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:

Epigenomic analysis of Parkinson’s disease neurons identifies Tet2 loss as neuroprotective

Abstract

Parkinson’s disease (PD) pathogenesis may involve the epigenetic control of enhancers that modify neuronal functions. Here, we comprehensively examine DNA methylation at enhancers, genome-wide, in neurons of patients with PD and of control individuals. We find a widespread increase in cytosine modifications at enhancers in PD neurons, which is partly explained by elevated hydroxymethylation levels. In particular, patients with PD exhibit an epigenetic and transcriptional upregulation of TET2, a master-regulator of cytosine modification status. TET2 depletion in a neuronal cell model results in cytosine modification changes that are reciprocal to those observed in PD neurons. Moreover, Tet2 inactivation in mice fully prevents nigral dopaminergic neuronal loss induced by previous inflammation. Tet2 loss also attenuates transcriptional immune responses to an inflammatory trigger. Thus, widespread epigenetic dysregulation of enhancers in PD neurons may, in part, be mediated by increased TET2 expression. Decreased Tet2 activity is neuroprotective, in vivo, and may be a new therapeutic target for PD.

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: Genome-wide hypermethylation at enhancers in PD prefrontal cortex neurons.
Fig. 2: Chromatin conformation analysis identifying gene targets of enhancers that are epigenetically dysregulated in PD neurons.
Fig. 3: Increased hydroxymethylation contributes to epigenetic abnormalities at enhancers in PD neurons.
Fig. 4: Transcriptomic changes in the PD prefrontal cortex converge on genes affecting neuronal signaling and immune activation.
Fig. 5: TET2 transcript levels are upregulated in PD and depletion of TET2 in vitro induces cytosine modification changes that are inverse to those observed in PD neurons.
Fig. 6: Tet2 inactivation protects against nigral dopaminergic neuronal loss induced by previous exposure to systemic inflammation.
Fig. 7: Tet2 loss suppresses immune signaling pathways induced by proinflammatory stimulus.

Similar content being viewed by others

Data availability

All sequencing data generated in this study are available from the NCBI Gene Expression Omnibus (GEO) database under the accession number GSE136010. DNA methylation padlock-seq and RNA-seq data used in this study are publicly available under the GEO accession number GSE135037. Hi-C data are available at GSM2322542. RNA-seq data from Parkinson’s disease models in Extended Data Fig. 7a–d are available under the GEO accession numbers GSE54795, GSE108370, GSE125239 and GSE130752. https://github.com/LeeLMarshall/Epigenomic-analysis-of-Parkinson-s-disease-neurons-identifies-Tet2-loss-as-neuroprotectiveSource data are provided with this paper.

Code availability

Code used for the analyses in this study is provided in the Supplementary Software and is available at https://github.com/LeeLMarshall/Epigenomic-analysis-of-Parkinson-s-disease-neurons-identifies-Tet2-loss-as-neuroprotective.

References

  1. GBD 2016 Parkinson’s Disease Collaborators. Global, regional, and national burden of Parkinson’s disease, 1990–2016: a systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 17, 939–953 (2018).

    Article  Google Scholar 

  2. Wirdefeldt, K., Gatz, M., Reynolds, C. A., Prescott, C. A. & Pedersen, N. L. Heritability of Parkinson disease in Swedish twins: a longitudinal study. Neurobiol. Aging 32, 1923.e1–1923.e8 (2011).

    Article  Google Scholar 

  3. Gomez-Esteban, J. C. et al. Factors influencing the symmetry of Parkinson’s disease symptoms. Clin. Neurol. Neurosurg. 112, 302–305 (2010).

    Article  PubMed  Google Scholar 

  4. Baldereschi, M. et al. Parkinson’s disease and parkinsonism in a longitudinal study: two-fold higher incidence in men. ILSA Working Group. Italian Longitudinal Study on Aging. Neurology 55, 1358–1363 (2000).

    Article  CAS  PubMed  Google Scholar 

  5. Labbe, C., Lorenzo-Betancor, O. & Ross, O. A. Epigenetic regulation in Parkinson’s disease. Acta Neuropathol. 132, 515–530 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Jakubowski, J. L. & Labrie, V. Epigenetic biomarkers for Parkinson’s disease: from diagnostics to therapeutics. J. Parkinsons Dis. 7, 1–12 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Masliah, E., Dumaop, W., Galasko, D. & Desplats, P. Distinctive patterns of DNA methylation associated with Parkinson disease: identification of concordant epigenetic changes in brain and peripheral blood leukocytes. Epigenetics 8, 1030–1038 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Kaut, O., Schmitt, I. & Wullner, U. Genome-scale methylation analysis of Parkinson’s disease patients’ brains reveals DNA hypomethylation and increased mRNA expression of cytochrome P450 2E1. Neurogenetics 13, 87–91 (2012).

