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.

  • Letter
  • Published:

Spatial and temporal heterogeneity of mouse and human microglia at single-cell resolution

An Author Correction to this article was published on 27 March 2019

This article has been updated

Abstract

Microglia have critical roles not only in neural development and homeostasis, but also in neurodegenerative and neuroinflammatory diseases of the central nervous system1,2,3,4. These highly diverse and specialized functions may be executed by subsets of microglia that already exist in situ, or by specific subsets of microglia that develop from a homogeneous pool of cells on demand. However, little is known about the presence of spatially and temporally restricted subclasses of microglia in the central nervous system during development or disease. Here we combine massively parallel single-cell analysis, single-molecule fluorescence in situ hybridization, advanced immunohistochemistry and computational modelling to comprehensively characterize subclasses of microglia in multiple regions of the central nervous system during development and disease. Single-cell analysis of tissues of the central nervous system during homeostasis in mice revealed specific time- and region-dependent subtypes of microglia. Demyelinating and neurodegenerative diseases evoked context-dependent subtypes of microglia with distinct molecular hallmarks and diverse cellular kinetics. Corresponding clusters of microglia were also identified in healthy human brains, and the brains of patients with multiple sclerosis. Our data provide insights into the endogenous immune system of the central nervous system during development, homeostasis and disease, and may also provide new targets for the treatment of neurodegenerative and neuroinflammatory pathologies.

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: Identification of spatiotemporal subclasses of microglia in the mouse.
Fig. 2: Characteristics of subsets of microglia during development.
Fig. 3: Specific disease-associated populations of microglia with distinct kinetics during demyelination and neurodegeneration.
Fig. 4: Presence of distinct subclasses of microglia in brains of healthy humans and patients with multiple sclerosis.

Similar content being viewed by others

Data availability

Raw data for mouse and human single cell RNA-sequencing have been deposited in the Gene Expression Omnibus, and are available at the following accession numbers: GSE120629 (mouse), GSE120747 (mouse) and GSE124335 (human). All other data are available from the corresponding author on reasonable request.

Change history

  • 27 March 2019

    In this Letter, Dominic Grün and Sagar have been added to the author list (affiliated with Max-Planck-Institute of Immunology and Epigenetics (MPI-IE), Freiburg, Germany). The author list, ‘Author contribution’ and ‘Acknowledgements’ sections have been corrected online. See accompanying Amendment.

References

  1. Prinz, M. & Priller, J. Microglia and brain macrophages in the molecular age: from origin to neuropsychiatric disease. Nat. Rev. Neurosci. 15, 300–312 (2014).

    Article  CAS  Google Scholar 

  2. Colonna, M. & Butovsky, O. Microglia function in the central nervous system during health and neurodegeneration. Annu. Rev. Immunol. 35, 441–468 (2017).

    Article  CAS  Google Scholar 

  3. Herz, J., Filiano, A. J., Smith, A., Yogev, N. & Kipnis, J. Myeloid cells in the central nervous system. Immunity 46, 943–956 (2017).

    Article  CAS  Google Scholar 

  4. Prinz, M., Erny, D. & Hagemeyer, N. Ontogeny and homeostasis of CNS myeloid cells. Nat. Immunol. 18, 385–392 (2017).

    Article  CAS  Google Scholar 

  5. Kierdorf, K. et al. Microglia emerge from erythromyeloid precursors via Pu.1- and Irf8-dependent pathways. Nat. Neurosci. 16, 273–280 (2013).

    Article  CAS  Google Scholar 

  6. Hagemeyer, N. et al. Microglia contribute to normal myelinogenesis and to oligodendrocyte progenitor maintenance during adulthood. Acta Neuropathol. 134, 441–458 (2017).

    Article  Google Scholar 

  7. Matcovitch-Natan, O. et al. Microglia development follows a stepwise program to regulate brain homeostasis. Science 353, aad8670 (2016).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  9. Jordao, M. J. et al. Single-cell profiling identifies myeloid cell subsets with distinct fates during neuroinflammation. Science 363, eaat7554 (2019).

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

    Article  CAS  Google Scholar 

  11. Grabert, K. et al. Microglial brain region-dependent diversity and selective regional sensitivities to aging. Nat. Neurosci. 19, 504–516 (2016).

