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:

TET2 modulates spatial relocalization of heterochromatin in aged hematopoietic stem and progenitor cells

Abstract

DNA methylation deregulation at partially methylated domains (PMDs) represents an epigenetic signature of aging and cancer, yet the underlying molecular basis and resulting biological consequences remain unresolved. We report herein a mechanistic link between disrupted DNA methylation at PMDs and the spatial relocalization of H3K9me3-marked heterochromatin in aged hematopoietic stem and progenitor cells (HSPCs) or those with impaired DNA methylation. We uncover that TET2 modulates the spatial redistribution of H3K9me3-marked heterochromatin to mediate the upregulation of endogenous retroviruses (ERVs) and interferon-stimulated genes (ISGs), hence contributing to functional decline of aged HSPCs. TET2-deficient HSPCs retain perinuclear distribution of heterochromatin and exhibit age-related clonal expansion. Reverse transcriptase inhibitors suppress ERVs and ISGs expression, thereby restoring age-related defects in aged HSPCs. Collectively, our findings deepen the understanding of the functional interplay between DNA methylation and histone modifications, which is vital for maintaining heterochromatin function and safeguarding genome stability in stem cells.

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: Tet2 depletion rescues age-related decline in HSPC self-renewal and repopulation capability.
Fig. 2: Distinct transcriptional alterations during WT and Tet2KO HSPCs aging.
Fig. 3: Tet2 loss prevents the spatial relocalization of H3K9me3-marked genomic regions from the nuclear periphery to the nucleoplasm during HSPC aging.
Fig. 4: DNA methylation dysregulation contributes to the spatial relocalization of H3K9me3 in HSPCs.
Fig. 5: Upregulation of repetitive elements in aged HSPCs contributes to their functional decline.

Similar content being viewed by others

Data availability

The sequencing datasets were deposited into the NCBI BioProject under the accession number GSE183675. The majority of data are available in the source data file. All other data supporting the findings of this study are available from the corresponding authors upon reasonable request.

References

  1. Kramer, A. & Challen, G. A. The epigenetic basis of hematopoietic stem cell aging. Semin. Hematol. 54, 19–24 (2017).

    Article  PubMed  Google Scholar 

  2. Sun, D. et al. Epigenomic profiling of young and aged HSCs reveals concerted changes during aging that reinforce self-renewal. Cell Stem Cell 14, 673–688 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Field, A. E. et al. DNA methylation clocks in aging: categories, causes, and consequences. Mol. Cell 71, 882–895 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Zhou, W. et al. DNA methylation loss in late-replicating domains is linked to mitotic cell division. Nat. Genet. 50, 591–602 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Hernando-Herraez, I. et al. Ageing affects DNA methylation drift and transcriptional cell-to-cell variability in mouse muscle stem cells. Nat. Commun. 10, 4361 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  6. Salhab, A. et al. A comprehensive analysis of 195 DNA methylomes reveals shared and cell-specific features of partially methylated domains. Genome Biol. 19, 150 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Brinkman, A. B. et al. Partially methylated domains are hypervariable in breast cancer and fuel widespread CpG island hypermethylation. Nat. Commun. 10, 1749 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Janssen, A., Colmenares, S. U. & Karpen, G. H. Heterochromatin: guardian of the genome. Annu. Rev. Cell Dev. Biol. 34, 265–288 (2018).

    Article  CAS  PubMed  Google Scholar 

  9. Padeken, J., Methot, S. P. & Gasser, S. M. Establishment of H3K9-methylated heterochromatin and its functions in tissue differentiation and maintenance. Nat. Rev. Mol. Cell Biol. 23, 623–640 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Allshire, R. C. & Madhani, H. D. Ten principles of heterochromatin formation and function. Nat. Rev. Mol. Cell Biol. 19, 229–244 (2018).

