCA7, HLA-DRB5, SLC24A4, and BIN1 with pathological diagnosis of Alzheimer disease. JAMA Neurol. 72, 15–24 (2015).

    PubMed  PubMed Central  Google Scholar 

  45. Lambert, J.-C. et al. Genome-wide haplotype association study identifies the FRMD4A gene as a risk locus for Alzheimer’s disease. Mol. Psychiatry 18, 461–470 (2013).

    CAS  PubMed  Google Scholar 

  46. Hokama, M. et al. Altered expression of diabetes-related genes in Alzheimer’s disease brains: the Hisayama study. Cereb. Cortex 24, 2476–2488 (2014).

    PubMed  Google Scholar 

  47. Shijo, M. et al. Association of adipocyte enhancer-binding protein 1 with Alzheimer’s disease pathology in human hippocampi. Brain Pathol. 28, 58–71 (2018).

    CAS  PubMed  Google Scholar 

  48. Maynard, M. A. et al. Human HIF-3α4 is a dominant-negative regulator of HIF-1 and is down-regulated in renal cell carcinoma. FASEB J. 19, 1396–1406 (2005).

    CAS  PubMed  Google Scholar 

  49. Martini-Stoica, H. et al. TFEB enhances astroglial uptake of extracellular tau species and reduces tau spreading. J. Exp. Med. 215, 2355–2377 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. de Toledo-Morrell, L., Goncharova, I., Dickerson, B., Wilson, R. S. & Bennett, D. A. From healthy aging to early Alzheimer’s disease: in vivo detection of entorhinal cortex atrophy. Ann. N. Y. Acad. Sci. 911, 240–253 (2000).

    PubMed  Google Scholar 

  51. McKenzie, A. T. et al. Brain cell type specific gene expression and co-expression network architectures. Sci. Rep. 8, 8868 (2018).

    PubMed  PubMed Central  Google Scholar 

  52. Habib, N. et al. Massively parallel single-nucleus RNA-seq with DroNc-seq. Nat. Methods 14, 955–958 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Tirosh, I. et al. Single-cell RNA-seq supports a developmental hierarchy in human oligodendroglioma. Nature 539, 309–313 (2016).

    PubMed  PubMed Central  Google Scholar 

  54. Zhu, Y., Wang, L., Yin, Y. & Yang, E. Systematic analysis of gene expression patterns associated with postmortem interval in human tissues. Sci. Rep. 7, 5435 (2017).

    PubMed  PubMed Central  Google Scholar 

  55. Butler, A., Hoffman, P., Smibert, P., Papalexi, E. & Satija, R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nat. Biotechnol. 36, 411–420 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Lummertz da Rocha, E. et al. Reconstruction of complex single-cell trajectories using CellRouter. Nat. Commun. 9, 892 (2018).

    PubMed  PubMed Central  Google Scholar 

  57. van der Maaten, L. & Hinton, G. Visualizing data using t-SNE. J. Mach. Learn. Res. 9, 2579–2605 (2008).

    Google Scholar 

  58. Lambert, S. A. et al. The human transcription factors. Cell 175, 598–599 (2018).

    CAS  PubMed  Google Scholar 

  59. Kolde, R. Pheatmap: pretty heatmaps. R package version 61, 926 (2012).

    Google Scholar 

  60. Wickham, H. ggplot2: Elegant Graphics for Data Analysis. (Springer: 2016)..

  61. Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Buniello, A. et al. The NHGRI-EBI GWAS Catalog of published genome-wide association studies, targeted arrays and summary statistics 2019. Nucleic Acids Res. 47, D1005–D1012 (2019).

    CAS  PubMed  Google Scholar 

  63. Skene, N. G. & Grant, S. G. N. Identification of vulnerable cell types in major brain disorders using single cell transcriptomes and expression weighted cell type enrichment. Front. Neurosci. 10, 16 (2016).

    PubMed  PubMed Central  Google Scholar 

  64. McKenzie, A. T., Katsyv, I., Song, W.-M., Wang, M. & Zhang, B. DGCA: a comprehensive R package for differential gene correlation analysis. BMC Syst. Biol. 10, 106 (2016).

    PubMed  PubMed Central  Google Scholar 

  65. Falcon, S. & Gentleman, R. Using GOstats to test gene lists for GO term association. Bioinformatics 23, 257–258 (2007).

    CAS  PubMed  Google Scholar 

  66. Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. B 57, 289–300 (1995).

    Google Scholar 

  67. Li, H. et al. Reference component analysis of single-cell transcriptomes elucidates cellular heterogeneity in human colorectal tumors. Nat. Genet. 49, 708–718 (2017).