    Article  CAS  Google Scholar 

  12. Herman, J. S., Sagar, & Grün, D. FateID infers cell fate bias in multipotent progenitors from single-cell RNA-seq data. Nat. Methods 15, 379–386 (2018).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  14. Tay, T. L. et al. A new fate mapping system reveals context-dependent random or clonal expansion of microglia. Nat. Neurosci. 20, 793–803 (2017).

    Article  CAS  Google Scholar 

  15. Mildner, A. et al. Microglia in the adult brain arise from Ly-6ChiCCR2+ monocytes only under defined host conditions. Nat. Neurosci. 10, 1544–1553 (2007).

    Article  CAS  Google Scholar 

  16. Benveniste, E. N. Role of macrophages/microglia in multiple sclerosis and experimental allergic encephalomyelitis. J. Mol. Med. (Berl.) 75, 165–173 (1997).

    Article  CAS  Google Scholar 

  17. Brück, W. et al. Monocyte/macrophage differentiation in early multiple sclerosis lesions. Ann. Neurol. 38, 788–796 (1995).

    Article  Google Scholar 

  18. Lawson, L. J., Perry, V. H. & Gordon, S. Turnover of resident microglia in the normal adult mouse brain. Neuroscience 48, 405–415 (1992).

    Article  CAS  Google Scholar 

  19. Doorn, K. J. et al. Brain region-specific gene expression profiles in freshly isolated rat microglia. Front. Cell. Neurosci. 9, 84 (2015).

    Article  Google Scholar 

  20. Wang, Y. et al. IL-34 is a tissue-restricted ligand of CSF1R required for the development of Langerhans cells and microglia. Nat. Immunol. 13, 753–760 (2012).

    Article  CAS  Google Scholar 

  21. Greter, M. et al. Stroma-derived interleukin-34 controls the development and maintenance of Langerhans cells and the maintenance of microglia. Immunity 37, 1050–1060 (2012).

    Article  CAS  Google Scholar 

  22. De Biase, L. M. et al. Local cues establish and maintain region-specific phenotypes of basal ganglia microglia. Neuron 95, 341–356.e6 (2017).

    Article  Google Scholar 

  23. Ayata, P. et al. Epigenetic regulation of brain region-specific microglia clearance activity. Nat. Neurosci. 21, 1049–1060 (2018).

    Article  CAS  Google Scholar 

  24. Hammond, T. R. et al. Single-cell RNA sequencing of microglia throughout the mouse lifespan and in the injured brain reveals complex cell-state changes. Immunity 50, 253–271.e6 (2019).

    Article  CAS  Google Scholar 

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

    Google Scholar 

  26. Liao, Y., Smyth, G. K. & Shi, W. The Subread aligner: fast, accurate and scalable read mapping by seed-and-vote. Nucleic Acids Res. 41, e108 (2013).

    Article  Google Scholar 

  27. Gu, Z., Eils, R. & Schlesner, M. Complex heatmaps reveal patterns and correlations in multidimensional genomic data. Bioinformatics 32, 2847–2849 (2016).

    Article  CAS  Google Scholar 

  28. Hashimshony, T. et al. CEL-Seq2: sensitive highly-multiplexed single-cell RNA-seq. Genome Biol. 17, 77 (2016).

    Article  Google Scholar 

  29. Li, H. & Durbin, R. Fast and accurate long-read alignment with Burrows–Wheeler transform. Bioinformatics 26, 589–595 (2010).

    Article  Google Scholar 

  30. Grün, D., Kester, L. & van Oudenaarden, A. Validation of noise models for single-cell transcriptomics. Nat. Methods 11, 637–640 (2014).

    Article  Google Scholar 

  31. Grün, D. et al. Single-cell messenger RNA sequencing reveals rare intestinal types. Nature 525, 251–255 (2015).

    Article  ADS  Google Scholar 

  32. Goldmann, T. et al. USP18 lack in microglia causes destructive interferonopathy of the mouse brain. EMBO J. 34, 1612–1629 (2015).

    Article  CAS  Google Scholar 

  33. Raasch, J. et al. IκB kinase 2 determines oligodendrocyte loss by non-cell-autonomous activation of NF-κB in the central nervous system. Brain 134, 1184–1198 (2011).

    Article  Google Scholar 

  34. Kuhn, A., Luthi-Carter, R. & Delorenzi, M. Cross-species and cross-platform gene expression studies with Bioconductor-compliant R package ‘annotationTools’. BMC Bioinformatics 9, 26 (2008).

    Article  Google Scholar 

  35. 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).

