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:

Locus-specific induction of gene expression from heterochromatin loci during cellular senescence

Abstract

Senescence is a fate-determined state, accompanied by reorganization of heterochromatin. Although lineage-appropriate genes can be temporarily repressed through facultative heterochromatin, stable silencing of lineage-inappropriate genes often involves the constitutive heterochromatic mark, histone H3 lysine 9 trimethylation (H3K9me3). The fate of these heterochromatic genes during senescence is unclear. In the present study, we show that a small number of lineage-inappropriate genes, exemplified by the LCE2 skin genes, are derepressed during senescence from H3K9me3 regions in fibroblasts. DNA FISH experiments reveal that these gene loci, which are condensed at the nuclear periphery in proliferative cells, are decompacted during senescence. Decompaction of the locus is not sufficient for LCE2 expression, which requires p53 and C/EBPβ signaling. NLRP3, which is predominantly expressed in macrophages from an open topologically associated domain (TAD), is also derepressed in senescent fibroblasts due to the local disruption of the H3K9me3-rich TAD that contains it. NLRP3 has been implicated in the amplification of inflammatory cytokine signaling in senescence and aging, highlighting the functional relevance of gene induction from ‘permissive’ H3K9me3 regions in senescent 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: Aberrant expression of lineage-inappropriate LCE2 genes during senescence.
Fig. 2: Decompaction of EDC locus in senescent IMR90 cells.
Fig. 3: Decompaction of EDC locus and senescence effectors.
Fig. 4: H3K9me3 peak characterization.
Fig. 5: Aberrant expression of lineage-specific genes and chromatin decompaction during senescence.
Fig. 6: Distinct mode of expression of lineage-appropriate and lineage-inappropriate genes during RIS.
Fig. 7: Differential 3D behavior of permissive H3K9me3 peaks.

Similar content being viewed by others

Data availability

Human KC RNA-seq, ChIP-seq data, IMR90 ATAC-seq data (RIS) and RNA-seq data generated for the present study have been deposited at the GEO (accession nos. GSE130457 and GSE180469). Control ATAC-seq data were previously published (accession no. GSE103590)9. Microarray datasets (Extended Data Fig. 1) have been previously published (accession nos. GSE53379 and GSE59522)15,26. RNA-seq data for RIS/DDIS IMR90 cells are available at accession no. GSE72407 (ref. 21). Other senescence RNA-seq data17,18 were downloaded from GSE61130 (ref. 17), GSE74238 (ref. 6), GSE45833 (ref. 18), GSE85082, GSE63577, GSE56293 and GSE130306 (ref. 19). SF RNA-seq was downloaded from accession no. GSE163749 (ref. 46). HPA RNA-seq data were downloaded from the HPA (https://www.proteinatlas.org), accession no. E-MTAB-2836 (ref. 34). THP-1 RNA-seq data were downloaded from accession no. GSE96800 (ref. 45). ChIP-seq for H3K9m3, H3K27me3 (ref. 10), H3K27ac9 and Lamin B1 (ref. 11) are available at accession nos. GSE38448, GSE103590 and GSE49341, respectively. ChIP-seq data for hES cells were downloaded from ENCODE, accession no. GSE29611 (ref. 63). H3K27ac ChIP-seq and ATAC-seq data for THP-1 cells were downloaded from the GEO, accession no. GSE96800 (ref. 45). Hi-C data (IMR90 cells) can be found at accession no. GSE135090. Hi-C data (human KCs) were downloaded from accession no. GSE84660. Hi-C data (THP-1 cells) were downloaded from the Sequence Read Archive accession no. PRJNA385337 (ref. 45). Hi-C data (WI38 cells) were downloaded from accession no. GSE130306 (ref. 19). Source data are provided with this paper.

Code availability

No customized code was used. FISH signal length measurement was performed using LAS-AF software (v.4.0.0). XZ re-sliced images were generated from Z-stack images using ImageJ (v.2.1.0). GraphPad Prism 6.0, 7.0, 8.0 and 9.0 were used for general statistical analyses. For sequence data analysis, BWA (v.0.7.12), Cutadapt (v.1.10.0), ChIPQC Bioconductor package (v.1.14.0), MACS2 (v.2.1.0), bedtools (v.2.26.0), EDD (v.1.1.19), THOR (v.0.11.4), STAR aligner (v.2.5.3a), Rsubread (v.1.5.2), edgeR (v.3.20.9), GEOquery package (v.2.60.0), limma package (v.3.48.3), HiC-Pro (v.2.10.0), FastQC (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/), diffHiC (v.1.10.0) and HiCvizR (v.3.4.4) were used.