    Article  CAS  PubMed  Google Scholar 

  11. Wu, H. & Zhang, Y. Reversing DNA methylation: mechanisms, genomics, and biological functions. Cell 156, 45–68 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Li, E. & Zhang, Y. DNA methylation in mammals. Cold Spring Harb. Perspect. Biol. 6, a019133 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Tahiliani, M. et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 324, 930–935 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Okano, M., Xie, S. & Li, E. Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat. Genet. 19, 219–220 (1998).

    Article  CAS  PubMed  Google Scholar 

  15. Genovese, G. et al. Clonal hematopoiesis and blood-cancer risk inferred from blood DNA sequence. N. Engl. J. Med. 371, 2477–2487 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Jaiswal, S. et al. Age-related clonal hematopoiesis associated with adverse outcomes. N. Engl. J. Med. 371, 2488–2498 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Couronne, L., Bastard, C. & Bernard, O. A. TET2 and DNMT3A mutations in human T-cell lymphoma. N. Engl. J. Med. 366, 95–96 (2012).

    Article  CAS  PubMed  Google Scholar 

  18. Richards, E. J. & Elgin, S. C. Epigenetic codes for heterochromatin formation and silencing: rounding up the usual suspects. Cell 108, 489–500 (2002).

    Article  CAS  PubMed  Google Scholar 

  19. Rose, N. R. & Klose, R. J. Understanding the relationship between DNA methylation and histone lysine methylation. Biochim. Biophys. Acta 1839, 1362–1372 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. Lopez-Moyado, I. F. et al. Paradoxical association of TET loss of function with genome-wide DNA hypomethylation. Proc. Natl. Acad. Sci. USA 116, 16933–16942 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Fang, S. et al. Tet inactivation disrupts YY1 binding and long-range chromatin interactions during embryonic heart development. Nat Commun 10, 4297 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Li, J. et al. Decoding the dynamic DNA methylation and hydroxymethylation landscapes in endodermal lineage intermediates during pancreatic differentiation of hESC.Nucleic Acids Res. 46, 2883–2900 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Busch, K. et al. Fundamental properties of unperturbed haematopoiesis from stem cells in vivo. Nature 518, 542–546 (2015).

    Article  CAS  PubMed  Google Scholar 

  24. 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 

  25. Rundberg Nilsson, A. et al. Human and murine hematopoietic stem cell aging is associated with functional impairments and intrinsic megakaryocytic/erythroid bias. PLoS One 11, e0158369 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Yan, H. et al. Distinct roles for TET family proteins in regulating human erythropoiesis. Blood 129, 2002–2012 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Qu, X. et al. TET2 deficiency leads to stem cell factor-dependent clonal expansion of dysfunctional erythroid progenitors. Blood 132, 2406–2417 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Pang, W. W. et al. Human bone marrow hematopoietic stem cells are increased in frequency and myeloid-biased with age. Proc. Natl. Acad. Sci. USA 108, 20012–20017 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ko, M. et al. Ten-Eleven-Translocation 2 (TET2) negatively regulates homeostasis and differentiation of hematopoietic stem cells in mice. Proc. Natl. Acad. Sci. USA 108, 14566–14571 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Kowalczyk, M. S. et al. Single-cell RNA-seq reveals changes in cell cycle and differentiation programs upon aging of hematopoietic stem cells. Genome Res. 25, 1860–1872 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Kirschner, K. et al. Proliferation drives aging-related functional decline in a subpopulation of the hematopoietic stem cell compartment. Cell Rep. 19, 1503–1511 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Bergen, V. et al. Generalizing RNA velocity to transient cell states through dynamical modeling. Nat. Biotechnol. 38, 1408–1414 (2020).

    Article  CAS  PubMed  Google Scholar 

  33. Pastor, W. A. et al. Genome-wide mapping of 5-hydroxymethylcytosine in embryonic stem cells. Nature 473, 394–397 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Huang, Y., Pastor, W. A., Zepeda-Martinez, J. A. & Rao, A. The anti-CMS technique for genome-wide mapping of 5-hydroxymethylcytosine. Nat. Protoc. 7, 1897–1908 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Poleshko, A. et al. H3K9me2 orchestrates inheritance of spatial positioning of peripheral heterochromatin through mitosis. Elife 8, e49278 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  36. von Meyenn, F. et al. Impairment of DNA methylation maintenance is the main cause of global demethylation in naive embryonic stem cells. Mol. Cell 62, 848–861 (2016).