    CAS  PubMed  Google Scholar 

  68. 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 

  69. Conway, J. R., Lex, A. & Gehlenborg, N. UpSetR: an R package for the visualization of intersecting sets and their properties. Bioinformatics 33, 2938–2940 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors acknowledge Flowcore and Micromon, Monash University, for the provision of instrumentation, training, and technical support. Tissues were received from the Victorian Brain Bank, supported by The Florey Institute of Neuroscience and Mental Health, The Alfred and the Victorian Institute of Forensic Medicine, and funded in part by Parkinson’s Victoria and MND Victoria. The Australian Regenerative Medicine Institute is supported by grants from the State Government of Victoria and the Australian Government. This work was supported by a Dementia Australia Research Foundation Grant to A.G., J.M.P., and E.P., a Monash Network of Excellence grant to J.M.P., E.P., and O.J.L.R., and a Yulgilbar Foundation grant to A.G. and G.S. A.G. was supported by a National Health and Medical Research Council (NHMRC) and Australian Research Council (ARC) Dementia Research Development Fellowship (GNT1097461). J.M.P. was supported by a Sylvia-Charles Viertel Fellowship. O.J.L.R. and J.F.O. were supported by a Singapore National Research Foundation Competitive Research Programme (NRF-CRP20-2017-0002). S.B. was supported by a National Health and Medical Research Council (NHMRC) and Australian Research Council (ARC) Dementia Research Development Fellowship (GNT1111206). This work was supported by a Sylvia and Charles Viertel Senior Medical Research Fellowship to R.L. and and a Howard Hughes Medical Institute International Research Scholarship to R.L. We acknowledge S. Freytag (University of Western Australia) for her advice regarding disentangling individuals from mixed-individual libraries.

Author information

Authors and Affiliations

Authors

Contributions

A.G. and J.M.P. conceived the study and designed experiments, and together with E.P. and O.J.L.R., designed the bioinformatics analyses. A.G., G.S., and X.Y.C. performed nuclei isolation and fluorescence-activated cell sorting (FACS). C.M. performed the pathological assessment of human control and Alzheimer’s disease cases. G.C. and E.P. performed GWAS integration, and CellRouter and network analyses. J.F.O. and O.J.L.R. performed cell-type and cell-subcluster identification and performed differential gene expression and GSEA analyses. J.F.O. and O.J.L.R. developed the shiny web interface. J.P., R.S., S.B., D.V.L., D.P., and R.L. worked up the protocol for single-nuclei sequencing from the human brain. A.G., G.C., J.F.O., O.J.L.R., E.P., and J.M.P. wrote the manuscript. All authors approved of, and contributed to, the final version of the manuscript.

Corresponding authors

Correspondence to Owen J. L. Rackham, Enrico Petretto or Jose M. Polo.

Ethics declarations

Competing interests

O.J.L.R. and J.M.P. are co-inventors of the patent (WO/2017/106932) and are co-founders, shareholders, and directors of Mogrify Ltd., a cell therapy company. All other 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 Gating strategy for FACS isolation of single nuclei.

Nuclei isolated using the EZ-Prep kit from entorhinal cortex of AD and control patients were FACS sorted on BD Influx using 70 μm nozzle, 21–22 psi. Single DAPI+ events were considered nuclei.

Extended data Fig. 2 Single nuclei metadata and analysis of hybrid cells.

a, cell proportion plots by library for eight libraries showing exclusion of two libraries due to high neuronal count and comparison of cell type proportions recovered in this study compared to cell proportions in other single nuclei studies24 and52. Cell proportions are shown by sequencing library, where this information is available. b, Frequency histogram of number of Unique Molecular Identifiers (UMIs) across all 8 cell type groups. c, Frequency histogram of number of detected genes across all 8 cell type groups. d, Barplot showing number of top two identified cell types in hybrid cell types in AD and control libraries as proportion. e, UpSetR69 plot showing the BRETIGEA number of distinct and overlapping cell type markers across the six major cell types - microglia, astrocyte, neuron, oligodendrocyte, OPC, and endothelial cells.

Source data

Extended data Fig. 3 Comparison of differentially expressed genes (DEGs) within microglia, OPCs and oligodendrocytes with Mathys et al. 2019.

a, d, g, Venn diagram showing overlap of DEGs detected in Mathys et al. (n = 24 AD-pathology vs n = 24 no-pathology individuals) and those identified in the present study (n = 6 AD vs n = 6 control individuals) in microglia (a), OPCs (d) and oligodendrocytes (g), and the table shows the number of and concordance of dysregulation in AD of genes within the overlap of DEGs in both studies. Hypergeometric test was used to test for significance of overlap. b, e, h, List of concordant and discordant overlapping DEGs in microglia (b), OPCs (e) and oligodendrocytes (h). c, f, i, Significant gene ontology (GO) terms for each gene set comparison in microglia (c), OPCs (f) and oligodendrocytes (i). Multiple testing correction was done using the Benjamini–Hochberg method. Enrichment analysis was performed using hypergeometric test.