    Article  CAS  Google Scholar 

  36. Lucchinetti, C. et al. Heterogeneity of multiple sclerosis lesions: implications for the pathogenesis of demyelination. Ann. Neurol. 47, 707–717 (2000).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank M. Ditter, E. Barleon, T. el Gaz and J. Bodinek-Wersing for technical assistance; C. Gross for proofreading; N. Mossadegh for helpful discussions; and A. Zeisel and S. Linnarsson (Karolinska Institute, Sweden) for sharing data and support. The samples and data used for this project were kindly provided by the West German Biobank. T.M. was supported by the KANAE Foundation and the Japan Society for the Promotion of Science. M.P. is supported by the Sobek Foundation, the Ernst-Jung Foundation, the DFG (SFB 992, SFB1160, SFB/TRR167, Reinhart Koselleck Grant), the Ministry of Science, Research and Arts, Baden-Wuerttemberg (Sonderlinie ‘Neuroinflammation’) and the BMBF-funded competence network of multiple sclerosis (KKNMS). This study was supported by the DFG under Germany’s Excellence Strategy (CIBSS – EXC-2189 – Project ID390939984). S. is supported by the DFG (GR4980) the Behrens-Weise-Foundation, and the Max Planck Society. D.G. is supported by the Max Planck Society. J.P. is supported by the DFG (SFB/TRR167 B05 and B07), BIH CRG 2a, and the UK DRI Momentum Award. C. Stadelmann is supported by the DFG (STA1389/2-1), the Deutsche Multiple Sklerose Gesellschaft (DMSG), the National MS Society of the USA (NMSS) and the Hertie Foundation. C.B. is supported by the DFG (SFB/TRR167 B05).

Reviewer information

Nature thanks Arnold Kriegstein and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

T.M., R.S., O.S., C.B., L.A., C. Scheiwe, S.N., P.K., G.v.L., V.A.C., P.C.R., A.M., U.S. and R.G. conducted experiments and analysed the data. M.P., C. Stadelmann and J.P. analysed the data, contributed to the in vivo studies and provided mice or reagents. D.G. and S. established the sequencing platform, supervised the sequencing and helped with bioinformatics. T.M. and M.P. supervised the project and wrote the manuscript.

Corresponding author

Correspondence to Marco Prinz.

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 figures and tables

Extended Data Fig. 1 Comprehensive analysis of microglial diversity by scRNA-seq.

a, Illustration depicting the workflow for the isolation of microglia from different regions of the CNS of embryonic (E16.5), juvenile (3 weeks of age) and adult (16 weeks of age) mice during homeostasis and during pathology (FNX or cuprizone-mediated demyelination for scRNA-seq). b, t-SNE plot showing 3,826 analysed microglial cells from different conditions tested in this study. Each dot represents a single cell. c, t-SNE plot depicting 13 clusters for all of the different conditions (3,826 cells). Colours represent each cluster. d, t-SNE plots depicting single microglia from the replicates from different regions of the CNS of individual embryos, juvenile and adult mice and diseased mice (3,826 cells).

Extended Data Fig. 2 Subpopulations of microglia with distinct gene expression during development, in mice.

a, Distribution of Malat1 gene expression in a t-SNE plot (2,966 microglia cells). Colour keys represent the respective expression levels. The right cloud represents the embryonic microglial population, whereas the left cloud combines both juvenile and adult microglia as shown in Fig. 1a. b, Left, single-molecule fluorescent in situ hybridization (smFISH) for Malat1 and Cx3cr1 shows the kinetics of Malat1+ microglia during development. Scale bar, 10 μm. Representative images out of two adult mice investigated are shown. Yellow and white arrowheads indicate Malat1+Cx3cr1+ microglia and Malat1+Cx3cr1 non-microglial cells, respectively. Right, frequency of Malat1+ microglia in the forebrain or cortex during development. Bar represents mean ± s.e.m. of 120 cells studied, from 3 animals per time point. c, t-SNE plot of Ctsd and Lamp1 gene expression; these genes were enriched in the C1 and C2 clusters, as shown in Fig. 1b. d, t-SNE plot of Eef1a1 and Rpl4 gene expression; these genes were enriched in the C6 cluster, as shown in Fig. 1b.