    Article  Google Scholar 

  37. Challen, G. A. et al. Dnmt3a is essential for hematopoietic stem cell differentiation. Nat. Genet. 44, 23–31 (2012).

    Article  CAS  Google Scholar 

  38. Pappalardi, M. B. et al. Discovery of a first-in-class reversible DNMT1-selective inhibitor with improved tolerability and efficacy in acute myeloid leukemia. Nat. Cancer 2, 1002–1017 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Chen, Q. et al. GSK-3484862 targets DNMT1 for degradation in cells. NAR. Cancer 5, zcad022 (2023).

    Article  Google Scholar 

  40. Hsieh, T. H. et al. Mapping nucleosome resolution chromosome folding in yeast by Micro-C. Cell 162, 108–119 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Zhang, X. et al. The loss of heterochromatin is associated with multiscale three-dimensional genome reorganization and aberrant transcription during cellular senescence. Genome Res. 31, 1121–1135 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Zhang, S. M. et al. KDM5B promotes immune evasion by recruiting SETDB1 to silence retroelements. Nature 598, 682–687 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Liu, X. et al. Resurrection of endogenous retroviruses during aging reinforces senescence. Cell 186, 287–304 e26 (2023).

    Article  CAS  PubMed  Google Scholar 

  44. Tsumura, A. et al. Maintenance of self-renewal ability of mouse embryonic stem cells in the absence of DNA methyltransferases Dnmt1, Dnmt3a and Dnmt3b. Genes Cells 11, 805–814 (2006).

    Article  CAS  PubMed  Google Scholar 

  45. Zhang, X. et al. DNMT3A and TET2 compete and cooperate to repress lineage-specific transcription factors in hematopoietic stem cells. Nat. Genet. 48, 1014–1023 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Wilkinson, A. C. et al. Long-term ex vivo haematopoietic-stem-cell expansion allows nonconditioned transplantation. Nature 571, 117–121 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Wolf, F. A., Angerer, P. & Theis, F. J. SCANPY: large-scale single-cell gene expression data analysis. Genome Biol. 19, 15 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Xi, Y. & Li, W. BSMAP: whole genome bisulfite sequence MAPping program. BMC Bioinformatics 10, 232 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  50. Stark R. & Brown G. DiffBind: differential binding analysis of ChIP-Seq peak data. Cancer Research UK – Cambridge Institute; University of Cambridge (2023).

  51. Sun, D. et al. MOABS: model based analysis of bisulfite sequencing data. Genome Biol. 15, R38 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Poleshko, A. et al. Genome-nuclear lamina interactions regulate cardiac stem cell lineage restriction. Cell 171, 573–587 e14 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Lund, E., Oldenburg, A. R. & Collas, P. Enriched domain detector: a program for detection of wide genomic enrichment domains robust against local variations. Nucleic Acids Res. 42, e92 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Robinson, J. T. et al. Integrative genomics viewer. Nat. Biotechnol. 29, 24–26 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Arifin, W. N. & Zahiruddin, W. M. Sample size calculation in animal studies using resource equation approach. Malays. J. Med. Sci. 24, 101–105 (2017).

    PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are grateful for the Epigenetic core at the Institute of Biosciences and Technology at the Texas A&M University. This work was supported by grants from Cancer Prevention and Research Institute of Texas (RP210070 to Y.Z.), National Institute of Health grants (R35HL166557, R01DK132286 and R01CA240258 to Y.H.; R01CA232017 and R21CA277257 to Y.Z.; P01AG036695, DK092883, CA183252, P30CA125123 and P01CA265748 to M.A.G.), Leukemia & Lymphoma Society (LLS-TRP 6680-24 to Y.Z.) and the Welch Foundation (BE-1913-20220331 to Y.Z.).

