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:

Histone H3K9 methylation is dispensable for Caenorhabditis elegans development but suppresses RNA:DNA hybrid-associated repeat instability

Abstract

Histone H3 lysine 9 (H3K9) methylation is a conserved modification that generally represses transcription. In Caenorhabditis elegans it is enriched on silent tissue-specific genes and repetitive elements. In met-2 set-25 double mutants, which lack all H3K9 methylation (H3K9me), embryos differentiate normally, although mutant adults are sterile owing to extensive DNA-damage-driven apoptosis in the germ line. Transposons and simple repeats are derepressed in both germline and somatic tissues. This unprogrammed transcription correlates with increased rates of repeat-specific insertions and deletions, copy number variation, R loops and enhanced sensitivity to replication stress. We propose that H3K9me2 or H3K9me3 stabilizes and protects repeat-rich genomes by suppressing transcription-induced replication stress.

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

Figure 1: Worms lacking H3K9me were viable but showed stochastically delayed development.
Figure 2: DNA-damage-checkpoint-dependent increase of apoptotic cells in the germ line of met-2 set-25 worms.
Figure 3: Differential enrichment of H3K9me2 and H3K9me3 on repeat element classes and gene types.
Figure 4: Temperature-dependent derepression of subsets of genes and repeat families in embryos and gonads in met-2 set-25 worms.
Figure 5: met-2 set-25 worms accumulate RNA:DNA hybrids at repeat elements.
Figure 6: The met-2 set-25 strain is hydroxyurea sensitive and accumulates mutations in repeat elements and reactivated transposable elements.
Figure 7: Somatic accumulation of indels leading to frameshift mutations in met-2 set-25 mutant larvae.
Figure 8: Transcribed REs in H3K9me-deficient strains can exacerbate replication stress provoking genomic instability.

Similar content being viewed by others

Accession codes

Primary accessions

Sequence Read Archive

References

  1. Bannister, A.J. et al. Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature 410, 120–124 (2001).

    CAS  PubMed  Google Scholar 

  2. Lachner, M., O'Carroll, D., Rea, S., Mechtler, K. & Jenuwein, T. Methylation of histone H3 lysine 9 creates a binding site for HP1 proteins. Nature 410, 116–120 (2001).

    CAS  PubMed  Google Scholar 

  3. Nakayama, J., Rice, J.C., Strahl, B.D., Allis, C.D. & Grewal, S.I. Role of histone H3 lysine 9 methylation in epigenetic control of heterochromatin assembly. Science 292, 110–113 (2001).

    CAS  PubMed  Google Scholar 

  4. Allshire, R.C., Javerzat, J.P., Redhead, N.J. & Cranston, G. Position effect variegation at fission yeast centromeres. Cell 76, 157–169 (1994).

    CAS  PubMed  Google Scholar 

  5. Cam, H.P. et al. Comprehensive analysis of heterochromatin- and RNAi-mediated epigenetic control of the fission yeast genome. Nat. Genet. 37, 809–819 (2005).

    CAS  PubMed  Google Scholar 

  6. Lehnertz, B. et al. Suv39h-mediated histone H3 lysine 9 methylation directs DNA methylation to major satellite repeats at pericentric heterochromatin. Curr. Biol. 13, 1192–1200 (2003).

    CAS  PubMed  Google Scholar 

  7. Peters, A.H. et al. Loss of the Suv39h histone methyltransferases impairs mammalian heterochromatin and genome stability. Cell 107, 323–337 (2001).

    CAS  PubMed  Google Scholar 

  8. Eissenberg, J.C. et al. Mutation in a heterochromatin-specific chromosomal protein is associated with suppression of position–effect variegation in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 87, 9923–9927 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Epsztejn-Litman, S. et al. De novo DNA methylation promoted by G9a prevents reprogramming of embryonically silenced genes. Nat. Struct. Mol. Biol. 15, 1176–1183 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Wissmann, M. et al. Cooperative demethylation by JMJD2C and LSD1 promotes androgen receptor–dependent gene expression. Nat. Cell Biol. 9, 347–353 (2007).