    Article  CAS  PubMed  Google Scholar 

  9. Young, J. I. et al. Genome-wide brain DNA methylation analysis suggests epigenetic reprogramming in Parkinson disease. Neurol. Genet. 5, e342 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Braak, H. et al. Staging of brain pathology related to sporadic Parkinson’s disease. Neurobiol. Aging 24, 197–211 (2003).

    Article  PubMed  Google Scholar 

  11. Guo, J. U. et al. Neuronal activity modifies the DNA methylation landscape in the adult brain. Nat. Neurosci. 14, 1345–1351 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Feng, J. et al. Dnmt1 and Dnmt3a maintain DNA methylation and regulate synaptic function in adult forebrain neurons. Nat. Neurosci. 13, 423–430 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Li, X. et al. Ten-eleven translocation 2 interacts with forkhead box O3 and regulates adult neurogenesis. Nat. Commun. 8, 15903 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Hon, G. C. et al. 5mC oxidation by Tet2 modulates enhancer activity and timing of transcriptome reprogramming during differentiation. Mol. Cell 56, 286–297 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Lister, R. et al. Global epigenomic reconfiguration during mammalian brain development. Science 341, 1237905 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Price, A. J. et al. Divergent neuronal DNA methylation patterns across human cortical development reveal critical periods and a unique role of CpH methylation. Genome Biol. 20, 196 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Roadmap Epigenomics Consortium et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015).

  18. Dao, L. T. M. et al. Genome-wide characterization of mammalian promoters with distal enhancer functions. Nat. Genet. 49, 1073–1081 (2017).

    Article  CAS  PubMed  Google Scholar 

  19. Dong, X. et al. Enhancers active in dopamine neurons are a primary link between genetic variation and neuropsychiatric disease. Nat. Neurosci. 21, 1482–1492 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Soldner, F. et al. Parkinson-associated risk variant in distal enhancer of ɑ-synuclein modulates target gene expression. Nature 533, 95–99 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Fernandez-Santiago, R. et al. Aberrant epigenome in iPSC-derived dopaminergic neurons from Parkinson’s disease patients. EMBO Mol. Med. 7, 1529–1546 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Li, P. et al. Hemispheric asymmetry in the human brain and in Parkinson’s disease is linked to divergent epigenetic patterns in neurons. Genome Biol. 21, 61 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Narayanan, N. S., Rodnitzky, R. L. & Uc, E. Y. Prefrontal dopamine signaling and cognitive symptoms of Parkinson’s disease. Rev. Neurosci. 24, 267–278 (2013).

    Article  PubMed  Google Scholar 

  24. Luo, C. et al. Single-cell methylomes identify neuronal subtypes and regulatory elements in mammalian cortex. Science 357, 600–604 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Farley, J. E. et al. Transcription factor Pebbled/RREB1 regulates injury-induced axon degeneration. Proc. Natl Acad. Sci. USA 115, 1358–1363 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Schmitt, A. D. et al. A compendium of chromatin contact maps reveals spatially active regions in the human genome. Cell Rep. 17, 2042–2059 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Lio, C. J. & Rao, A. TET enzymes and 5hmC in adaptive and innate immune systems. Front. Immunol. 10, 210 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Nalls, M. A. et al. Identification of novel risk loci, causal insights, and heritable risk for Parkinson’s disease: a meta-analysis of genome-wide association studies. Lancet Neurol. 18, 1091–1102 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Figge, D. A., Eskow Jaunarajs, K. L. & Standaert, D. G. Dynamic DNA methylation regulates levodopa-induced dyskinesia. J. Neurosci. 36, 6514–6524 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Risso, D., Ngai, J., Speed, T. P. & Dudoit, S. Normalization of RNA-seq data using factor analysis of control genes or samples. Nat. Biotechnol. 32, 896–902 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Donega, V. et al. Transcriptome and proteome profiling of neural stem cells from the human subventricular zone in Parkinson’s disease. Acta Neuropathol. Commun. 7, 84 (2019).

    Article  PubMed  Google Scholar 

  32. Johnson, M. E., Stecher, B., Labrie, V., Brundin, L. & Brundin, P. Triggers, facilitators, and aggravators: redefining Parkinson’s disease pathogenesis. Trends Neurosci. 42, 4–13 (2019).

    Article  CAS  PubMed  Google Scholar 

  33. Qin, L. et al. Systemic LPS causes chronic neuroinflammation and progressive neurodegeneration. Glia 55, 453–462 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Brichta, L. et al. Identification of neurodegenerative factors using translatome-regulatory network analysis. Nat. Neurosci. 18, 1325–1333 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Sodersten, E. et al. A comprehensive map coupling histone modifications with gene regulation in adult dopaminergic and serotonergic neurons. Nat. Commun. 9, 1226 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Wu, T. T. et al. TET2-mediated Cdkn2A DNA hydroxymethylation in midbrain dopaminergic neuron injury of Parkinson’s disease. Hum. Mol. Genet. 29, 1239–1252 (2020).