Extended data Fig. 4 Comparison of differentially expressed genes (DEGs) within astrocytes, excitatory neurons and inhibitory neurons with Mathys et al. 2019.

a, d, g, Venn diagram showing overlap of DEGs detected in Mathys et al. (n = 24 AD-pathology vs n = 24 no-pathology individuals) and those identified in the present study (n = 6 AD vs n = 6 control individuals) in astrocytes (a), excitatory neurons (d), and inhibitory neurons (g), and the table shows the number of and concordance of dysregulation in AD of genes within the overlap of DEGs in both studies. Hypergeometric test was used to test for significance of overlap. b, e, h, List of concordant and discordant overlapping DEGs in astrocytes (b), excitatory neurons (e), and inhibitory neurons (h). c, f, i, Significant gene ontology (GO) terms for each gene set comparison in astrocytes (c), excitatory neurons (f), and inhibitory neurons (i). Multiple testing correction was done using the Benjamini–Hochberg method. Enrichment analysis was performed using hypergeometric test.

Extended data Fig. 5 Analysis of sources of variation.

a, Box plot (center line: median, box limits: upper and lower quartiles, whiskers: 1.5x interquartile range, points outside whiskers: outliers) showing percentage variance explained by each covariate for genes that are differentially expressed between AD and Control groups in any of the identified cell types. b, Proportion of AD-vs-Control differentially expressed genes (DEGs) that are also individual-associated genes and genes that are only DE between AD and Control. c, Proportion of AD-vs-Control DEGs that are also sex-associated genes and genes that are only DE between AD and Control. The number of DEGs are given in parenthesis.

Source data

Extended data Fig. 6 Cell-type-specific and shared AD related changes.

a–i, Bubble plot showing the top 15 differentially expressed genes (DEGs) within the DEG1-DEG9 groups (see Fig. 1f), where the colour represents the relative gene expression and the bubble size is the proportion of cells in the group expressing the gene.

Extended data Fig. 7 AD related changes in excitatory and inhibitory neurons.

Scatter plot showing the relationship between cell-type-specific and common differentially expressed genes (DEGs) in control and AD, excitatory and inhibitory neurons. Cutoff for significance is abs(log fold change) > 0.5 and false discovery rate (FDR) < 0.01.

Source data

Extended data Fig. 8 Single nuclei sequencing of human AD and control entorhinal cortex reveals homeostatic, AD-specific and shared ontological cell subclusters.

a,e,i, Uniform Manifold Approximation and Projection (UMAP) visualization of subclusters of microglia (a), OPCs (e) and endothelial cells (i) showing b, f, j, the composition of cells in subclusters by disease state, c, g, k, hierarchical clustering and heatmap colored by single-cell gene expression of subcluster-specific genes (top eight genes were shown per cluster). d, h, l, Gene Set Enrichment Analysis (GSEA) results of subcluster specific genes coloured by normalised gene set enrichment scores for the gene ontologies shown in each cell subcluster.

Source data

Extended data Fig. 9 Comparison of signatures across subclusters.

Comparison of astrocyte signatures to ai;17 aii25. b, Comparison of neuronal subcluster signatures to24(Ex1–8, In1–8). c, Functional annotation of gene regulatory networks (GRNs) controlled by TFs in Fig. 5a. GRNs are shown in Extended data Fig. 10, ETS2 (n = 9 genes), GTF2IRD1 (n = 339 genes), HIF3A-neuron (n = 112 genes), HIF3A-OPC (n = 96 genes), KDM5B (n = 44 genes), MAX (n = 273 genes), NKX6–2 (n = 332 genes), RELA (n = 224 genes), SOX10 (n = 114 genes), and ZKSCAN1 (n = 37 genes). Multiple correction testing was done using the Benjamini–Hochberg method. Enrichment analysis was performed using the hypergeometric test in R package GOstats.

Source data

Extended data Fig. 10 Gene regulatory network analysis predicts transcription factors regulating GWAS genes for conversion of control to AD subcluster signatures.

Trajectories of the top transcription factors (TFs): NKX6–2, ETS2, ZKSCAN1, HIF3A, GTF2IRD1, RELA, SOX10, KDM5B and MAX, with downstream GWAS gene targets. For each cell type, the reported TFs-driven gene regulatory networks include GWAS gene hits, which are direct target genes of the TF and have the highest average log fold change in gene expression between the source subcluster/s and target subcluster/s. Each gene is colored according to its average log fold change between the source subcluster/s and target subcluster/s. Only significantly differentially expressed GWAS gene hits are reported (false discovery rate < 0.05).

Supplementary information

Source data

Source Data Fig. 1

Statistical Source Data

Source Data Fig. 2

Statistical Source Data

Source Data Fig. 3

Statistical Source Data

Source Data Fig. 5

Statistical Source Data

Source Data Extended Data Fig. 2

Statistical Source Data

Source Data Extended Data Fig. 5

Statistical Source Data

Source Data Extended Data Fig. 7

Statistical Source Data

Source Data Extended Data Fig. 8

Statistical Source Data

Source Data Extended Data Fig. 9

Statistical Source Data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grubman, A., Chew, G., Ouyang, J.F. et al. A single-cell atlas of entorhinal cortex from individuals with Alzheimer’s disease reveals cell-type-specific gene expression regulation. Nat Neurosci 22, 2087–2097 (2019). https://doi.org/10.1038/s41593-019-0539-4

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-019-0539-4

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