    CAS  PubMed  Google Scholar 

  11. Chen, M.W. et al. H3K9 histone methyltransferase G9a promotes lung cancer invasion and metastasis by silencing the cell adhesion molecule Ep-CAM. Cancer Res. 70, 7830–7840 (2010).

    CAS  PubMed  Google Scholar 

  12. Hua, K.T. et al. The H3K9 methyltransferase G9a is a marker of aggressive ovarian cancer that promotes peritoneal metastasis. Mol. Cancer 13, 189 (2014).

    PubMed  PubMed Central  Google Scholar 

  13. Kumari, D. & Usdin, K. The distribution of repressive histone modifications on silenced FMR1 alleles provides clues to the mechanism of gene silencing in fragile X syndrome. Hum. Mol. Genet. 19, 4634–4642 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Lee, J. et al. Epigenetic regulation of cholinergic receptor M1 (CHRM1) by histone H3K9me3 impairs Ca2+ signaling in Huntington's disease. Acta Neuropathol. 125, 727–739 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Loh, Y.H., Zhang, W., Chen, X., George, J. & Ng, H.H. Jmjd1a and Jmjd2c histone H3 Lys 9 demethylases regulate self-renewal in embryonic stem cells. Genes Dev. 21, 2545–2557 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Shi, Y. et al. A combined chemical and genetic approach for the generation of induced pluripotent stem cells. Cell Stem Cell 2, 525–528 (2008).

    CAS  PubMed  Google Scholar 

  17. Treangen, T.J. & Salzberg, S.L. Repetitive DNA and next-generation sequencing: computational challenges and solutions. Nat. Rev. Genet. 13, 36–46 (2011).

    PubMed  PubMed Central  Google Scholar 

  18. Dodge, J.E., Kang, Y.K., Beppu, H., Lei, H. & Li, E. Histone H3-K9 methyltransferase ESET is essential for early development. Mol. Cell. Biol. 24, 2478–2486 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Kondo, Y. et al. Downregulation of histone H3 lysine 9 methyltransferase G9a induces centrosome disruption and chromosome instability in cancer cells. PLoS One 3, e2037 (2008).

    PubMed  PubMed Central  Google Scholar 

  20. Tachibana, M. et al. G9a histone methyltransferase plays a dominant role in euchromatic histone H3 lysine 9 methylation and is essential for early embryogenesis. Genes Dev. 16, 1779–1791 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Mellone, B.G. et al. Centromere silencing and function in fission yeast is governed by the amino terminus of histone H3. Curr. Biol. 13, 1748–1757 (2003).

    CAS  PubMed  Google Scholar 

  22. Peng, J.C. & Karpen, G.H. H3K9 methylation and RNA interference regulate nucleolar organization and repeated DNA stability. Nat. Cell Biol. 9, 25–35 (2007).

    CAS  PubMed  Google Scholar 

  23. Andersen, E.C. & Horvitz, H.R. Two C. elegans histone methyltransferases repress lin-3 EGF transcription to inhibit vulval development. Development 134, 2991–2999 (2007).

    CAS  PubMed  Google Scholar 

  24. Towbin, B.D. et al. Step-wise methylation of histone H3K9 positions heterochromatin at the nuclear periphery. Cell 150, 934–947 (2012).

    CAS  PubMed  Google Scholar 

  25. Harris, J. et al. Mutator phenotype of Caenorhabditis elegans DNA damage checkpoint mutants. Genetics 174, 601–616 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Bernard, P. et al. Requirement of heterochromatin for cohesion at centromeres. Science 294, 2539–2542 (2001).

    CAS  PubMed  Google Scholar 

  27. Batista, P.J. et al. PRG-1 and 21U-RNAs interact to form the piRNA complex required for fertility in C. elegans. Mol. Cell 31, 67–78 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Ketting, R.F., Haverkamp, T.H., van Luenen, H.G. & Plasterk, R.H. Mut-7 of C. elegans, required for transposon silencing and RNA interference, is a homolog of Werner syndrome helicase and RNaseD. Cell 99, 133–141 (1999).