    Article  CAS  PubMed  Google Scholar 

  37. Kriaucionis, S. & Heintz, N. The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 324, 929–930 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kozlenkov, A. et al. A unique role for DNA (hydroxy)methylation in epigenetic regulation of human inhibitory neurons. Sci. Adv. 4, eaau6190 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  39. Szulwach, K. E. et al. 5-hmC-mediated epigenetic dynamics during postnatal neurodevelopment and aging. Nat. Neurosci. 14, 1607–1616 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Herrup, K. & Yang, Y. Cell cycle regulation in the postmitotic neuron: oxymoron or new biology? Nat. Rev. Neurosci. 8, 368–378 (2007).

    Article  CAS  PubMed  Google Scholar 

  41. Ellison, E. M., Abner, E. L. & Lovell, M. A. Multiregional analysis of global 5-methylcytosine and 5-hydroxymethylcytosine throughout the progression of Alzheimer’s disease. J. Neurochem. 140, 383–394 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Stoger, R., Scaife, P. J., Shephard, F. & Chakrabarti, L. Elevated 5hmC levels characterize DNA of the cerebellum in Parkinson’s disease. NPJ Parkinsons Dis. 3, 6 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Carrillo-Jimenez, A. et al. TET2 regulates the neuroinflammatory response in microglia. Cell Rep. 29, e698 (2019).

    Article  CAS  Google Scholar 

  44. Jain, N. et al. Global modulation in DNA epigenetics during pro-inflammatory macrophage activation. Epigenetics 14, 1183–1193 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Zhang, Q. et al. Tet2 is required to resolve inflammation by recruiting Hdac2 to specifically repress IL-6. Nature 525, 389–393 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Pronier, E. et al. Inhibition of TET2-mediated conversion of 5-methylcytosine to 5-hydroxymethylcytosine disturbs erythroid and granulomonocytic differentiation of human hematopoietic progenitors. Blood 118, 2551–2555 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Ichiyama, K. et al. The methylcytosine dioxygenase Tet2 promotes DNA demethylation and activation of cytokine gene expression in T cells. Immunity 42, 613–626 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Izzo, F. et al. DNA methylation disruption reshapes the hematopoietic differentiation landscape. Nat. Genet. 52, 378–387 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Gagne, J. J. & Power, M. C. Anti-inflammatory drugs and risk of Parkinson disease: a meta-analysis. Neurology 74, 995–1002 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Peter, I. et al. Anti-tumor necrosis factor therapy and incidence of Parkinson disease among patients with inflammatory bowel disease. JAMA Neurol. 75, 939–946 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  51. Caligiore, D. et al. Parkinson’s disease as a system-level disorder. NPJ Parkinsons Dis. 2, 16025 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Weintraub, D. et al. Neurodegeneration across stages of cognitive decline in Parkinson disease. Arch. Neurol. 68, 1562–1568 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Kordower, J. H. et al. Disease duration and the integrity of the nigrostriatal system in Parkinson’s disease. Brain 136, 2419–2431 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Li, P. et al. Epigenetic dysregulation of enhancers in neurons is associated with Alzheimer’s disease pathology and cognitive symptoms. Nat. Commun. 10, 2246 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Pai, S. et al. Differential methylation of enhancer at IGF2 is associated with abnormal dopamine synthesis in major psychosis. Nat. Commun. 10, 2046 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  56. He, Y. & Wang, T. EpiCompare: an online tool to define and explore genomic regions with tissue or cell type-specific epigenomic features. Bioinformatics 33, 3268–3275 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Ernst, J. & Kellis, M. Chromatin-state discovery and genome annotation with ChromHMM. Nat. Protoc. 12, 2478–2492 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Labrie, V. et al. Lactase nonpersistence is directed by DNA-variation-dependent epigenetic aging. Nat. Struct. Mol. Biol. 23, 566–573 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Diep, D. et al. Library-free methylation sequencing with bisulfite padlock probes. Nat. Methods 9, 270–272 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Krueger, F. & Andrews, S. R. Bismark: a flexible aligner and methylation caller for Bisulfite-seq applications. Bioinformatics 27, 1571–1572 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. The 1000 Genomes Project Consortium et al. A global reference for human genetic variation. Nature 526, 68–74 (2015).