Extended Data Fig. 3 CST3 is enriched in adult microglia.

ac, Representative sections of the cortex from adult mice using immunofluoresence for CST3 (green), NeuN for neurons (red, a) and adenomatous polyposis coli (APC) for oligodendrocytes (red, b). The astrocyte marker ALDH1L1 (red, c), combined with CST3, was used on the hippocampal sections. Scale bars, 50 μm (overview), 20 μm (magnification). Representative images out of three mice investigated are shown. d, Quantification of CST3 immunoreactivity in the brain of the adult mouse. Bar represents mean ± s.e.m. of 3 animals (393 microglia, 1,817 neurons, 298 oligodendrocytes and 461 astrocytes).

Extended Data Fig. 4 Diverse clusters of microglia during demyelination and neurodegeneration.

a, Projection of 1,564 single microglia isolated from different regions of the CNS during homeostasis or FNX or cuprizone treatment, shown as a t-SNE plot. b, Heat map of top differentially regulated genes that were up- or downregulated in each cluster. The genes with the highest differentially expression are highlighted. c, t-SNE plot exhibiting 13 clusters for the 1,564 individual microglia isolated from different regions of the CNS during homeostasis or FNX or cuprizone treatment.

Extended Data Fig. 5 Molecular characterization of subpopulations of microglia during demyelination and remyelination.

ac, e, g, i, j, t-SNE plots showing expression of Axl, Igf1, Lyz2, Itgax, Gpnmb and Apoc1 (a); Fam20c, Ccl6, Psat1 and Ank (b); Fn1 (c); Spp1 (e); Cybb (g); H2-Aa and H2-Ab1 (i); and Ccl4 (j) transcripts after cuprizone challenge (3,826 microglial cells). Genes shown in a were upregulated in both demyelination and remyelination, whereas genes depicted in b, c and e or g and i were increased in demyelination-associated clusters C12, or in the remyelination- associated clusters C13, respectively. Colour keys represent the respective expression levels. d, Left and middle, smFISH for Fn1 and Cx3cr1 reveals subpopulations of microglia after a five-week cuprizone treatment in the corpus callosum. ‘1’ indicates Fn1+Cx3cr1+ microglia (yellow arrowheads, Fn1 mRNA); ‘2’ indicates Fn1Cx3cr1+ microglia. Scale bar, 10 μm (overview), 3 μm (inset). Representative images out of three mice investigated are shown. Non-filled arrowheads in the image of control mice indicate non-specific signals. Right, percentage of Fn1+ Cx3cr1+ microglia in the corpus callosum. Bar represents mean ± s.e.m. of 3 animals (168 cells investigated). f, Left and middle, smFISH for Spp1 and Cx3cr1 reveals subpopulations of microglia after a five-week cuprizone treatment in the corpus callosum. ‘1’ indicates Spp1+Cx3cr1+ microglia (yellow arrowheads, Spp1 mRNA); ‘2’ indicates Spp1Cx3cr1+ microglia. Scale bars, 10 μm (overview), 3 μm (insets). Representative images out of three mice investigated are shown. Right, percentage of Spp1+Cx3cr1+ microglia in the corpus callosum. Bar represents mean ± s.e.m. of 3 animals (165 investigated cells). h, Left and middle, smFISH for Cybb and Cx3cr1 reveals subpopulations of microglia after a five-week cuprizone treatment in the corpus callosum. ‘1’ indicates Cybb+Cx3cr1+ microglia (yellow arrowheads, Cybb mRNA); ‘2’ indicates CybbCx3cr1+ microglia. Scale bars, 10 μm (overview) and 3 μm (insets). Representative images out of three mice investigated are shown. Right, percentage of Cybb+Cx3cr1+ microglia in the corpus callosum. Bar represents mean ± s.e.m. of 3 animals (165 investigated cells). The colour key represents the expression levels.

Extended Data Fig. 6 Microglial subtypes in healthy human brains.

a, t-SNE plot of 1,180 human microglia showing the distribution of individual microglia from 5 patients. Each dot represents a single cell. Different colours indicate different patients. b, t-SNE plot of 1,180 individual human microglia isolated from 5 individual non-pathological brains depicts four major clusters (HHu-C1 to HHu-C4). Each dot represents a single cell. Colours correspond to each cluster. c, Heat map showing the distribution of the healthy human clusters in each individual patient. d, Bar graphs representing the relative abundance of microglial cells in the respective clusters from five individual non-pathological brains. Colours represent distinct clusters. e, Heat map of the top differentially regulated genes that were up- or downregulated in each cluster. f, t-SNE plots for CST3, P2RY13, CCL4 and EGR2 mRNA expression. CCL4 and EGR2 are enriched in the cluster HHu-C4.