    CAS  PubMed  Google Scholar 

  29. Sijen, T. & Plasterk, R.H. Transposon silencing in the Caenorhabditis elegans germ line by natural RNAi. Nature 426, 310–314 (2003).

    CAS  PubMed  Google Scholar 

  30. Zhou, Z., Hartwieg, E. & Horvitz, H.R. CED-1 is a transmembrane receptor that mediates cell corpse engulfment in C. elegans. Cell 104, 43–56 (2001).

    CAS  PubMed  Google Scholar 

  31. Greiss, S., Schumacher, B., Grandien, K., Rothblatt, J. & Gartner, A. Transcriptional profiling in C. elegans suggests DNA damage dependent apoptosis as an ancient function of the p53 family. BMC Genomics 9, 334 (2008).

    PubMed  PubMed Central  Google Scholar 

  32. Gartner, A., Milstein, S., Ahmed, S., Hodgkin, J. & Hengartner, M.O. A conserved checkpoint pathway mediates DNA damage–induced apoptosis and cell cycle arrest in C. elegans. Mol. Cell 5, 435–443 (2000).

    CAS  PubMed  Google Scholar 

  33. Schumacher, B., Hofmann, K., Boulton, S. & Gartner, A. The C. elegans homolog of the p53 tumor suppressor is required for DNA damage–induced apoptosis. Curr. Biol. 11, 1722–1727 (2001).

    CAS  PubMed  Google Scholar 

  34. Gerstein, M.B. et al. Integrative analysis of the Caenorhabditis elegans genome by the modENCODE project. Science 330, 1775–1787 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Liu, T. et al. Broad chromosomal domains of histone modification patterns in C. elegans. Genome Res. 21, 227–236 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Cangiano, G. & La Volpe, A. Repetitive DNA sequences located in the terminal portion of the Caenorhabditis elegans chromosomes. Nucleic Acids Res. 21, 1133–1139 (1993).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Padeken, J., Zeller, P. & Gasser, S.M. Repeat DNA in genome organization and stability. Curr. Opin. Genet. Dev. 31, 12–19 (2015).

    CAS  PubMed  Google Scholar 

  38. Das, P.P. et al. Piwi and piRNAs act upstream of an endogenous siRNA pathway to suppress Tc3 transposon mobility in the Caenorhabditis elegans germline. Mol. Cell 31, 79–90 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Tabara, H. et al. The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell 99, 123–132 (1999).

    CAS  PubMed  Google Scholar 

  40. Wicker, T. et al. A unified classification system for eukaryotic transposable elements. Nat. Rev. Genet. 8, 973–982 (2007).

    CAS  PubMed  Google Scholar 

  41. Aguilera, A. & García-Muse, T. Causes of genome instability. Annu. Rev. Genet. 47, 1–32 (2013).

    CAS  PubMed  Google Scholar 

  42. Azvolinsky, A., Giresi, P.G., Lieb, J.D. & Zakian, V.A. Highly transcribed RNA polymerase II genes are impediments to replication fork progression in Saccharomyces cerevisiae. Mol. Cell 34, 722–734 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Deshpande, A.M. & Newlon, C.S. DNA replication fork pause sites dependent on transcription. Science 272, 1030–1033 (1996).

    CAS  PubMed  Google Scholar 

  44. Hoffman, E.A., McCulley, A., Haarer, B., Arnak, R. & Feng, W. Break-seq reveals hydroxyurea-induced chromosome fragility as a result of unscheduled conflict between DNA replication and transcription. Genome Res. 25, 402–412 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Merrikh, H., Machón, C., Grainger, W.H., Grossman, A.D. & Soultanas, P. Co-directional replication–transcription conflicts lead to replication restart. Nature 470, 554–557 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Wahba, L., Amon, J.D., Koshland, D. & Vuica-Ross, M. RNase H and multiple RNA biogenesis factors cooperate to prevent RNA:DNA hybrids from generating genome instability. Mol. Cell 44, 978–988 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Nakamori, M., Pearson, C.E. & Thornton, C.A. Bidirectional transcription stimulates expansion and contraction of expanded (CTG)*(CAG) repeats. Hum. Mol. Genet. 20, 580–588 (2011).