    Article  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chen, B., Khodadoust, M. S., Liu, C. L., Newman, A. M. & Alizadeh, A. A. Profiling tumor infiltrating immune cells with CIBERSORT. Methods Mol. Biol. 1711, 243–259 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Akalin, A. et al. methylKit: a comprehensive R package for the analysis of genome-wide DNA methylation profiles. Genome Biol. 13, R87 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Kwon, A. T., Arenillas, D. J., Worsley Hunt, R. & Wasserman, W. W. oPOSSUM-3: advanced analysis of regulatory motif over-representation across genes or ChIP-seq datasets. G3 (Bethesda) 2, 987–1002 (2012).

    Article  CAS  Google Scholar 

  66. Rao, S. S. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Mishra, A. & Hawkins, R. D. Three-dimensional genome architecture and emerging technologies: looping in disease. Genome Med. 9, 87 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Wingett, S. et al. HiCUP: pipeline for mapping and processing Hi-C data. F1000Res. 4, 1310 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Forcato, M. et al. Comparison of computational methods for Hi-C data analysis. Nat. Methods 14, 679–685 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Levy-Leduc, C., Delattre, M., Mary-Huard, T. & Robin, S. Two-dimensional segmentation for analyzing Hi-C data. Bioinformatics 30, i386–i392 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. McLean, C. Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nat. Biotechnol. 28, 495–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Reimand, J., Kull, M., Peterson, H., Hansen, J. & Vilo, J. g:Profiler—a web-based toolset for functional profiling of gene lists from large-scale experiments. Nucleic Acids Res. 35, W193–W200 (2007).

    Article  PubMed  PubMed Central  Google Scholar 

  73. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Reimand, J. et al. Pathway enrichment analysis and visualization of omics data using g:Profiler, GSEA, Cytoscape and EnrichmentMap. Nat. Protoc. 14, 482–517 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Ramirez, F., Dundar, F., Diehl, S., Gruning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Diaz, A., Park, K., Lim, D. A. & Song, J. S. Normalization, bias correction, and peak calling for ChIP-seq. Stat. Appl. Genet. Mol. Biol. 11, 9 (2012).

    Article  CAS  Google Scholar 

  78. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  80. 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  CAS  PubMed  Google Scholar 

  81. Yu, Q. & He, Z. Comprehensive investigation of temporal and autism-associated cell type composition-dependent and independent gene expression changes in human brains. Sci. Rep. 7, 4121 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  82. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  83. Konnova, E. A. & Swanberg, M. Animal models of Parkinson’s disease. in Parkinson’s Disease: Pathogenesis and Clinical Aspects (eds Stoker, T. B. & Greenland, J. C.) (Codon Publications, 2018).

  84. Caiazzo, M. et al. Direct generation of functional dopaminergic neurons from mouse and human fibroblasts. Nature 476, 224–227 (2011).

    Article  CAS  PubMed  Google Scholar 

  85. Cheng, L. et al. Gene dysregulation is restored in the Parkinson’s disease MPTP neurotoxic mice model upon treatment of the therapeutic drug Cu(II)(atsm). Sci. Rep. 6, 22398 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Chen, X. et al. Parkinson’s disease-linked D620N VPS35 knockin mice manifest tau neuropathology and dopaminergic neurodegeneration. Proc. Natl Acad. Sci. USA 116, 5765–5774 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Maco, B. et al. Semiautomated correlative 3D electron microscopy of in vivo-imaged axons and dendrites. Nat. Protoc. 9, 1354–1366 (2014).

    Article  CAS  PubMed  Google Scholar 

  88. McQuin, C. et al. CellProfiler 3.0: next-generation image processing for biology. PLoS Biol. 16, e2005970 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We thank the Van Andel Institute Flow Cytometry, Genomics, Bioinformatics and Biostatistics, Pathology and Vivarium Cores. We also thank the CAMH Sequencing Facility. We thank the Parkinson’s UK Brain Bank, the NIH NeuroBioBank and the Michigan Brain Bank for the brain tissue provided. This work was funded by a Department of Defense grant (grant no. W81XWH1810512) and a VAI Innovation Award to V.L. V.L. is also supported by grants from the National Institute of Neurological Disorders and Stroke (grant no. 1R21NS112614-01, grant no. 1R01NS114409-01A1 and grant no. 1R01NS113894-01A1), the Farmer Family Foundation Parkinson’s Research Initiative and a Gibby & Friends vs. Parky Award. S.J. and M.W. are supported by the Richard and Helen DeVos Foundation.