Extended Data Fig. 7 Detailed neuropathological characterization of lesions of multiple sclerosis.

Histology of the brains of patients with multiple sclerosis (labelled as patients MS-1 to MS-5) using haematoxylin and eosin (HE), luxol fast blue (LFB-PAS), 2′,3′-cyclic-nucleotide 3′-phosphodiesterase (CNPase) and myelin basic protein (MBP) for myelin, human leukocyte antigen DR isotype (HLA-DR) and CD68 for myeloid cells, CD3 for T cells, CD20 for B cells and Bielschowsky (Biel) for axons. Scale bar, 50 μm. Lesions are typical early active multiple-sclerosis lesions, according to the standard classification system36.

Extended Data Fig. 8 Molecular profile of subsets of microglia during multiple sclerosis.

a, t-SNE plots representing the core signature genes for lymphocytes (TRAC, TRBC2, CD52 and IL32), myeloid cells (ITGAM, MS4A6A, TYROBP and CD14) and monocytes (CCR2, PLAC8, CLEC12A and FCN1) in the brains of patients with multiple sclerosis. Colour keys reflect the expression levels. bf, t-SNE plots of genes that are enriched in clusters Hu-C5 to Hu-C7 (b), Hu-C2 (c), Hu-C3 (d), Hu-C4 (e) and Hu-C8 (f) are shown (1,602 microglia cells). Colour codes represent expression levels. g, t-SNE plots depicting genes that are upregulated in the clusters Hu-C2, Hu-C3 and Hu-C8. Colour codes represent expression levels. h, t-SNE plots of genes that were upregulated in the disease-associated subsets of microglia in the mouse demyelination model, but not in the microglia in the brains of patients with multiple sclerosis. Colour codes represent expression levels. i, Immunofluorescence images for TMEM119 and IBA1 in healthy brains or the brains of patients with multiple sclerosis. Arrowheads indicate TMEM119+IBA1+ cells (filled) in the healthy brains, and TMEM119IBA1+ microglia (open) during multiple sclerosis. Representative images out of four mice investigated, per condition, are shown. Scale bar, 50 μm. j, Representative immunofluorescence images for IBA1+MRP14 (indicating microglia) and IBA1+MRP14+ cells (representing infiltrating early activated monocytes) in the healthy brain and brains of patients with multiple sclerosis. Insets show microglia (top row) and monocytes (bottom row) in the multiple-sclerosis lesion. Right, quantification. Bars represent means ± s.e.m. (n = 4 for each condition). Each symbol represents one patient. Scale bars, 50 μm (overview), 4 μm (inset).

Extended Data Fig. 9 Canonical correlation analysis of scRNA-seq data from mouse and human microglia.

a, Canonical correlation analysis (Seurat alignment procedure) visualizing shared correlation structures (that is, canonical correlation vectors (CCs)) between mouse and human datasets. Each dot represents single cell. b, CC plot of cells assigned as mouse clusters C7 to C13 and human clusters Hu-C1 to Hu-C10. Mouse demyelination-related clusters of microglia (C12 and C13) are transcriptionally close to human multiple sclerosis-associated clusters of microglia (Hu-C2, Hu-C3 and Hu-C8). Each dot represents single cell. c, Violin plots depicting a shared gene correlation structure that is conserved between mouse and human clusters. Values for CC1 and CC2 vectors for individual cells are shown.

Extended Data Fig. 10

Graphical abstract of experimental findings.

Supplementary information

Supplementary Figure

This file contains Supplementary Figure 1: FACS gating strategy for microglia isolation. CNS cells were gated for G1 and G2 (singlets), followed by being gated for living cell (G3, fixable viability dye), CD45intCD11b+ (G4), Ly6C-Ly6G- (G5), and CD206- (G6).

Reporting Summary

Supplementary Table

This file contains Supplementary Table split between mouse and human data. The mouse data shows information on mice and cells used in each scRNA-seq analysis including genotype, strain, sex, condition, age, CNS region and cell numbers. The human data shows details on patients and cells used in each scRNAseq analysis including sex, condition, age, CNS region and cell numbers.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Masuda, T., Sankowski, R., Staszewski, O. et al. Spatial and temporal heterogeneity of mouse and human microglia at single-cell resolution. Nature 566, 388–392 (2019). https://doi.org/10.1038/s41586-019-0924-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-0924-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