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:

Molecular characterization of selectively vulnerable neurons in Alzheimer’s disease

Abstract

Alzheimer’s disease (AD) is characterized by the selective vulnerability of specific neuronal populations, the molecular signatures of which are largely unknown. To identify and characterize selectively vulnerable neuronal populations, we used single-nucleus RNA sequencing to profile the caudal entorhinal cortex and the superior frontal gyrus—brain regions where neurofibrillary inclusions and neuronal loss occur early and late in AD, respectively—from postmortem brains spanning the progression of AD-type tau neurofibrillary pathology. We identified RORB as a marker of selectively vulnerable excitatory neurons in the entorhinal cortex and subsequently validated their depletion and selective susceptibility to neurofibrillary inclusions during disease progression using quantitative neuropathological methods. We also discovered an astrocyte subpopulation, likely representing reactive astrocytes, characterized by decreased expression of genes involved in homeostatic functions. Our characterization of selectively vulnerable neurons in AD paves the way for future mechanistic studies of selective vulnerability and potential therapeutic strategies for enhancing neuronal resilience.

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: AD progression differentially affects the cell-type composition of the EC and SFG.
Fig. 2: RORB-expressing excitatory neuron subpopulations in the EC are selectively vulnerable.
Fig. 3: Immunofluorescence of the EC validates selective vulnerability of RORB-expressing excitatory neurons.
Fig. 4: Inhibitory neuron subpopulations do not consistently show differences in resilience or vulnerability to AD progression.
Fig. 5: GFAPhigh astrocytes show signs of dysfunction in glutamate homeostasis and synaptic support.

Similar content being viewed by others

Data availability

The raw snRNA-seq sequencing data and unfiltered UMI count matrices are available on the Gene Expression Omnibus (GEO) under accession code GSE147528. Single-cell data after quality control is available for download at Synapse.org under the Synapse ID syn21788402. Post-quality-control data can also be explored interactively through the CellXGene platform at https://kampmannlab.ucsf.edu/ad-brain. Data from Mathys et al.14 was downloaded from Synapse under Synapse ID syn18485175.

Code availability

We provide the full bioinformatics pipeline for the analysis of snRNA-seq data in this paper at https://kampmannlab.ucsf.edu/ad-brain-analysis.

References

  1. Braak, H. & Braak, E. Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 82, 239–259 (1991).

    Article  PubMed  CAS  Google Scholar 

  2. Scholl, M. et al. PET imaging of tau deposition in the aging human brain. Neuron 89, 971–982 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Seeley, W. W., Crawford, R. K., Zhou, J., Miller, B. L. & Greicius, M. D. Neurodegenerative diseases target large-scale human brain networks. Neuron 62, 42–52 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  4. Braak, H. & Braak, E. Staging of Alzheimer’s disease-related neurofibrillary changes. Neurobiol. Aging 16, 271–278 (1995).

    Article  PubMed  CAS  Google Scholar 

  5. Price, J. L. et al. Neuron number in the entorhinal cortex and CA1 in preclinical Alzheimer disease. Arch. Neurol. 58, 1395–1402 (2001).

    Article  PubMed  CAS  Google Scholar 

  6. Stranahan, A. M. & Mattson, M. P. Selective vulnerability of neurons in layer II of the entorhinal cortex during aging and Alzheimer’s disease. Neural Plast. 2010, 108190 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Van Hoesen, G. W., Hyman, B. T. & Damasio, A. R. Entorhinal cortex pathology in Alzheimer’s disease. Hippocampus 1, 1–8 (1991).

    Article  PubMed  Google Scholar 

  8. Gomez-Isla, T. et al. Profound loss of layer II entorhinal cortex neurons occurs in very mild Alzheimer’s disease. J. Neurosci. 16, 4491–4500 (1996).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Braak, H. & Braak, E. The human entorhinal cortex: normal morphology and lamina-specific pathology in various diseases. Neurosci. Res. 15, 6–31 (1992).