Author information

Authors and Affiliations

Authors

Contributions

L.L.M. contributed to experimental design, computational analyses, immunohistochemistry experiments and cell culture study. B.A.K. performed neuronal nuclei isolations and the bisulfite padlock probe sequencing of human patient samples, and was involved in the immunohistochemistry and behavioral analyses in mice. E.E. performed the TET2 mRNA analysis in patient samples and RNA isolation of mouse samples, and contributed to the immunohistochemistry experiments. P.L. was involved in the Hi-C analysis and experimental design of computational approaches. W.C. performed the bisulfite padlock probe sequencing of the cell culture study. K.X.L. contributed to the immunohistochemistry analysis. M.W. and S.J. were involved in the flow sorting of neuronal nuclei. J.G. contributed to the hMEDIP analysis. G.A.C. contributed to experimental design. J.M. and X.W. contributed to the analysis of mouse models. V.L. was involved with study design and overseeing the experiments. The manuscript was written by V.L. and L.L.M. and commented on by all authors.

Corresponding author

Correspondence to Viviane Labrie.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Schahram Akbarian and Tiago Outeiro 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 Analysis of sample distribution and genomic context of modified cytosine sites at enhancers.

a, Dendrogram showing unsupervised clustering of samples based on cytosine sites. DNA methylation at top 10,000 most variable cytosine sites. Clustering of corresponding technical (nuclei sorting and/or library preparation replicated for select DNA samples) and sequencing replicates (sequencing replicated across lanes/flow cells for select libraries) are shown in pink and purple, respectively. Pearson’s correlation of technical replicates is 0.972 ± 0.0012 s.e.m. and of sequencing replicates is 0.977 ± 0.0010 s.e.m., supporting a high technical reproducibility. b, The proportion of each CpH subtype for the differentially methylated cytosine sites in PD neurons (significant: orange) as compared to the total number of cytosine sites profiled (background: gray). CpH context for cytosine sites in enhancers in this study was similar to that of previous whole genome studies in neurons24. c, The proportion of differentially modified cytosine sites in PD located within an enhancer or promoter. Differentially modified cytosines in neurons of PD patients relative to controls were identified by logistic regression after adjusting for age, sex, postmortem interval, and neuronal subtypes (n = 57 PD, 48 controls; q < 0.05, multiple testing corrected). d, The proportion of 21 specific neuronal subtypes present in the prefrontal cortex of PD patients (orange) and controls (gray). There are 12 glutamatergic and 9 GABAergic neuronal types in the NeuN+ nuclei from PD and controls (n = 57 and 48 individuals, respectively). Neuronal proportions were calculated by computational deconvolution, using neuronal subtype specific markers (n = 563 cytosine sites)24. No significant differences were determined between PD and controls, as determined by a repeated-measures ANOVA and post hoc Tukey HSD test. Error bars represent s.e.m. e, Proportion of glutamatergic neurons and GABAergic neurons in PD and controls. Neuronal proportions determined by the sum of all glutamatergic or GABAergic neuronal subtypes identified by computational deconvolution in panel a. Averaged 76.95% glutamate neuronal subtype proportion ± 0.814 s.e.m. There were no significant differences in glutamatergic and GABAergic neuronal proportions between PD patients and controls, as determined by a repeated-measures ANOVA and post hoc Tukey HSD test. Boxplot center line is the mean, the bounds of the box are the lower and upper quartiles, and the whiskers extend to minimum and maximum data points within 1.5 times the interquartile range.

Extended Data Fig. 2 Genome-wide hypermethylation of differentially methylated enhancer regions in PD neurons.

Differentially methylated regions at enhancers in prefrontal cortex neurons of 57 PD patients and 48 controls were identified (methylKit, sliding window: 1000 bp length, 50 bp step increment). a, Manhattan plot showing differentially methylated regions at enhancers in PD neurons (q < 0.05, logistic regression, adjusting for age, sex, postmortem interval, and neuronal subtypes). -log10(PValue) refers to the significance of region methylation change in PD, with the sign corresponding to the direction of change (red: hypermethylated, blue: hypomethylated region in PD). Threshold for genome-wide significance is q < 0.05 (black line, multiple testing corrected). Differentially methylated region in an enhancer of TET2 in PD neurons is highlighted. b, Proportion of hypermethylated (red) and hypomethylated (blue) regions at enhancers in PD. Enrichment determined by Fisher’s exact test with OR referring to the odds ratio and P = 10-38 significance of hypermethylation enrichment. Gray is all regions in the analysis (background). c, Intersection of the genes with differentially methylated cytosines or regions at enhancers in PD neurons. P = 10−16 is the significance of overlap determined by Fisher’s exact test.

Extended Data Fig. 3 Epigenetic changes in the prefrontal cortex precede the arrival of PD pathology.