References

  1. Gorgoulis, V. et al. Cellular senescence: Defining a path forward. Cell 179, 813–827 (2019).

    Article  CAS  PubMed  Google Scholar 

  2. Chan, A. S. L. & Narita, M. Short-term gain, long-term pain: the senescence life cycle and cancer. Gene Dev. 33, 127–143 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Faget, D. V., Ren, Q. & Stewart, S. A. Unmasking senescence: context-dependent effects of SASP in cancer. Nat. Rev. Cancer 19, 439–453 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Rai, T. S. & Adams, P. D. Lessons from senescence: chromatin maintenance in non-proliferating cells. Biochim. Biophys. Acta 1819, 322–331 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Nacarelli, T., Liu, P. & Zhang, R. Epigenetic basis of cellular senescence and its implications in aging. Genes 8, 343 (2017).

    Article  PubMed Central  Google Scholar 

  6. Tasdemir, N. et al. BRD4 connects enhancer remodeling to senescence immune surveillance. Cancer Discov. 6, 612–629 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Sen, P. et al. Histone acetyltransferase p300 induces de novo super-enhancers to drive cellular senescence. Mol. Cell 73, 684–698.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Cecco, M. et al. Genomes of replicatively senescent cells undergo global epigenetic changes leading to gene silencing and activation of transposable elements. Aging Cell 12, 247–256 (2013).

    Article  PubMed  Google Scholar 

  9. Parry, A. J. et al. NOTCH-mediated non-cell autonomous regulation of chromatin structure during senescence. Nat. Commun. 9, 1840 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Chandra, T. et al. Independence of repressive histone marks and chromatin compaction during senescent heterochromatic layer formation. Mol. Cell 47, 203–214 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Sadaie, M. et al. Redistribution of the Lamin B1 genomic binding profile affects rearrangement of heterochromatic domains and SAHF formation during senescence. Gene Dev. 27, 1800–1808 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Swanson, E. C., Manning, B., Zhang, H. & Lawrence, J. B. Higher-order unfolding of satellite heterochromatin is a consistent and early event in cell senescence. J. Cell Biol. 203, 929–942 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Dambacher, S., Hahn, M. & Schotta, G. Epigenetic regulation of development by histone lysine methylation. Heredity 105, 24–37 (2010).

    Article  CAS  PubMed  Google Scholar 

  14. Becker, J. S., Nicetto, D. & Zaret, K. S. H3K9me3-dependent heterochromatin: barrier to cell fate changes. Trends Genet. 32, 29–41 (2016).

    Article  CAS  PubMed  Google Scholar 

  15. Kirschner, K. et al. Phenotype specific analyses reveal distinct regulatory mechanism for chronically activated p53. PLoS Genet. 11, e1005053 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Jackson, B. et al. Late cornified envelope family in differentiating epithelia—response to calcium and ultraviolet irradiation. J. Invest. Dermatol. 124, 1062–1070 (2005).

    Article  CAS  PubMed  Google Scholar 

  17. Herranz, N. et al. mTOR regulates MAPKAPK2 translation to control the senescence-associated secretory phenotype. Nat. Cell Biol. 17, 1205–1217 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Loayza-Puch, F. et al. p53 induces transcriptional and translational programs to suppress cell proliferation and growth. Genome Biol. 14, R32 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Sati, S. et al. 4D genome rewiring during oncogene-induced and replicative senescence. Mol. Cell 78, 522–538.e9 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Bergboer, J. G. M. et al. Psoriasis risk genes of the late cornified envelope-3 group are distinctly expressed compared with genes of other LCE groups. Am. J. Pathol. 178, 1470–1477 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Hoare, M. et al. NOTCH1 mediates a switch between two distinct secretomes during senescence. Nat. Cell Biol. 18, 979–992 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Deng, Z. et al. Late cornified envelope group I, a novel target of p53, regulates PRMT5 activity. Neoplasia 16, 656–664 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Ezhkova, E. et al. Ezh2 orchestrates gene expression for the stepwise differentiation of tissue-specific stem cells. Cell 136, 1122–1135 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Narita, M. et al. Rb-mediated heterochromatin formation and silencing of E2F target genes during cellular senescence. Cell 113, 703–716 (2003).