    CAS  PubMed  Google Scholar 

  48. Prak, E.T. & Kazazian, H.H. Jr. Mobile elements and the human genome. Nat. Rev. Genet. 1, 134–144 (2000).

    CAS  PubMed  Google Scholar 

  49. Boguslawski, S.J. et al. Characterization of monoclonal antibody to DNA.RNA and its application to immunodetection of hybrids. J. Immunol. Methods 89, 123–130 (1986).

    CAS  PubMed  Google Scholar 

  50. Castellano-Pozo, M., García-Muse, T. & Aguilera, A. The Caenorhabditis elegans THO complex is required for the mitotic cell cycle and development. PLoS One 7, e52447 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Castellano-Pozo, M. et al. R loops are linked to histone H3 S10 phosphorylation and chromatin condensation. Mol. Cell 52, 583–590 (2013).

    CAS  PubMed  Google Scholar 

  52. Balk, B. et al. Telomeric RNA–DNA hybrids affect telomere-length dynamics and senescence. Nat. Struct. Mol. Biol. 20, 1199–1205 (2013).

    CAS  PubMed  Google Scholar 

  53. Wahba, L., Costantino, L., Tan, F.J., Zimmer, A. & Koshland, D. S1-DRIP-seq identifies high expression and polyA tracts as major contributors to R-loop formation. Genes Dev. 30, 1327–1338 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Rosenkranz, H.S. & Levy, J.A. Hydroxyurea: a specific inhibitor of deoxyribonucleic acid synthesis. Biochim. Biophys. Acta 95, 181–183 (1965).

    CAS  PubMed  Google Scholar 

  55. Gan, W. et al. R-loop-mediated genomic instability is caused by impairment of replication fork progression. Genes Dev. 25, 2041–2056 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Zhang, H. & Freudenreich, C.H. An AT-rich sequence in human common fragile site FRA16D causes fork stalling and chromosome breakage in S. cerevisiae. Mol. Cell 27, 367–379 (2007).

    PubMed  PubMed Central  Google Scholar 

  57. Collins, J., Saari, B. & Anderson, P. Activation of a transposable element in the germ line but not the soma of Caenorhabditis elegans. Nature 328, 726–728 (1987).

    CAS  PubMed  Google Scholar 

  58. Pothof, J. et al. Identification of genes that protect the C. elegans genome against mutations by genome-wide RNAi. Genes Dev. 17, 443–448 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Towbin, B.D., Meister, P., Pike, B.L. & Gasser, S.M. Repetitive transgenes in C. elegans accumulate heterochromatic marks and are sequestered at the nuclear envelope in a copy-number- and lamin-dependent manner. Cold Spring Harb. Symp. Quant. Biol. 75, 555–565 (2010).

    CAS  PubMed  Google Scholar 

  60. Tijsterman, M., Pothof, J. & Plasterk, R.H.A. Frequent germline mutations and somatic repeat instability in DNA mismatch-repair-deficient Caenorhabditis elegans. Genetics 161, 651–660 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Wagner, T. & Jung, M. New lysine methyltransferase drug targets in cancer. Nat. Biotechnol. 30, 622–623 (2012).

    CAS  PubMed  Google Scholar 

  62. Helmrich, A., Ballarino, M. & Tora, L. Collisions between replication and transcription complexes cause common fragile site instability at the longest human genes. Mol. Cell 44, 966–977 (2011).