Differentially methylated cytosines in neurons of the PD prefrontal cortex before the arrival of Lewy pathology (PD Braak stage 3–4, n = 22 individuals, green), after the arrival of Lewy pathology (PD Braak stage 5–6, n = 34 individuals, yellow) and in the combined PD group (n = 57 individuals, red), in comparison to healthy controls (n = 48 individuals). a, b, Manhattan plot showing differentially methylated cytosines in PD neurons before (a) and after (b) the arrival of Lewy pathology. -log10(PValue) refers to the significance of cytosine methylation change in PD, with the sign corresponding to the direction of change (red: hypermethylated, blue: hypomethylated cytosine in PD). Threshold for genome-wide significance is q < 0.05 (black line, multiple testing corrected). c, Density plot of differentially methylated cytosines in PD prefrontal cortex neurons before and after arrival of Lewy pathology. Enrichment determined by Fisher’s exact test with OR referring to the odds ratio and p-value is significance of hypermethylation enrichment. d, Intersection of differentially methylated cytosines in PD prefrontal cortex neurons in the combined PD group and before or after arrival of Lewy pathology.

Extended Data Fig. 4 Independent validation of DNA methylation changes at enhancers, genome-wide, in PD neurons.

Differential methylation at enhancers in PD prefrontal cortex neurons identified in the discovery cohort were validated with an independent replication cohort of prefrontal cortex neurons from 27 PD patients and 31 controls. a, Manhattan plot showing differentially methylated cytosines at enhancers in PD neurons, identified by logistic regression after adjusting for age, sex, postmortem interval, and neuronal subtypes. -log10(PValue) refers to the significance of cytosine methylation change in PD, with the sign corresponding to the direction of change (red: hypermethylated, blue: hypomethylated cytosine in PD). Threshold for genome-wide significance is q < 0.05 (black line, multiple testing corrected). b, Proportion of hypermethylated (red) and hypomethylated (blue) cytosines at enhancers in PD replication cohort. Enrichment determined by Fisher’s exact test with OR referring to the odds ratio and P = 10−6 is the significance of hypermethylation enrichment. Gray is all cytosines in the analysis (background). c, Correlation between the discovery and replication cohort in the DNA methylation fold changes at enhancer cytosine sites in PD neurons. P = 10−111 is the significance of Pearson’s correlation by two-tailed t-test. d, Intersection of the genes with differentially methylated cytosines at their enhancers in the discovery and replication cohort. P = 10−79 is the significance of overlap determined by Fisher’s exact test.

Extended Data Fig. 5 Medications are not a major contributor to the PD-associated DNA methylation changes identified in this study.

a, Heatmap correlations and unsupervised clustering of patient covariates (yellow: positively correlated, blue: negatively correlated; Pearson’s correlation). b, Venn diagram showing the overlap of differentially methylated cytosines at enhancers in PD patients (n = 57 PD, 48 controls) with differentially methylated cytosines related to PD patients taking dopamine agonists, COMT inhibitors, and/or MAO-B inhibitors. Differentially methylated cytosines related to PD medications were identified by logistic regression, after adjusting for age, sex, postmortem interval, and neuronal subtypes (q < 0.05, multiple testing corrected). c, Intersection of genes with differential methylation in PD and with differential methylation in response to levodopa (L-Dopa) in the brain of a rat model of PD (dorsal striatum of 6-OHDA–treated rats receiving L-Dopa)9,29. Enrichment determined by Fisher’s exact test.

Extended Data Fig. 6 The prefrontal cortex of PD patients does not exhibit differences in the proportion of brain cell types.

Cell-type deconvolution in 24 PD and 12 control prefrontal cortex transcriptomes was performed using 834 cell-type-specific gene signatures63,81. a, Cell-type proportions in the prefrontal cortex of PD patients (orange) and controls (gray). There were no significant differences between PD patients and controls in brain cell types, as determined by a repeated-measures ANOVA and post hoc Tukey HSD test. Boxplot center line is the mean, the bounds of the box are the lower and upper quartiles, and the whiskers extend to minimum and maximum data points within 1.5 times the interquartile range. b, Correlation of neuronal proportions in transcriptomic data with the correction for sources of unknown variation. Statistical analysis of transcriptomic changes in the prefrontal cortex of PD patients was corrected for unknown sources variation, as determined by RUVseq30. P = 10−9 is the significance of Pearson’s correlation by two-tailed t-test, with 95% confidence intervals shown in gray. c, Comparison of statistical models that identifies differentially expressed genes in the PD prefrontal cortex after controlling for unknown sources of variation or neuronal proportion. Differentially expressed genes in the prefrontal cortex of PD patients were identified using a generalized linear model, adjusting for age, sex, RIN, and neuronal proportion or sources of unknown variation (q < 0.05, absolute logFC≥0.5, multiple testing corrected). The outcome of both models is comparable, showing significant overlap of differentially expressed genes in PD (P = 10−228, Fisher’s exact test). d, TET1 and TET3 transcript levels in the prefrontal cortex of PD patients relative to controls. Differentially expressed genes in the prefrontal cortex of PD patients were identified using a generalized linear model, adjusting for age, sex, RIN, and sources of unknown variation. There are no significant changes in TET1 and TET3 transcript levels between PD patients and controls (nominal P = 0.96 and 0.09, respectively). Boxplot center line is the mean, the bounds of the box are the lower and upper quartiles, and the whiskers extend to minimum and maximum data points within 1.5 times the interquartile range. a‒d, n = 12 controls, 24 PD. e, Increased TET2 transcript levels are correlated with reduced TH protein levels in the prefrontal cortex. n = 11 individuals. P = 0.03 is the significance of inverse association between TET2 mRNA and TH protein levels by linear regression. The 95% confidence intervals are shown in gray.