    Article  PubMed  CAS  Google Scholar 

  10. Kordower, J. H. et al. Loss and atrophy of layer II entorhinal cortex neurons in elderly people with mild cognitive impairment. Ann. Neurol. 49, 202–213 (2001).

    Article  PubMed  CAS  Google Scholar 

  11. Drummond, E. & Wisniewski, T. Alzheimer’s disease: experimental models and reality. Acta Neuropathol. 133, 155–175 (2017).

    Article  PubMed  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  13. Hodge, R. D. et al. Conserved cell types with divergent features in human versus mouse cortex. Nature 573, 61–68 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Grubman, A. 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).

    Article  PubMed  CAS  Google Scholar 

  16. Nelson, P. T. et al. Correlation of Alzheimer disease neuropathologic changes with cognitive status: a review of the literature. J. Neuropathol. Exp. Neurol. 71, 362–381 (2012).

    Article  PubMed  Google Scholar 

  17. Hof, P. R. et al. Parvalbumin-immunoreactive neurons in the neocortex are resistant to degeneration in Alzheimer’s disease. J. Neuropathol. Exp. Neurol. 50, 451–462 (1991).

    Article  PubMed  CAS  Google Scholar 

  18. Fu, H. et al. A tau homeostasis signature is linked with the cellular and regional vulnerability of excitatory neurons to tau pathology. Nat. Neurosci. 22, 47–56 (2019).

    Article  PubMed  CAS  Google Scholar 

  19. Heinsen, H. et al. Quantitative investigations on the human entorhinal area: left-right asymmetry and age-related changes. Anat. Embryol. (Berl.) 190, 181–194 (1994).

    Article  PubMed  CAS  Google Scholar 

  20. Kobro-Flatmoen, A. & Witter, M. P. Neuronal chemo-architecture of the entorhinal cortex: a comparative review. Eur. J. Neurosci. 50, 3627–3662 (2019).

    Article  PubMed  Google Scholar 

  21. Naumann, R. K. et al. Conserved size and periodicity of pyramidal patches in layer 2 of medial/caudal entorhinal cortex. J. Comp. Neurol. 524, 783–806 (2016).

    Article  PubMed  CAS  Google Scholar 

  22. Ramsden, H. L., Surmeli, G., McDonagh, S. G. & Nolan, M. F. Laminar and dorsoventral molecular organization of the medial entorhinal cortex revealed by large-scale anatomical analysis of gene expression. PLoS Comput. Biol. 11, e1004032 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Jabaudon, D., Shnider, S. J., Tischfield, D. J., Galazo, M. J. & Macklis, J. D. RORbeta induces barrel-like neuronal clusters in the developing neocortex. Cereb. Cortex 22, 996–1006 (2012).

    Article  PubMed  Google Scholar 

  24. Oishi, K., Aramaki, M. & Nakajima, K. Mutually repressive interaction between Brn1/2 and Rorb contributes to the establishment of neocortical layer 2/3 and layer 4. Proc. Natl Acad. Sci. USA 113, 3371–3376 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Nakagawa, Y. & O’Leary, D. D. Dynamic patterned expression of orphan nuclear receptor genes RORalpha and RORbeta in developing mouse forebrain. Dev. Neurosci. 25, 234–244 (2003).

    Article  PubMed  CAS  Google Scholar 

  26. Marinaro, F., et al. Molecular and cellular pathology of monogenic Alzheimer’s disease at single cell resolution. Preprint at bioRxiv https://doi.org/10.1101/2020.07.14.202317 (2020).

  27. Liang, W. S. et al. Altered neuronal gene expression in brain regions differentially affected by Alzheimer’s disease: a reference data set. Physiol. Genomics 33, 240–256 (2008).