    Article  CAS  PubMed  Google Scholar 

  25. Mardaryev, A. N. et al. p63 and Brg1 control developmentally regulated higher-order chromatin remodelling at the epidermal differentiation complex locus in epidermal progenitor cells. Development 141, 101–111 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Young, A. R. J. et al. Autophagy mediates the mitotic senescence transition. Gene Dev. 23, 798–803 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Narita, M. et al. A novel role for high-mobility group A proteins in cellular senescence and heterochromatin formation. Cell 126, 503–514 (2006).

    Article  CAS  PubMed  Google Scholar 

  28. Shimi, T. et al. The role of nuclear lamin B1 in cell proliferation and senescence. Gene Dev. 25, 2579–2593 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Rodier, F. et al. Persistent DNA damage signalling triggers senescence-associated inflammatory cytokine secretion. Nat. Cell Biol. 11, 973–979 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Kuilman, T. et al. Oncogene-induced senescence relayed by an interleukin-dependent inflammatory network. Cell 133, 1019–1031 (2008).

    Article  CAS  PubMed  Google Scholar 

  31. Thurman, R. E. et al. The accessible chromatin landscape of the human genome. Nature 489, 75–82 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Stergachis, A. B. et al. Developmental fate and cellular maturity encoded in human regulatory DNA landscapes. Cell 154, 888–903 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Jordan, I. K., Rogozin, I. B., Glazko, G. V. & Koonin, E. V. Origin of a substantial fraction of human regulatory sequences from transposable elements. Trends Genet. 19, 68–72 (2003).

    Article  CAS  PubMed  Google Scholar 

  34. Uhlén, M. et al. Proteomics. Tissue-based map of the human proteome. Science 347, 1260419 (2015).

    Article  PubMed  Google Scholar 

  35. Lin, A. W. & Lowe, S. W. Oncogenic ras activates the ARF-p53 pathway to suppress epithelial cell transformation. Proc. Natl Acad. Sci. USA 98, 5025–5030 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Gaspar-Maia, A., Alajem, A., Meshorer, E. & Ramalho-Santos, M. Open chromatin in pluripotency and reprogramming. Nat. Rev. Mol. Cell Biol. 12, 36–47 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Chandra, T. et al. Global reorganization of the nuclear landscape in senescent cells. Cell Rep. 10, 471–483 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Iwasaki, O. et al. Involvement of condensin in cellular senescence through gene regulation and compartmental reorganization. Nat. Commun. 10, 5688 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Olan, I. et al. Transcription-dependent cohesin repositioning rewires chromatin loops in cellular senescence. Nat. Commun. 11, 6049 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Rubin, A. J. et al. Lineage-specific dynamic and pre-established enhancer–promoter contacts cooperate in terminal differentiation. Nat. Genet. 49, 1522–1528 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Schoenfelder, S. & Fraser, P. Long-range enhancer-promoter contacts in gene expression control. Nat. Rev. Genet. 20, 437–455 (2019).

    Article  CAS  PubMed  Google Scholar 

  42. Swanson, K. V., Deng, M. & Ting, J. P.-Y. The NLRP3 inflammasome: molecular activation and regulation to therapeutics. Nat. Rev. Immunol. 19, 477–489 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Acosta, J. C. et al. A complex secretory program orchestrated by the inflammasome controls paracrine senescence. Nat. Cell Biol. 15, 978–990 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Kong, F. et al. Curcumin represses NLRP3 inflammasome activation via TLR4/MyD88/NF-κB and P2X7R signaling in PMA-induced macrophages. Front. Pharmacol. 7, 369 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Phanstiel, D. H. et al. Static and dynamic DNA loops form AP-1-bound activation hubs during macrophage development. Mol. Cell 67, 1037–1048.e6 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Friščić, J. et al. The complement system drives local inflammatory tissue priming by metabolic reprogramming of synovial fibroblasts. Immunity 54, 1002–1021.e10 (2021).