Source data

Extended Data Fig. 7 TET2 expression is increased in neurons of PD-relevant models and PD patient neurons.

a, TET2 transcript levels in induced dopaminergic (iDA) neurons of PD patients and controls (n = 8 control fibroblasts, 7 PD fibroblasts, 8 control iDA neurons, 9 PD iDA neurons). b, TET2 transcript levels in bulk subventricular zone (SVZ), isolated neuronal stem cells (NSCs), and isolated microglia cells from PD patients and controls (n = 3 control SVZ, 4 PD SVZ, 4 control NSCs, 5 PD NSCs, 2 control microglia, 4 PD microglia). c, d, Tet2 transcript levels in neurons of PD-relevant models. Tet2 mRNA levels in dopaminergic neurons of mice treated with (c) MPTP (n = 4 saline, 4 MPTP) or (d) 6-OHDA (n = 4 non-injected side, 4 6-OHDA). ad, Transcriptomic analyses of RNA-seq datasets were performed using a generalized linear model adjusting for sources of unknown variation, with contrast fit for group-wise comparisons. TET2 transcript levels in counts per million (cpm) with nominal P < 0.05 significance. Datasets: GSE54795, GSE108370, GSE125239, and GSE130752. e, TET2 mRNA depletion by siRNA in the neuronal cell model, SH-SY5Y cells. TET2 mRNA transcript levels in SH-SY5Y cells 72 h after incubation with TET2 siRNA (TET2i) or non-target control (NT). TET2 mRNA levels determined by qPCR and normalized to housekeeping genes B2M, HPRT, and RPL13 (n = 6 NT, 6 TET2i samples). ****P = 10−5, by two-tailed, unpaired equal variance t-test. ae, Boxplot center line is the mean, the bounds of the box are the lower and upper quartiles, and the whiskers extend to minimum and maximum data points within 1.5 times the interquartile range. f, Cytosine modification changes in response to TET2 mRNA depletion in vitro. Differentially modified cytosines induced by TET2 reduction in SH-SY5Y cells were identified using logistic regression (n = 8 NT, 8 TET2i). Enrichment in hypomethylation in response to TET2 inactivation determined by Fisher’s exact test with odds ratio (OR) and P = 10−234 for significance of enrichment. The proportion of significantly hypomethylated (blue) and hypermethylated (red) cytosines induced by TET2 inactivation is shown, with all cytosines in the analysis (background) in gray.

Extended Data Fig. 8 Immunohistochemistry analysis of microglia activation and proteinase-K resistant α-synuclein aggregates in mice with Tet2 inactivation 11 months after exposure to inflammation.

Immunohistochemistry image analysis of microglial activation in the (a) dorsal striatum and the (b) substantia nigra. Representative images of dorsal striatum and substantia nigra microglia (IBA1-positive). n = 9 wild-type + saline, 10 wild-type + LPS, 14 Tet2 knockout + saline, and 10 Tet2 knockout + LPS. Immunohistochemistry image analysis of proteinase-K resistant α-synuclein aggregates in the (c) dorsal striatum and the (d) substantia nigra. Data analyzed by two-way ANOVA followed by post hoc Tukey HSD test. n = 9 wild-type + saline, 9 wild-type + LPS, 14 Tet2 knockout + saline, and 7 Tet2 knockout + LPS. Error bars represent s.e.m. e, Transcript analysis of inflammatory marker genes in the midbrain of wild-type and Tet2 knockout mice 11 months post-LPS or saline injection. Transcript levels of IL-1β, IL-6, NFKB2, RELA and TNF-α were analyzed by qPCR and normalized to housekeeping genes HPRT and GAPDH. n = 6 wild-type + saline, 8 wild-type + LPS, 5 Tet2 knockout + saline, 10 Tet2 knockout + LPS. f, Tet2 inactivation does not induce differences in midbrain Tet1 or Tet3 transcript levels. Tet1 and Tet3 transcript levels in the midbrain of wild-type (Tet2+/+) and Tet2 knockout (Tet2-/-) mice 24 h after saline or LPS treatment (n = 4 wild-type + saline, 4 wild-type + LPS, 3 Tet2 knockout + saline, 3 Tet2 knockout + LPS). There was no significant change in Tet1 or Tet3 expression between mouse groups, as determined by a generalized linear model. Transcript levels in counts per million (cpm) from RNA-seq data. e, f, Boxplot center line is the mean, the bounds of the box are the lower and upper quartiles, and the whiskers extend to minimum and maximum data points within 1.5 times the interquartile range.