    Article  PubMed  CAS  Google Scholar 

  28. Franjic, D., et al. Molecular diversity among adult human hippocampal and entorhinal cells. Preprint at bioRxiv https://doi.org/10.1101/2019.12.31.889139 (2019).

  29. Ehrenberg, A. J., et al. A manual multiplex immunofluorescence method for investigating neurodegenerative diseases. J. Neurosci. Meth. https://doi.org/10.1016/j.jneumeth.2020.108708. (2020).

  30. Keren-Shaul, H. et al. A unique microglia type associated with restricting development of alzheimer’s disease. Cell 169, 1276–1290 e1217 (2017).

    Article  PubMed  CAS  Google Scholar 

  31. Srinivasan, K., et al. Alzheimer’s patient brain myeloid cells exhibit enhanced aging and unique transcriptional activation. Preprint at bioRxiv https://doi.org/10.1101/610345 (2019).

  32. Thrupp, N., et al. Single nucleus sequencing fails to detect microglial activation in human tissue. Preprint at bioRxiv https://doi.org/10.1101/2020.04.13.035386 (2020).

  33. Chen, W. T. et al. Spatial transcriptomics and in situ sequencing to study Alzheimer’s disease. Cell 182, 976–991 (2020).

    Article  PubMed  CAS  Google Scholar 

  34. Perez-Nievas, B. G. & Serrano-Pozo, A. Deciphering the Astrocyte reaction in Alzheimer’s disease. Front Aging Neurosci. 10, 114 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Simpson, J. E. et al. Microarray analysis of the astrocyte transcriptome in the aging brain: relationship to Alzheimer’s pathology and APOE genotype. Neurobiol. Aging 32, 1795–1807 (2011).

    Article  PubMed  CAS  Google Scholar 

  36. Sekar, S. et al. Alzheimer’s disease is associated with altered expression of genes involved in immune response and mitochondrial processes in astrocytes. Neurobiol. Aging 36, 583–591 (2015).

    Article  PubMed  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Laywell, E. D. et al. Enhanced expression of the developmentally regulated extracellular matrix molecule tenascin following adult brain injury. Proc. Natl Acad. Sci. USA 89, 2634–2638 (1992).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Zamanian, J. L. et al. Genomic analysis of reactive astrogliosis. J. Neurosci. 32, 6391–6410 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Anderson, M. A. et al. Astrocyte scar formation aids central nervous system axon regeneration. Nature 532, 195–200 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  41. Kampmann, M. A CRISPR approach to neurodegenerative diseases. Trends Mol. Med. 23, 483–485 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Kampmann, M. CRISPR-based functional genomics for neurological disease. Nat. Rev. Neurol. 16, 465–480 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Tian, R. et al. CRISPR interference-based platform for multimodal genetic screens in human iPSC-Derived neurons. Neuron 104, 239–255.e212 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Hof, P. R. & Morrison, J. H. Neocortical neuronal subpopulations labeled by a monoclonal antibody to calbindin exhibit differential vulnerability in Alzheimer’s disease. Exp. Neurol. 111, 293–301 (1991).

    Article  PubMed  CAS  Google Scholar 

  45. Hof, P. R., Cox, K. & Morrison, J. H. Quantitative analysis of a vulnerable subset of pyramidal neurons in Alzheimer’s disease: I. Superior frontal and inferior temporal cortex. J. Comp. Neurol. 301, 44–54 (1990).

    Article  PubMed  CAS  Google Scholar 

  46. Mikkonen, M., Alafuzoff, I., Tapiola, T., Soininen, H. & Miettinen, R. Subfield- and layer-specific changes in parvalbumin, calretinin and calbindin-D28K immunoreactivity in the entorhinal cortex in Alzheimer’s disease. Neuroscience 92, 515–532 (1999).

    Article  PubMed  CAS  Google Scholar 

  47. Montine, T. J. et al. National Institute on Aging-Alzheimer’s association guidelines for the neuropathologic assessment of Alzheimer’s disease: a practical approach. Acta Neuropathol. 123, 1–11 (2012).