    CAS  PubMed  Google Scholar 

  63. Vermezovic, J., Stergiou, L., Hengartner, M.O. & d'Adda di Fagagna, F. Differential regulation of DNA damage response activation between somatic and germline cells in Caenorhabditis elegans. Cell Death Differ. 19, 1847–1855 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Chan, Y.A., Hieter, P. & Stirling, P.C. Mechanisms of genome instability induced by RNA-processing defects. Trends Genet. 30, 245–253 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. El Hage, A., Webb, S., Kerr, A. & Tollervey, D. Genome-wide distribution of RNA–DNA hybrids identifies RNase H targets in tRNA genes, retrotransposons and mitochondria. PLoS Genet. 10, e1004716 (2014).

    PubMed  PubMed Central  Google Scholar 

  66. Kilchert, C., Wittmann, S. & Vasiljeva, L. The regulation and functions of the nuclear RNA exosome complex. Nat. Rev. Mol. Cell Biol. 17, 227–239 (2016).

    CAS  PubMed  Google Scholar 

  67. Skourti-Stathaki, K., Proudfoot, N.J. & Gromak, N. Human senataxin resolves RNA/DNA hybrids formed at transcriptional pause sites to promote Xrn2-dependent termination. Mol. Cell 42, 794–805 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Kim, N. & Jinks-Robertson, S. Transcription as a source of genome instability. Nat. Rev. Genet. 13, 204–214 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Kouzine, F., Levens, D. & Baranello, L. DNA topology and transcription. Nucleus 5, 195–202 (2014).

    PubMed  PubMed Central  Google Scholar 

  70. Yuan, Y. et al. Gossypol and an HMT G9a inhibitor act in synergy to induce cell death in pancreatic cancer cells. Cell Death Dis. 4, e690 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Liu, S. et al. Setdb1 is required for germline development and silencing of H3K9me3-marked endogenous retroviruses in primordial germ cells. Genes Dev. 28, 2041–2055 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Meister, P., Towbin, B.D., Pike, B.L., Ponti, A. & Gasser, S.M. The spatial dynamics of tissue-specific promoters during C. elegans development. Genes Dev. 24, 766–782 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. 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).

    CAS  PubMed  Google Scholar 

  74. Ikegami, K., Egelhofer, T.A., Strome, S. & Lieb, J.D. Caenorhabditis elegans chromosome arms are anchored to the nuclear membrane via discontinuous association with LEM-2. Genome Biol. 11, R120 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Gaidatzis, D., Lerch, A., Hahne, F. & Stadler, M.B. QuasR: quantification and annotation of short reads in R. Bioinformatics 31, 1130–1132 (2015).

    CAS  PubMed  Google Scholar 

  76. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S.L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    PubMed  PubMed Central  Google Scholar 

  77. Jurka, J. et al. Repbase Update, a database of eukaryotic repetitive elements. Cytogenet. Genome Res. 110, 462–467 (2005).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  80. Ye, K., Schulz, M.H., Long, Q., Apweiler, R. & Ning, Z. Pindel: a pattern growth approach to detect break points of large deletions and medium sized insertions from paired-end short reads. Bioinformatics 25, 2865–2871 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Craig, A.L., Moser, S.C., Bailly, A.P. & Gartner, A. Methods for studying the DNA damage response in the Caenorhabdatis elegans germ line. Methods Cell Biol. 107, 321–352 (2012).

    CAS  PubMed  Google Scholar 

  82. Ginno, P.A., Lott, P.L., Christensen, H.C., Korf, I. & Chédin, F. R-loop formation is a distinctive characteristic of unmethylated human CpG island promoters. Mol. Cell 45, 814–825 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

A number of strains were provided by the Caenorhabditis Genetics Center (CGC), which is funded by NIH Office of Research Infrastructure Programs (P40 OD010440). We thank R. Ciosk and P. Pasero for reagents and materials, I. Katiç and members of the Friedrich Miescher Institute Genomics and Microscopy facilities for advice and discussion, and P. Ginno and L. Constantino for advice on R-loop detection. J.P. is supported by a long-term EMBO fellowship. S.M.G. thanks the Swiss National Science Foundation as well as the Novartis Research Foundation for support.