Author information

Authors and Affiliations

Authors

Contributions

Y.H. and Y.Z. directed and supervised the project. T.H. performed most animal-related work, molecular characterization and sequencing library preparation. J.L. performed all the bioinformatics analysis on high-throughput sequencing data. L.G., T.W. and S.F. supported sequencing library preparation. Y.D. performed cell sorting. A.D. and M.C. performed genotyping and supported molecular cloning. A.G., K.W., C.R. and C.K. supported animal-related work. Y.Y., C.C.Y., S.L. and M.J.Y. provided human bone marrow samples. M.A.G. and X.C. provided essential resources and key intellectual inputs to support this study. Y.H. and Y.Z. wrote the paper. All the authors participated in the discussion, data interpretation and paper editing or discussion.

Corresponding authors

Correspondence to Jia Li, Yubin Zhou or Yun Huang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Aging thanks Gerald de Haan, Francesco Neri, and Karl Lenhard Rudolph for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Tet2 depletion mitigates age-related decline in HSPC self-renewal and repopulation capability (Related to Fig. 1).

a, Immunoblot analysis of Tet2 expression in cKit+ cells from bone marrow of WT or Tet2KO mice (6 ~ 8-week-old, n = 3 male and 2 female mice). b, The total cell numbers (Top) and the fraction of LSK cells (Bottom) in the indicated replating times. CFU assay was used to assess HSPCs function. Data are shown as mean ± S.D; n = 3 biological replicates. (Two-way ANOVA test; Supplementary Table 1). c, The reconstitution efficiency of lethally irradiated CD45.1 mice transferred with LincKit+Sca1+ (LSK) cells (n = 3 male and 2 female donor mice / group). The reconstitution efficiency was measured biweekly from the peripheral blood till 16 weeks post-BMT (n = 5-7 recipient mice with equal number of male and female mice / group, each dot represents an individual recipient mouse). (Two-way ANOVA test; Supplementary Table 1). d, Lineage analysis in the peripheral blood collected from lethally irradiated CD45.1 mice transferred with LSK cells (n = 5-7 recipient mice /group, each dot represents an individual recipient mouse). Data are shown as mean ± S.D. (Two-way ANOVA test; Supplementary Table 1). e, The gating strategies. Lin, lineage-negative; LSK, LincKit+Sca1+; LT-HSC, long-term HSC; MPP, multipotential progenitor; ST-HSC, short-term HSC. f, Lineage analysis in the peripheral blood collected from recipient mice undergoing a second bone marrow transplantation using donor cells from the recipient mice transferred with young and old Tet2KO HSPCs listed in Extended Data Fig. 1c, d. Data are shown as mean ± S.D. (n = 3 male donor mice and 3-5 male recipient mice / group). g, Representative Hematoxylin and Eosin (H&E) stained sections of heart, liver, lung, and spleen tissues collected from the CD45.1 recipient mice with secondary BMT (18 weeks post-BMT). Donor cells were total bone marrow cells isolated from mice transplanted Tet2KO LSKs (young or old) as shown in Extended Data Fig. 1c, d (n = 3 donor mice). Scale bar, 200 μm. h, Lineage analysis in the peripheral blood collected from the recipient mice transferred with mixed HSPCs as shown in Fig. 1e. Data are shown as mean ± S.D. (Two-way ANOVA test; Supplementary Table 1).

Extended Data Fig. 2 Distinct transcriptional changes during WT and Tet2KO HSPCs aging (Related to Fig. 2).