    Article  PubMed  CAS  Google Scholar 

  48. Hughes, C. P., Berg, L., Danziger, W. L., Coben, L. A. & Martin, R. L. A new clinical scale for the staging of dementia. Br. J. Psychiatry 140, 566–572 (1982).

    Article  PubMed  CAS  Google Scholar 

  49. Grinberg, L. T. et al. Brain bank of the brazilian aging brain study group - a milestone reached and more than 1,600 collected brains. Cell Tissue Bank 8, 151–162 (2007).

    Article  PubMed  Google Scholar 

  50. Hyman, B. T. et al. National institute on Aging-Alzheimer’s association guidelines for the neuropathologic assessment of Alzheimer’s disease. Alzheimers Dement. 8, 1–13 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Suemoto, C. K. et al. Neuropathological diagnoses and clinical correlates in older adults in Brazil: A cross-sectional study. PLoS Med. 14, e1002267 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Cairns, N. J. et al. Neuropathologic diagnostic and nosologic criteria for frontotemporal lobar degeneration: consensus of the consortium for frontotemporal lobar degeneration. Acta Neuropathol. 114, 5–22 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Ferrer, I., Santpere, G. & van Leeuwen, F. W. Argyrophilic grain disease. Brain 131, 1416–1432 (2008).

    Article  PubMed  Google Scholar 

  54. Rodriguez, R. D. & Grinberg, L. T. Argyrophilic grain disease: An underestimated tauopathy. Dement. Neuropsychol. 9, 2–8 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  55. Rodriguez, R. D. et al. Argyrophilic grain disease: Demographics, clinical, and neuropathological features from a large autopsy study. J. Neuropathol. Exp. Neurol. 75, 628–635 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. Braak, H., Thal, D. R., Ghebremedhin, E. & Del Tredici, K. Stages of the pathologic process in Alzheimer disease: age categories from 1 to 100 years. J. Neuropathol. Exp. Neurol. 70, 960–969 (2011).

    Article  PubMed  CAS  Google Scholar 

  57. Mo, A. et al. Epigenomic signatures of neuronal diversity in the mammalian brain. Neuron 86, 1369–1384 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Tabula Muris, C. et al. Single-cell transcriptomics of 20 mouse organs creates a Tabula Muris. Nature 562, 367–372 (2018).

    Article  Google Scholar 

  59. Lun, A. T. L. et al. EmptyDrops: distinguishing cells from empty droplets in droplet-based single-cell RNA sequencing data. Genome Biol. 20, 63 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  60. Lun, A. T., Bach, K. & Marioni, J. C. Pooling across cells to normalize single-cell RNA sequencing data with many zero counts. Genome Biol. 17, 75 (2016).

    Article  PubMed  Google Scholar 

  61. Lun, A. T., McCarthy, D. J. & Marioni, J. C. A step-by-step workflow for low-level analysis of single-cell RNA-seq data with Bioconductor. F1000Res 5, 2122 (2016).

    PubMed  PubMed Central  Google Scholar 

  62. McCarthy, D. J., Campbell, K. R., Lun, A. T. & Wills, Q. F. Scater: pre-processing, quality control, normalization and visualization of single-cell RNA-seq data in R. Bioinformatics 33, 1179–1186 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  63. Stuart, T. et al. Comprehensive Integration of Single-Cell data. Cell 177, 1888–1902 e1821 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Johansen, N. & Quon, G. scAlign: a tool for alignment, integration, and rare cell identification from scRNA-seq data. Genome Biol. 20, 166 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  65. 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).

    Article  PubMed  CAS  Google Scholar 

  66. Amezquita, R. A. et al. Orchestrating single-cell analysis with bioconductor. Nat. Methods 17, 137–145 (2020).

    Article  PubMed  CAS  Google Scholar 

  67. Szklarczyk, D. et al. STRING v11: protein-protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 47, D607–D613 (2019).