    Article  PubMed  Google Scholar 

  47. Brown, S. J. et al. Rodent LCE gene clusters: new nomenclature, gene organization, and divergence of human and rodent genes. J. Invest. Dermatol. 127, 1782–1786 (2007).

    Article  CAS  PubMed  Google Scholar 

  48. Robbins, P. D. et al. Senolytic drugs: reducing senescent cell viability to extend health span. Annu. Rev. Pharmacol 61, 779–803 (2020).

    Article  Google Scholar 

  49. Shailesh, H., Zakaria, Z. Z., Baiocchi, R. & Sif, S. Protein arginine methyltransferase 5 (PRMT5) dysregulation in cancer. Oncotarget 9, 36705–36718 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Li, Y.-H. et al. PRMT5-TRIM21 interaction regulates the senescence of osteosarcoma cells by targeting the TXNIP/p21 axis. Aging 12, 2507–2529 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Croft, A. P. et al. Rheumatoid synovial fibroblasts differentiate into distinct subsets in the presence of cytokines and cartilage. Arthritis Res. Ther. 18, 270 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Castets, M. et al. DCC constrains tumour progression via its dependence receptor activity. Nature 482, 534–537 (2011).

    Article  PubMed  Google Scholar 

  53. Kimura, H., Hayashi-Takanaka, Y., Goto, Y., Takizawa, N. & Nozaki, N. The organization of histone H3 modifications as revealed by a panel of specific monoclonal antibodies. Cell Struct. Funct. 33, 61–73 (2008).

    Article  CAS  PubMed  Google Scholar 

  54. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. Embnet J. 17, 10–12 (2011).

    Article  Google Scholar 

  55. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  57. 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 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Karolchik, D. et al. The UCSC table browser data retrieval tool. Nucleic Acids Res. 32, 493D–496D (2004).

    Article  Google Scholar 

  59. Davis, S. & Meltzer, P. S. GEOquery: a bridge between the Gene Expression Omnibus (GEO) and bioconductor. Bioinformatics 23, 1846–1847 (2007).

    Article  PubMed  Google Scholar 

  60. 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  Google Scholar 

  61. Kryuchkova-Mostacci, N. & Robinson-Rechavi, M. A benchmark of gene expression tissue-specificity metrics. Brief Bioinform. https://doi.org/10.1093/bib/bbw008 (2016).

  62. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Dunham, I. et al. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

    Article  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  65. 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 

  66. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Allhoff, M., Seré, K., Pires, J. F., Zenke, M. & Costa, I. G. Differential peak calling of ChIP-seq signals with replicates with THOR. Nucleic Acids Res. 44, gkw680 (2016).

    Article  Google Scholar 

  68. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  69. Servant, N. et al. HiC-Pro: an optimized and flexible pipeline for Hi-C data processing. Genome Biol. 16, 259 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  70. Lun, A. T. L. & Smyth, G. K. diffHic: a Bioconductor package to detect differential genomic interactions in Hi-C data. BMC Bioinform. 16, 258 (2015).

    Article  Google Scholar 

  71. Imakaev, M. et al. Iterative correction of Hi-C data reveals hallmarks of chromosome organization. Nat. Methods 9, 999–1003 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank all members of the Narita laboratory for helpful discussions, A. Young for editing and the staff of the CRUK-CI core facilities (Microscopy, Genomics, Biorepository, Bioinformatics and Research Instrumentation) for technical support, and T. Tsuda and E. Nishihama at the Ehime University for isolation and expansion of human primary KCs. This work is supported by a Cancer Research UK Cambridge Institute core grant (no. C9545/A29580) to Masashi N.’s laboratory, and in part by the Human Frontier Science Program (no. RGY0078/2010) (to Masashi N.). Masashi N. is also supported by the Medical Research Council (MRC) (MR/M013049/1 and MR/R010013/1), BBSRC (BB/S013466/1) and Diabetes UK via BIRAX and the British Council (65BX18MNIB). I.O. is supported by the Cancer Research UK Pioneer grant (no. C63389/A30462). S.A.S and D.B. are supported by the MRC (MC_UU_12022/10). S.S. is supported by the UKRI Biotechnology and Biological Science Research Council (BB/J004480/1), the UKRI MRC (MR/T016787/1) and a Career Progression Fellowship from the Babraham Institute. P.J.R.-G. is supported by the BBSRC (BBS/E/B/000C0422), the MRC (MR/T011769/1 and MR/N018419/1) and the Wellcome Trust (215116/Z/18/Z). H.K. is supported by JSPS KAKENHI (nos. JP17H01417 and JP18H05527). The funders had no role in study design, data collection and analysis, the decision to publish or the preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