a, Heatmap showing the expression of the marker genes used to annotate cell populations. b, Pseudotime inference reveals the differentiation process projected in UMAP-based embedding (shown in Fig. 1a) in all the analyzed cells. Dark blue represents cells in earlier differentiation stages; while dark red represents cells in later differentiation stages. c, The cell type ratio correlation analysis of scRNA-seq data collected in this study (WT-Young and Tet2KO-young) with the previously published dataset (GSE124822). d, Lineage analysis in the bone marrow collected from gender- and aged-matched congenic WT-young, WT-old, Tet2KO-young, and Tet2KO-old mice. Young, 6-8 weeks; Old, 18–24 months. Data are shown as mean ± S.D; n = 4-13 (equal number of male and female mice) / group, each dot represents an individual mouse. (Two-tailed Student’s t-test). e, (Left) Erythroid lineage analysis on bone marrow cells collected from gender- and aged-matched congenic WT-young, WT-old, Tet2KO-young, and Tet2KO-old mice. Data are shown as mean ± S.D; n = 4-12 (equal number of male and female mice) / group, each dot represents an individual recipient mouse (Two-tailed Student’s t-test). (Right) The gating strategies of erythroid progenitors. f, Heatmap showing the expression of marker genes used to annotate cell cycles. g, Heatmap representing genes associated with the pseudotime trajectory for HSC as shown in Fig. 2e.

Extended Data Fig. 3 Measurements of Tet expression and 5hmC in young and aged HSPC (Related to Fig. 3).

a, RT-qPCR analysis of Tet1 and Tet3 expression in LSK cells from WT-young, WT-old, Tet2KO-young, and Tet2KO-old mice. Data are shown as mean ± S.D (n = 3 independent experiments). b, Summary of Pearson’s correlation coefficients of 5hmC (left) and H3K9me3 (right) in the indicated groups. c, The percentage of age-related DHMRs in various genomic regions (distal, intergenic, genebody, 3’UTR, or promoter regions) for the indicated comparison groups (left, WT-Young vs WT-Old; right, Tet2KO-Young vs Tet2KO-Old). d, The top 5 enriched biological functions of age-related DHMRs identified from WT or Tet2KO HSPCs (young vs old) using the GREAT analysis. e, The genome browser views of 5hmC (blue, CMSIP) (scale: 0 to 5) in LSK cells purified from WT-young and WT-old mice. H3K9me3 track (scale: −0.5 to 0.5) in LSK cells purified from the WT-young group is used as the reference to discriminate between heterochromatin and euchromatin.

Extended Data Fig. 4 Tet2 loss prevents the spatial relocalization of H3K9me3-marked heterochromatin from the nuclear periphery to the nucleoplasm during HSPC aging (Related to Fig. 3).

a, Additional representative images of immunofluorescence (IF) staining of H3K9me3 in LSK cells purified from WT-young (6-8 weeks old) and WT-old (18 months old) mice. DAPI was used to stain the whole nuclei. Scale bar, 5 µm. b, Immunoblot analysis (left) and quantification (right) of the indicated histone marks in LK cells purified from WT-young, WT-old, Tet2KO-young, and Tet2KO-old mice. Data are shown as mean ± S.D; n = 3 independent biological replicates. c, Immunofluorescence analysis (top) and quantification (bottom) of H3K9me3 in CD34+ cells purified from human bone marrow at the indicated ages. Data are shown as mean ± S.D; n = 35-40 cells, Scale bar, 5 µm. d, (Left) Immunofluorescence staining of H3K4me3, H3K27ac, and lamin B1 in LSK cells. Scale bar, 5 µm. (Right) The quantification of fluorescence intensities of H3K4me3 and H3K27ac. Data are shown as mean ± S.D; n = 35-45 cells (n = 3-10 mice with equal number of male and female mice).

Extended Data Fig. 5 DNA methylation deregulation contributes to the spatial relocalization of H3K9me3 in HSPCs. (Related to Fig. 4).