    Article  PubMed  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Stelzer, G. et al. The GeneCards Suite: From gene data mining to disease genome sequence analyses. Curr. Protoc. Bioinforma. 54, 13031–313033 (2016).

    Article  Google Scholar 

  70. Arriza, J. L. et al. Functional comparisons of three glutamate transporter subtypes cloned from human motor cortex. J. Neurosci. 14, 5559–5569 (1994).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  71. Borden, L. A. et al. Cloning of the human homologue of the GABA transporter GAT-3 and identification of a novel inhibitor with selectivity for this site. Receptors Channels 2, 207–213 (1994).

    PubMed  CAS  Google Scholar 

  72. Gendreau, S. et al. A trimeric quaternary structure is conserved in bacterial and human glutamate transporters. J. Biol. Chem. 279, 39505–39512 (2004).

    Article  PubMed  CAS  Google Scholar 

  73. Häberle, J. et al. Congenital glutamine deficiency with glutamine synthetase mutations. N. Engl. J. Med. 353, 1926–1933 (2005).

    Article  PubMed  Google Scholar 

  74. Kawakami, H., Tanaka, K., Nakayama, T., Inoue, K. & Nakamura, S. Cloning and expression of a human glutamate transporter. Biochem. Biophys. Res. Commun. 199, 171–176 (1994).

    Article  PubMed  CAS  Google Scholar 

  75. Melzer, N., Biela, A. & Fahlke, C. Glutamate modifies ion conduction and voltage-dependent gating of excitatory amino acid transporter-associated anion channels. J. Biol. Chem. 278, 50112–50119 (2003).

    Article  PubMed  CAS  Google Scholar 

  76. Südhof, T. C. Synaptic neurexin complexes: A molecular code for the logic of neural circuits. Cell 171, 745–769 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  77. Pellissier, F., Gerber, A., Bauer, C., Ballivet, M. & Ossipow, V. The adhesion molecule Necl-3/SynCAM-2 localizes to myelinated axons, binds to oligodendrocytes and promotes cell adhesion. BMC Neurosci. 8, 90 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  78. González-Castillo, C., Ortuño-Sahagún, D., Guzmán-Brambila, C., Pallàs, M. & Rojas-Mayorquín, A. E. Pleiotrophin as a central nervous system neuromodulator, evidences from the hippocampus. Front. Cell Neurosci. 8, 443 (2014).

    PubMed  Google Scholar 

  79. Siddiqui, T. J. et al. An LRRTM4-HSPG complex mediates excitatory synapse development on dentate gyrus granule cells. Neuron 79, 680–695 (2013).

    Article  PubMed  CAS  Google Scholar 

  80. Heinsen, H., Arzberger, T. & Schmitz, C. Celloidin mounting (embedding without infiltration) - a new, simple and reliable method for producing serial sections of high thickness through complete human brains and its application to stereological and immunohistochemical investigations. J. Chem. Neuroanat. 20, 49–59 (2000).

    Article  PubMed  CAS  Google Scholar 

  81. Insausti, R. & Amaral, D. G. Entorhinal cortex of the monkey: IV. Topographical and laminar organization of cortical afferents. J. Comp. Neurol. 509, 608–641 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Insausti, R., Munoz-Lopez, M., Insausti, A. M. & Artacho-Perula, E. The human periallocortex: layer pattern in presubiculum, parasubiculum and entorhinal cortex. A Review. Front. Neuroanat. 11, 84 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  83. Rose, S. Vergleichende messungen im allocortex bei tier und mensch. J. Psychol. Neurol. 34, 250–255 (1927).

    Google Scholar 

  84. Fu, H. et al. Tau pathology induces excitatory neuron loss, grid cell dysfunction, and spatial memory deficits reminiscent of early Alzheimer’s disease. Neuron 93, 533–541.e535 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. Hevner, R. F. et al. Tbr1 regulates differentiation of the preplate and layer 6. Neuron 29, 353–366 (2001).