Masashi N., S.A.S. and K.T. conceived the project. K.T., A.J.P., S.S., Y.I., K.K., C.B.R., Masako N., T.S., H. Kugoh, P.J.R.-G., P.F. and Masashi N. designed and performed experiments, and analyzed the data. D.B., I.O., G.St.C.S., A.S.L.C. and S.A.S analyzed the sequencing data. I.O. analyzed the Hi-C data. H. Kimura provided histone antibodies. K. Shiraishi and K. Sayama provided the human KCs. Masashi N. and K.T. wrote the manuscript with input from all authors.

Corresponding authors

Correspondence to Masako Narita, Shamith A. Samarajiwa or Masashi Narita.

Ethics declarations

Competing interests

P.F. and S.S. are co-founders of Enhanc3D Genomics Ltd. All other authors declare no competing interests.

Peer review information Nature Aging thanks Peter Adams, Rugang Zhang and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

Peer review information Nature Aging thanks Peter Adams, Rugang Zhang 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 Transcriptional profile of genes in the EDC locus.

(a) Image corrected, quantile normalized and log2 transformed microarray intensity values were downloaded from the indicated NCBI GEO entries. Control, proliferative; RIS, HRAS-G12V-induced senescent; Qui, quiescent; DDIS, DNA damage-induced senescent; acDDR, acute DNA damage response to etoposide (24 or 48 hours); E1A + RAS, transformed/senescence-bypassed. Genes for which there is no available transcript probe on the microarray platform used are greyed out. (b) Meta-analysis of senescence-associated transcriptomic changes of genes in the EDC locus using RNA-seq datasets, downloaded from indicated NCBI GEO, comparing between different cellular phenotypes. The datasets were processed using the same analysis pipeline. Values for the indicated genes are shown for each of the senescent samples and their corresponding growing controls. Values are Transcripts-Per-Million (TPM). Genes are ordered according to their position along Chr. 1. Genes for which there were no aligned reads in the specified RNA-seq runs are greyed out. (c) Immunoblot analysis of IMR90 cells stably expressing vector (-) or HRAS-G12V and control miR30 vector (-) or sh-p53 (+) for the proteins indicated. Arrow indicates specific bands. p21 is a p53-dependent cell cycle inhibitor that is upregulated during senescence. Experiments were performed at least 3 times. (d-e) IMR90 cells expressing ectopic mVenus-LCE2A and control vector were assessed for cumulative population doublings (d) and colony formation capacity P-value: 0.0193 (Two-tailed Paired t-test) (e) (n = 3). Error bars indicate mean + /− SD. The inset shows representative immunoblots of indicated protein (d).

Extended Data Fig. 2 Correlation between EDC decompaction and SAHF formation.

(a) 3D confocal images of DNA-FISH for the EDC locus in the cells indicated (see Fig. 2a for two alternate color probe designs). The perinuclear localization of the EDC locus was quantified on visual inspections by counting FISH signal, which attached at least 1 of the 7 tiled signals to the nuclear periphery (white arrowhead). Bars are % of signals on the nuclear periphery (n > 100 alleles from 3 biological replicates). (b) Representative images of DNA-FISH for the EDC and OR loci using the probes indicated and quantification of signal length. Note, for the EDC, 5 out of 7 probes (Fig. 2a) were used. Bars are median with inter-quartile range. Two-tailed Mann-Whitney-Wilcoxon test (n ≥ 107 alleles from 3 biological replicates) (c) FISH signal length of EDC on SAHF-negative or positive RIS cells. Bars are median with inter-quartile range. Two-tailed Mann-Whitney-Wilcoxon test (n = 100 alleles from 3 biological replicates). (d) LCE2 immunofluorescence signal on SAHF-negative or positive RIS cells (n = 3). (e-h) Quantification of SAHF-positive cells in the indicated conditions (e, g) and comparative analysis of LCE2 expression by immunoblotting in the conditions indicated (f, h). Arrows indicate specific bands on the immunoblotting. Bars indicate mean + /- s.e.m. One-way ANOVA with Tukey’s multiple comparisons test (n = 3 biological replicates) (e left, g). Two-tailed unpaired t-test (n = 3 biological replicates) (e right). (i) Quantification of FISH signal length in the conditions indicated for BJ and WI38 cells. Bars are median with interquartile range. Two-tailed Mann-Whitney-Wilcoxon test (n = 150 alleles from 3 biological replicates).