a, Average DNA methylation levels within the coding regions in LSK cells purified WT-young (6-8 weeks), WT-old (18 months), Tet2KO-young (6-8 weeks), and Tet2KO-old (18 months) mice. b, Violin plot of DNA methylation levels in the indicated groups. c, Scatterplot of age-related DMRs (young vs old) identified from WGBS data in this study and the published WGBS data (GSE47819). The Pearson’s correlation coefficients were listed to compare the DNA methylation levels within age-related DMRs for the indicated groups. d, Genomic distribution of age-related DMRs (young vs old) identified from WT and Tet2KO groups. e, Genome browser views illustrating H3K9me3, WGBS, 5mC changes (∆5mC), the standard deviation of 5mC changes (5mC S.D.), and 5hmC in WT-young and WT-old HSPCs. f, PCR-based genotyping results using genomic DNA isolated from the tails collected from MxCre and MxCre-Dnmt3af/f mice treated with pIpC. Genotyping was performed 4 weeks after pIpC treatment. g, Average DNA methylation levels within the coding regions in LSK cells purified from MxCre (blue) and MxCre-Dnmt3af/f mice (red) treated with pIpC and analyzed 4 weeks after pIpC treatment. h, RT-qPCR analysis of Dnmt1 expression in LSK cells purified from WT-young, WT-old, Tet2KO-young and Tet2KO-old mice. Data are shown as mean ± S.D (n = 3). (Two-tailed Student’s t-test). i, Summary of the Pearson’s correlation coefficients of H3K9me3 ChIP-seq data for LSK cells purified from MxCre (control) or MxCre-Dnmt3af/f (Dnmt3aKO) mice treated with pIpC. j, The scatterplot analysis to identify genomic regions exhibiting significant differential H3K9me3 enrichment between MxCre (control) and MxCre-Dnmt3af/f (Dnmt3aKO) HSPCs. k, Genome browser view of H3K9me3 in LSK cells purified from MxCre (control) and MxCre-Dnmt3af/f mice treated with pIpC. l, Schematic illustrating the impact of aging, Tet2 and Dnmt3a in regulating the spatial localization of H3K9me3-marked heterochromatin.

Extended Data Fig. 6 Upregulation of repetitive elements in aged HSPCs contributes to their functional decline (Related to Fig. 5).

a, Single-cell expression profiles of repetitive elements from each cell cluster in all the analyzed Lin cells (n = 8,352). b, UMAP dimensionality reduction showing the cell distribution, taking into account the expression of repetitive elements as listed in Extended Data Fig. 6a, of all analyzed Lin cells with high coverage (more than 300 genes and 30000 UMIs) (n = 8,352) purified from WT-young, WT-old, Tet2KO-young and Tet2KO-old mice. Young mice: 6-8 weeks old; Old mice: 18–24 months old. HSC, hematopoietic stem cells; Mono, monocyte progenitor; Ery, erythroid progenitor; MkP, megakaryocyte progenitor; Bcell-P: B cell progenitors. c, (Left) UMAP dimensionality reduction (n = 929 cells) in HSCs identified from the analysis performed in panel A. Four distinct clusters were identified based on the gene expression signatures including repetitive elements listed in Extended Data Fig. 6b. (Right) UMAP plot showing the distribution of HSCs in the indicated groups. d, Venn diagram to illustrate the overlapping genomic regions with age-related compartment changes in WT and Tet2KO groups. e, Violin plot of 5hmC levels in H3K9me3-marked genomic regions with age-related compartment changes in the indicated groups. f, Normalized expression of SINE and LINE elements within H3K9me3-marked genomic regions with age-related compartment changes (B-to-A or A-to-B) in the indicated groups.

Supplementary information

Reporting Summary

Supplementary Table 1

ANOVA test.

Supplementary Table 2

Marker genes in scRNA-seq analysis.

Supplementary Table 3

Age-related DEGs.

Supplementary Table 4

Age-related DHMRs.

Supplementary Table 5

Age-related DMR analysis.

Supplementary Table 6

List of qPCR primers.

Source data

Source Data Fig. 1

Unprocessed western blots.

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 5

Unprocessed western blots.

Statistical Source Data

Source data for statistical analysis.

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hong, T., Li, J., Guo, L. et al. TET2 modulates spatial relocalization of heterochromatin in aged hematopoietic stem and progenitor cells. Nat Aging 3, 1387–1400 (2023). https://doi.org/10.1038/s43587-023-00505-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43587-023-00505-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