    Article  PubMed  CAS  Google Scholar 

  86. Ferrari, S. L. P. & Cribari-Neto, F. Beta regression for modelling rates and proportions. J. Appl. Stat. 31, 799–815 (2004).

    Article  Google Scholar 

Download references

Acknowledgements

We thank A. Pisco, A. Maynard, S. Darmanis and the MACA team at the Chan Zuckerberg Biohub for advice on analysis 10X library preparation and reagents. We thank members of the Kampmann laboratory (A. Samelson, X. Guo, R. Tian and B. Rooney) for feedback on the manuscript. This work was supported by National Institutes of Health (NIH) awards F30 AG066418 (to K.L.), K08 AG052648 (to S.S.), R56 AG057528 (to M.K. and L.T.G.), K24 AG053435 (to L.T.G.), U54 NS100717 (to L.T.G. and M

Author information

Authors and Affiliations

Authors

Contributions

K.L., E.L., L.T.G. and M.K. conceptualized and led the overall project. K.L., L.T.G. and M.K. wrote the manuscript, with input from all co-authors. K.L. analyzed snRNA-Seq data and visualized results. E.L. generated snRNA-Seq data, with support from R.S., M.T., and N.N. R.D.R., C.K.S., R.E.P.L., A.E., C.A.P. W.W.S., and S.S. contributed to neuropathological data generation and analysis, R.E., A.P. and H.H. contributed to neuropathological data analysis, and S.H.L. contributed to neuropathological method development.

Corresponding authors

Correspondence to Lea T. Grinberg or Martin Kampmann.

Ethics declarations

Ethics Declarations

This project was approved the ethical committee of the University of Sao Paulo (for tissue transfer) and deemed non-human subject research by UCSF.

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Asgeir Kobro-Flatmoen, Menno Witter, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Data quality and initial clustering without cross-sample alignment.

a-b, Mean number of genes (a) or UMIs (b) detected per cell across individual samples for major cell types identified in each dataset. Grubman et al.15 did not resolve excitatory neurons from inhibitory neurons. Pericytes were identified only in Mathys et al.14 Cell type abbreviations: Exc – excitatory neurons, Oligo – oligodendrocytes, Astro – astrocytes, Inh – inhibitory neurons, OPC – oligodendrocyte precursor cells, Micro – microglia, Endo – endothelial cells, Per – pericytes. c-d, tSNE projection of cells from the EC (c) and SFG (d) clustered without first performing cross-sample alignment, colored by individual of origin (center) or cluster assignment (outer). e-f, Heatmap and hierarchical clustering of clusters and cluster marker expression (top subpanels); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Expression of cell type markers (bottom subpanels).

Extended Data Fig. 2 Expression of selected EC excitatory neuron subpopulation markers and pathway enrichment analysis of differentially expressed genes in selectively vulnerable EC excitatory neuron subpopulations.

a, Expression heatmap of genes that are specifically expressed by four or fewer EC excitatory neuron subpopulations; “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). b-d, Enrichment analysis against Gene Ontology Cellular Component terms or Reactome Pathways (b,d) and functional association network analysis (c,e; see Methods) of genes with higher (b-c) or lower expression (d-e) in RORB+ vulnerable EC excitatory neurons, with selected terms highlighted by color. In panels c and e, genes with stronger associations are connected by thicker lines, and genes without known associations are not shown.

Extended Data Fig. 3 Differential expression analysis across Braak stages for EC excitatory neuron subpopulations.

a-b, Number of differentially expressed genes in EC excitatory neuron subpopulations with higher (a) or lower (b) expression in Braak stage 6 vs. Braak stage 0. c-f, Enrichment analysis against Gene Ontology Cellular Component terms (c-d) or Reactome Pathways (e-f) of differentially expressed genes in EC excitatory neuron subpopulations with higher (c,e) or lower (d,f) expression in Braak stage 6 vs. Braak stage 0.