Extended Data Fig. 9 Tet2 inactivation protects against LPS induced motor learning deficits. Behavioral analysis in wild-type and Tet2 knockout mice 8 and 10 months after exposure to LPS or saline.

a, Rotarod learning assessed in wild-type and Tet2 knockout mice 8 months post-LPS or saline injection (*P = 0.04, **P < 0.01, repeated measures ANOVA). b, Rotarod analysis of wild-type and Tet2 knockout mice at 10 months post-LPS or saline injection. Open field analysis of wild-type and Tet2 knockout mice at (c) 8 months and (d) 10 months post-LPS or saline injection. Gait analysis using the CatWalk system for wild-type and Tet2 knockout mice measuring (e, f) average speed or (g, h) cadence at 8 months and 10 months post-LPS or saline injection. ah, All statistical analyses were performed using a two-way or repeated measures ANOVA followed by post hoc Tukey HSD test. 8 month: n = 11 wild-type + saline, 7 wild-type + LPS, 13–14 Tet2 knockout + saline, 11 Tet2 knockout + LPS; 10 month: n = 11 wild-type + saline, 7 wild-type + LPS, 11–12 Tet2 knockout + saline, 8 Tet2 knockout + LPS. Error bars representing s.e.m.

Extended Data Fig. 10 Proposed model displaying the contribution of TET2 to the development and progression of PD.

TET2 is responsible for the conversion of DNA methylation to hydroxymethylation, and TET2 inactivation results in cytosine modification changes relevant to those observed in PD neurons. In PD, there is epigenetic dysregulation of the TET2 locus and an upregulation of TET2 expression, which leads to increased hydroxymethylation. Based on our study and the existing literature13,27,43, this could have two main consequences. First, elevated TET2 levels increases susceptibility to immune activation, promoting neuroinflammation. Second, TET2 upregulation activates genes involved in neurogenesis in post-mitotic neurons. Both of these aberrant processes may lead to the neurodegeneration observed in PD patients.

Supplementary information

Supplementary Information

Supplementary Fig. 1.

Reporting Summary

Supplementary Software

Analysis code

Supplementary Table 1

Demographic and clinical information for study samples

Supplementary Table 2

Differentially methylated cytosine sites at enhancers in PD neurons

Supplementary Table 3

Differentially methylated regions at enhancers in PD neurons

Supplementary Table 4

Differentially methylated sites in PD neurons before and after the arrival of α-synuclein pathology in the prefrontal cortex

Supplementary Table 5

Differentially methylated cytosine sites at enhancers in PD neurons in replication cohort

Supplementary Table 6

Gene targets of differentially methylated enhancers in PD determined by chromatin interaction analysis in prefrontal cortex

Supplementary Table 7

Differentially methylated sites of PD medications at enhancers in PD neurons

Supplementary Table 8

Hydroxymethylation changes, genome-wide, in PD neurons determined by hMeDIP-seq

Supplementary Table 9

Hydroxymethylation changes at enhancers in PD neurons

Supplementary Table 10

Overlap of hydroxymethylation changes with differentially modified enhancers in PD

Supplementary Table 11

Differentially expressed genes in the prefrontal cortex of PD patients determined by RNA-seq

Supplementary Table 12

Pathway analysis integrating DNA methylation and transcriptomic alterations in PD brain

Supplementary Table 13

Epigenetically and transcriptionally dysregulated genes in PD

Supplementary Table 14

Differentially modified cytosines at the TET2 gene locus and surrounding genomic area

Supplementary Table 15

Differentially modified cytosines induced by TET2 knockdown in SH-SY5Y cells

Supplementary Table 16

Differentially modified cytosines induced by TET2 knockdown in vitro that overlap enhancers dysregulated in PD neurons

Supplementary Table 17

Transcriptomic alterations in wild-type and Tet2 knockout mice given saline or LPS

Supplementary Table 18

Pathways altered in wild-type and Tet2 knockout mice given saline or LPS

Supplementary Table 19

Primers

Supplementary Table 20

Antibodies

Supplementary Table 21

Computational tools used for study analyses

Source data

Source Data 1

Source data for Extended Data Fig. 6. Correlation of TET2 expression and TH protein levels.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Marshall, L.L., Killinger, B.A., Ensink, E. et al. Epigenomic analysis of Parkinson’s disease neurons identifies Tet2 loss as neuroprotective. Nat Neurosci 23, 1203–1214 (2020). https://doi.org/10.1038/s41593-020-0690-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-020-0690-y

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