Author information

Authors and Affiliations

Authors

Contributions

P.Z. and J.P. planned and executed most experiments, evaluated results and wrote the paper; S.M.G. planned experiments, evaluated results and wrote the paper; R.v.S. and M.T. helped with evaluation of genome mutations and provided the LacZ mutagenesis assay; and V.K. provided invaluable technical help.

Corresponding author

Correspondence to Susan M Gasser.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Immunofluorescence (IF) confirms absence of H3K9me in met-2 set-25 worms.

IF images of wild-type (wt) and met-2 set-25 worms showing the loss of H3K9me2/me3 at the indicated developmental stages grown at 20 °C (Online Methods). The staining of H4 pan-acetyl (H4ac) served as a positive control. (a) Gonads (bar, 20 μm). (b) Embryos and L2-stage larvae (bar, 20 μm).

Supplementary Figure 2 Phenotypes of strains losing H3K9me in the germ line.

(a) Number of viable progeny of wt and met-2 set-25 per worm at 26 °C, over three generations, starting at generation 3 after transfer of the worms from 20 °C (number of independent experiments (N) = 3, number of worms counted per experiment (n) = 60). (b) Number of viable progeny of wt, met-2 set-25 and mutants of the PIWI pathway per worm at 15 °C, 20 °C and 25 °C by generation 3 after transfer from 20 °C to the indicated temperature (N = 1, n = 25). (c) Percentage of worms developing full gonad arms at 20 °C. mex-5:gfp-h2b was used to visualize gonad cells at all stages (N = 7, n = 2). (d) Analysis of RNA-seq data showing average fold change in expression of apoptosis response genes in the gonads of met-2 set-25 mutant versus wt (N = 3, P < 0.05 adjusted for multiple testing; FDR < 0.05). (e) Percentage of laid eggs hatching at 20 °C. Strains deficient for CEP-1 (p53) additionally expressed the CED-1::GFP apoptosis reporter (N = 2, n = 80). (f) Analysis of RAD-51 foci detected by IF in the mitotic zone of the indicated genotypes to quantify the presence of resected DNA double-strand breaks. The percentage in images is equal to the frequency of mitotic tip cells with detectable RAD-51 foci. The bar graph further segregates positive cells by number of foci per cell (N = 3, n = 40 gonads).

Supplementary Figure 3 H3K9me distribution on genes and repeats.

(a) H3K9me2 and H3K9me3 ChIP-seq were performed on early embryos at 20 °C (N = 2) in a wild-type (N2) strain. Distribution along chr. I in relation to repetitive elements (REs) is shown. Quantification to the right shows the ratio of the coverage of the indicated histone modification on chromosome arms (outer two-thirds) versus the center (inner one-third). H3K27me3 and H3K4me3 mapping data are from the modENCODE project based on ChIP performed on mixed-population embryos, 20 °C. (b) Metaplot and heat map showing log2 enrichment of H3K9me2 and H3K9me3 (IP versus input) over gene bodies. Each row represents one gene, displaying the binned coding region plus 1 kb upstream of the transcription start site (TSS) and 1 kb downstream of the transcription termination sites (TES). Blue is most enriched, red is least enriched. (c) High-density scatterplot showing H3K9me2 and me3 enrichment over muscle-specific genes and pseudogenes. The upper number indicates the percentage of genes that are H3K9me3 positive (including K9me2 positive and negative), and the lower number indicates the percentage of genes that are H3K9me2 positive but H3K9me3 negative (d,e) Distribution of H3K9me2 and H3K9me3 determined by ChIP-seq over the gene body of a gene with tissue-spec expression (d) neuronal unc-54, and a cluster of pseudogenes (e).

Supplementary Figure 4 Gene and repeat element derepression in the absence of H3K9me.