Extended Data Fig. 4 Alignment of EC and SFG maps homologous excitatory neuron subpopulations.

a, tSNE projection of excitatory neurons from the EC and SFG in the joint alignment space, colored by subpopulation identity (top), individual of origin (middle), or brain region (bottom). b, Heatmap and hierarchical clustering of subpopulations and subpopulation marker expression (top subpanel); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Relative abundance of subpopulations across Braak stages (second and third subpanels); for each brain region, statistical significance of differences in relative abundance across Braak stages (Braak 0 n = 3, Braak 2 n = 4, Braak 6 n = 3, where n is the number of individuals sampled) was determined by beta regression and adjusted for multiple comparisons (see Methods). Expression heatmap of EC layer-specific genes identified from Ramsden et al.22 (fourth subpanel). Expression heatmap of neocortical layer-specific genes from Lake et al.12 (fifth subpanel). Expression of selectively vulnerable EC excitatory neuron subpopulation markers by excitatory neurons in the EC (sixth subpanel) or SFG (bottom subpanel). Significant beta regression P values (adjusted for multiple testing) are shown in a table at the bottom of the panel. c, Sankey diagram connecting subpopulation identity of excitatory neurons in the EC alignment space and the SFG alignment space to subpopulation identity in the EC+SFG alignment space. The links connecting EC:Exc.s2 and EC:Exc.s4 to SFG:Exc.s2 and SFG:Exc.s4, respectively, are highlighted.

Extended Data Fig. 5 Cross-sample alignment of excitatory neurons from Mathys et al. recapitulates selective vulnerability in a RORB-expressing subpopulation.

a, tSNE projection of excitatory neurons from Mathys et al.14 in the alignment space, colored by subpopulation identity (top) or individual of origin (bottom). b, Heatmap and hierarchical clustering of subpopulations and subpopulation marker expression (top subpanel); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Relative abundance of subpopulations in in AD cases vs. controls, separated by sex (second and third subpanels); for each sex, statistical significance of differences in relative abundance between AD cases vs. controls (cases n = 12, controls n=12, where n is the number of individuals sampled) was determined by beta regression and adjusted for multiple comparisons (see Methods). Expression heatmap of neocortical layer-specific genes from Lake et al.12 (fourth subpanel). Expression of selectively vulnerable EC excitatory neuron subpopulation markers (bottom subpanel). c, Heatmap of Pearson correlation between the gene expression profiles of excitatory neuron subpopulations from the EC vs. those from the prefrontal cortex in Mathys et al.14.

Extended Data Fig. 6 Delineation of the EC for each case used in immunofluorescence validation.

a, The borders of the caudal EC delineated on sections stained with hematoxylin and eosin (H&E) for all 26 cases used in immunofluorescence validation (Table 1). b, Borders of the EC were determined with the aid of 400 um thick serial coronal sections of whole-brain hemispheres stained with gallocyanin (see Methods). Each H&E section (left) along with its corresponding immunofluorescence image (middle) was aligned to the most approximate gallocyanin section (right), in which the the dissecans layers (diss-1, diss-2, and diss-ext) characteristic of the caudal EC were easier to visualize. This was then used to guide delineation of the EC on the H&E and immunofluorescence sections. For more details on the cytoarchitectonic definitions used to define the caudal EC, please consult Heinsen et al.19.

Extended Data Fig. 7 Inhibitory neurons from Mathys et al. also do not show differences in resilience or vulnerability to AD.

a, tSNE projection of inhibitory neurons from Mathys et al.14 in the alignment space, colored by subpopulation identity (top) or individual of origin (bottom). b, Heatmap and hierarchical clustering of subpopulations and subpopulation markers (top subpanel); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Relative abundance of subpopulations in in AD cases vs. controls, separated by sex (second and third subpanels); for each sex, statistical significance of differences in relative abundance between AD cases vs. controls (cases n=12, controls n=12, where n is the number of individuals sampled) was determined by beta regression and adjusted for multiple comparisons (see Methods). Expression heatmap of inhibitory neuron subtype markers from Lake et al.12 (bottom subpanel).