Extended Data Fig. 3 H3K9me3 exhibits increases in accessibility and H3K27ac ChIP-seq signal in RIS.

(a) ATAC-seq and H3K27ac THOR-normalized signal in proliferative and RIS cells centred on three classes of ATAC-seq peaks: proliferative-specific, RIS-specific, and common between the two conditions. (b) Number of ATAC-seq (top) and H3K27ac (bottom) peaks (in the consensus peak sets in the n = 3 proliferative and n = 3 RIS ATAC-seq samples and n = 4 proliferative and n = 4 RIS H3K27ac ChIP-seq samples) across the three groups of H3K9me3 peaks defined as permissive (group 2 n = 110 peaks and group 3 n = 26 peaks) and non-permissive (group 1 n = 627 peaks). The boxplots correspond to the median, 25th to 75th percentiles, and the whiskers correspond to the 10th to 90th percentiles. (c) IGV tracks of genes of interest and position of SINE and Alu repeats (UCSC annotation) over their gene bodies. (d) Region surrounding the SERPINA1 gene characterised by H3K9me3 ChIP-seq signal, as well as protruding H3K27ac and ATAC-seq peaks.

Extended Data Fig. 4 Tissue specificity of genes in H3K9me3 peaks.

(a) Tissue specificity score of the genes within the three groups of H3K9me3 peaks (group 1 n = 627 peaks, group 2 n = 110 peaks and group 3 n = 26 peaks) against all other genes. The Tau score ranges from 0 - ubiquitous expression across tissues to 1 - tissue specific; the 8 genes of interest in group 3 are highlighted. Significance testing was performed with two-sided Student’s t-tests, ***P < 0.001, n.s., not significant. The exact P-values are: 1 vs 2 0.7492, 2 vs 3 1.281e-06, 1 vs 3 1.594e-13, 1 vs Other 2.2e-16, 2 vs Other 2.2e-16, 3 vs Other 1.944e-07. The boxplots correspond to the median, 25th to 75th percentiles, and the whiskers correspond to the 10th to 90th percentiles. (b) DCC (left) and TMEM132B (right) genes represented by multiprobe DNA-FISH signals in the proliferative and RIS conditions. Probe sets are shown in Fig. 5b. Note, single colour probe images are shown for TMEM132B.

Extended Data Fig. 5 Three-dimensional interaction changes occurring during RIS.

(a) 200 kb resolution Hi-C maps in proliferative and RIS cells as well as significant interaction changes between the two conditions, where green arcs represent increases and blue arcs represent decreases. (b) Significant H3K9me3-H3K9me3 interaction changes (200 kb bins): log-fold changes (logFC) against distance between the interacting regions, dotted line corresponds to 10 Mb distance. (c) Hi-C interaction profiles at 40 kb resolution in proliferating (Prolif) and RIS IMR90 cells as well as differentiating keratinocytes at the EDC/OR loci. (d) Epigenetic profile of the NLRP3 locus in terms of H3K9me3, LaminB1 and H3K27me3 THOR-normalized ChIP-seq signal in proliferating and RIS cells. (e) Normalized gene expression (log-counts per million) of pro-inflammatory genes, EDC genes and senescence-representative genes in -/MSU and MSU/MSU synovial fibroblasts. *Adjusted P < 0.05, **P < 0.01, ***P < 0.001.

Supplementary information

Reporting Summary.

43587_2021_147_MOESM2_ESM.xlsx

Supplementary Table: Genes included in each H3K9me3-peak group. List of genes within the three types of H3K9me3 peaks determined; 38 genes significantly upregulated in group 3 are shown in bold text.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tomimatsu, K., Bihary, D., Olan, I. et al. Locus-specific induction of gene expression from heterochromatin loci during cellular senescence. Nat Aging 2, 31–45 (2022). https://doi.org/10.1038/s43587-021-00147-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43587-021-00147-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