(a,b) H3K9me2 and H3K9me3 ChIP-seq was performed on early embryos at 20 °C (N = 2) in a wt strain, and gene expression was determined by two replicas of RNA-seq from embryos (20 °C and 25 °C) and from isolated gonads (20 °C) from wt and met-2 set-25 strains. Scatterplots show H3K9me2 and H3K9me3 enrichment over the genes that are derepressed in early embryos at 20 °C and 25 °C, and in the gonads of met-2 set-25 animals, versus wt. Number indicates the percentage of derepressed genes enriched for either H3K9me2 or H3K9me3 in wt embryos. (b) Scatterplot of H3K9me2 and H3K9me3 enrichment over repeat subfamilies. Red dots mark repeat subfamilies that are derepressed in the indicated tissue and conditions; black dots represent non-affected repeat subfamilies. The red number indicates the percentage of derepressed repeat subfamilies that are either H3K9me2 or H3K9me3 positive. (c) qPCR verification of a subset of REs that were detected as derepressed (>2 fold, met-2 set-25/wt) in met-2 set-25 embryos at 20 °C. Expression of the same REs was additionally analyzed in gonads and L1-stage larval RNA. REs of all three main classes are detected (N = 3), but clearly there are strong stage-specific expression differences, with larvae and embryos showing more similarity.

Supplementary Figure 5 R-loop accumulation on repeat elements (REs).

(a) RNA:DNA hybrids were detected on dot plots of genomic DNA isolated from the indicated genotypes grown at 20 °C or 25 °C (three blots at each temperature were quantified by scanning for Fig. 5a). Equal amounts of DNA (determined by OD260/280) extracted from adult worms were spotted with or without RNase H treatment in decreasing concentrations (4, 2 and 1 μg). The nitrocellulose membrane was probed with the RNA:DNA-specific S9.6 antibody (n(20 °C) = 3, n(25 °C) = 3). Quantification on the right side with signals normalized to the background of each blot. (b) DRIP-seq signals in wt embryos are shown for genes grouped on the basis of their transcriptional activity. The upper box blot shows the level of transcription of the separate groups (N = 1). This enhancement of R loops on very highly expressed genes has been observed in many organisms. (c,d) Graphs show the accumulation of RNA:DNA hybrids in wt and met-2 set-25 embryos relative to the distribution of H3K9me2 and me3 over the rDNA cluster (c) or the right telomere of chr. I (d), in wt embryos. (e) DRIP-seq examples showing the R-loop signal over two RE clusters. The ChIP signal from antibody S9.6, which is specific for RNA:DNA hybrids, was normalized to input, and the RNase H control values were subtracted.

Supplementary Figure 6 Germline mutations in met-2 set-25 worms.

Further verification and characterization of the mutations detected in the genome sequencing experiment described in Figure 6. (a) Number and distribution of single-nucleotide variants (SNVs) or polymorphisms observed in the sequencing experiment described in Figure 6a. No dinucleotide preferences were found among SNVs. (b) Sketch of a complex rearrangement involving a Tc3 transposon, found exclusively in the met-2 set-25 genome. The rearrangement was identified by genome sequencing. The graph to the left indicates the precise site of Tc3 insertion. Below are H3K9me2 and H3K9me3 ChIP-seq tracks for the region around the Tc3 transposon that provoked the inversion: it bears high levels of H3K9me3 and showed roughly a twofold change in expression in met-2 set-25 over wt gonads at 20 °C. (c) Southern blotting with a probe against Tc3 shows a novel band detected in the met-2 set-25 mutant. (d) Depiction of the rearrangement shown in b indicating the position of the primer pairs used to verify the rearranged genomic context by PCR. PCR confirmation of the rearrangement is shown to the right. (e) Copy number ratios of met-2 set-25 worms relative to wt for the entire telomeric repeat subfamily, and for single telomere repeats. (f) Copy number ratios of met-2 set-25 worms relative to wt for rDNA repeats.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–6 and Supplementary Tables 1–3. (PDF 1797 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zeller, P., Padeken, J., van Schendel, R. et al. Histone H3K9 methylation is dispensable for Caenorhabditis elegans development but suppresses RNA:DNA hybrid-associated repeat instability. Nat Genet 48, 1385–1395 (2016). https://doi.org/10.1038/ng.3672

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3672

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