Extended Data Fig. 8 Subclustering of microglia does not sufficiently resolve disease associated microglia signature.

a-c, tSNE projection of astrocytes from the EC (a), SFG (b), and Mathys et al.14 (c) in their respective alignment spaces, colored by subpopulation identity (left) or individual of origin (right). d-f, Heatmap and hierarchical clustering of subpopulations and subpopulation marker expression (top subpanels); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Relative abundance of subpopulations (middle subpanels) across Braak stages in the EC and SFG (for each brain region, Braak 0 n=3, Braak 2 n=4, Braak 6 n=3, where n is the number of individuals sampled) or between AD cases vs. controls in Mathys et al.14 (for each sex, cases n =12, controls n = 12, where n is the number of individuals sampled); statistical significance of differences in relative abundance was determined by beta regression and adjusted for multiple comparisons (see Methods). Expression of disease associated microglia markers, with median expression level marked by line (bottom subpanels).

Extended Data Fig. 9 Subclustering of oligodendrocytes identifies subpopulations with higher expression of AD-associated oligodendrocyte markers from Mathys et al.

a-c, tSNE projection of oligodendrocytes from the EC (a), SFG (b), and Mathys et al.14 (c) in their respective alignment spaces, colored by subpopulation identity (left) or individual of origin (right). d-f, Heatmap and hierarchical clustering of subpopulations and subpopulation marker expression (top subpanels); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Relative abundance of subpopulations (middle subpanels) across Braak stages in the EC and SFG (for each brain region, Braak 0 n=3, Braak 2 n=4, Braak 6 n=3, where n is the number of individuals sampled) or between AD cases vs. controls in Mathys et al.14 (for each sex, cases n =12, controls n = 12, where n is the number of individuals sampled); statistical significance of differences in relative abundance was determined by beta regression and adjusted for multiple comparisons (see Methods). Relative expression of AD-associated oligodendrocyte subpopulation markers from Mathys et al.14 (bottom subpanels).

Extended Data Fig. 10 Astrocyte subpopulations with high GFAP expression from Mathys et al. are highly similar to those from the EC and SFG.

a, tSNE projection of astrocytes from Mathys et al.14 in the alignment subspace, colored by subpopulation identity (top) or individual of origin (bottom). b, Heatmap and hierarchical clustering of subpopulations and subpopulation marker expression (top subpanel); “High” and “Low” relative expression reflect above- and below-average expression, respectively (see Methods). Relative abundance of subpopulations in in AD cases vs. controls, separated by sex (middle subpanels); for each sex, statistical significance of differences in relative abundance between AD cases vs. controls (cases n=12, controls n=12, where n is the number of individuals sampled) was determined by beta regression and adjusted for multiple comparisons (see Methods). Expression of genes associated with reactive astrocytes, with median expression level marked by line (bottom subpanel). c, Enrichment analysis of overlap between differentially expressed genes in astrocytes with high GFAP expression from Mathys et al.14 vs. differentially expressed genes in astrocytes with high GFAP expression from the EC and SFG; the number of genes in each gene set and the number of overlapping genes are shown in parentheses, and the hypergeometric test p-values are shown without parentheses.

Supplementary information

Reporting Summary

Supplementary Tables

Supplementary Tables 1 and 2.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Leng, K., Li, E., Eser, R. et al. Molecular characterization of selectively vulnerable neurons in Alzheimer’s disease. Nat Neurosci 24, 276–287 (2021). https://doi.org/10.1038/s41593-020-00764-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-020-00764-7

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