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:

The disordered N-terminal domain of DNMT3A recognizes H2AK119ub and is required for postnatal development

Abstract

DNA methyltransferase 3a (DNMT3A) plays a crucial role during mammalian development. Two isoforms of DNMT3A are differentially expressed from stem cells to somatic tissues, but their individual functions remain largely uncharacterized. Here we report that the long isoform DNMT3A1, but not the short DNMT3A2, is essential for mouse postnatal development. DNMT3A1 binds to and regulates bivalent neurodevelopmental genes in the brain. Strikingly, Dnmt3a1 knockout perinatal lethality could be partially rescued by DNMT3A1 restoration in the nervous system. We further show that the intrinsically disordered N terminus of DNMT3A1 is required for normal development and DNA methylation at DNMT3A1-enriched regions. Mechanistically, a ubiquitin-interacting motif embedded in a putative α-helix within the N terminus binds to mono-ubiquitinated histone H2AK119, probably mediating recruitment of DNMT3A1 to Polycomb-regulated regions. These data demonstrate an isoform-specific role for DNMT3A1 in mouse postnatal development and reveal the N terminus as a necessary regulatory domain for DNMT3A1 chromatin occupancy and functions in the nervous system.

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: DNMT3A1, but not DNMT3A2, is essential for mouse postnatal development.
Fig. 2: DNMT3A1 binds to and regulates neurodevelopmental genes in the cerebral cortex.
Fig. 3: DNMT3A1 restoration in the nervous system by Nestin-Cre partially rescued Dnmt3a1 KO lethality.
Fig. 4: Deletion of DNMT3A1 N terminus leads to impaired postnatal development.
Fig. 5: The N terminus is required for DNMT3A1-regulated DNA methylation in the cerebral cortex.
Fig. 6: Integrative analyses of DNA methylation and gene expression changes in Dnmt3a1–/– neuron nuclei.
Fig. 7: The N terminus facilitates DNMT3A1 enrichment around bivalent promoters by binding to H2AK119ub.

Similar content being viewed by others

Data availability

All new data sets generated for this work have been deposited to GEO under accession number GSE164265. Gene ontology-based mouse gene set collections for GSEA were downloaded from GO2MSIG database (http://www.bioinformatics.org/go2msig/, April 2015 release). The genomic coordinates for mouse transcripts and CpG islands were downloaded from GENCODE (https://www.gencodegenes.org/mouse/) and UCSC genome browser (http://hgdownload.cse.ucsc.edu/) respectively. Source data are provided with this paper.

Code availability

Custom scripts developed for this study are available at https://github.com/DapengHao/DNMT3A1N_terminus.

References

  1. Smith, Z. D. & Meissner, A. DNA methylation: roles in mammalian development. Nat. Rev. Genet. 14, 204–220 (2013).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Greenberg, M. V. C. & Bourc’his, D. The diverse roles of DNA methylation in mammalian development and disease. Nat. Rev. Mol. Cell Biol. 20, 590–607 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Okano, M., Bell, D. W., Haber, D. A. & Li, E. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99, 247–257 (1999).

    Article  CAS  PubMed  Google Scholar 

  5. Nguyen, S., Meletis, K., Fu, D., Jhaveri, S. & Jaenisch, R. Ablation of de novo DNA methyltransferase Dnmt3a in the nervous system leads to neuromuscular defects and shortened lifespan. Dev. Dyn. 236, 1663–1676 (2007).

    Article  CAS  PubMed  Google Scholar 

  6. Lavery, L. A. et al. Losing dnmt3a dependent methylation in inhibitory neurons impairs neural function by a mechanism impacting Rett syndrome. eLife 9, 1–27 (2020).

    Article  Google Scholar 

  7. Christian, D. et al. DNMT3A haploinsufficiency results in behavioral deficits and global epigenomic dysregulation shared across neurodevelopment disorders. Cell Rep. 33, 108416 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Ley, T. J. et al. DNMT3A mutations in acute myeloid leukemia. N. Engl. J. Med. 363, 2424–2433 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Yang, L., Rau, R. & Goodell, M. A. DNMT3A in haematological malignancies. Nat. Rev. Cancer 15, 152–165 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Tatton-Brown, K. et al. Mutations in the DNA methyltransferase gene DNMT3A cause an overgrowth syndrome with intellectual disability. Nat. Genet. 46, 385–388 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Heyn, P. et al. Gain-of-function DNMT3A mutations cause microcephalic dwarfism and hypermethylation of Polycomb-regulated regions. Nat. Genet. 51, 96–105 (2019).

    Article  CAS  PubMed  Google Scholar 

  12. Otani, J. et al. Structural basis for recognition of H3K4 methylation status by the DNA methyltransferase 3A ATRX-DNMT3-DNMT3L domain. EMBO Rep. 10, 1235–1241 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Guo, X. et al. Structural insight into autoinhibition and histone H3-induced activation of DNMT3A. Nature 517, 640–644 (2015).

    Article  CAS  PubMed  Google Scholar 

  14. Dhayalan, A. et al. The Dnmt3a PWWP domain reads histone 3 lysine 36 trimethylation and guides DNA methylation. J. Biol. Chem. 285, 26114–26120 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Baubec, T. et al. Genomic profiling of DNA methyltransferases reveals a role for DNMT3B in genic methylation. Nature 520, 243–247 (2015).

    Article  CAS  PubMed  Google Scholar 

  16. Weinberg, D. N. et al. The histone mark H3K36me2 recruits DNMT3A and shapes the intergenic DNA methylation landscape. Nature 573, 281–286 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Chen, T., Ueda, Y., Xie, S. & Li, E. A novel Dnmt3a isoform produced from an alternative promoter localizes to euchromatin and its expression correlates with active de novo methylation. J. Biol. Chem. 277, 38746–38754 (2002).

    Article  CAS  PubMed  Google Scholar 

  18. Chen, T., Ueda, Y., Dodge, J. E., Wang, Z. & Li, E. Establishment and maintenance of genomic methylation patterns in mouse embryonic stem cells by Dnmt3a and Dnmt3b. Mol. Cell. Biol. 23, 5594–5605 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Feng, J., Chang, H., Li, E. & Fan, G. Dynamic expression of de novo DNA methyltransferases Dnmt3a and Dnmt3b in the central nervous system. J. Neurosci. Res. 79, 734–746 (2005).

    Article  CAS  PubMed  Google Scholar 

  20. Wu, H. et al. Dnmt3a-dependent nonpromoter DNA methylation facilitates transcription of neurogenic genes. Science 329, 444–447 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Gu, T. et al. DNMT3A and TET1 cooperate to regulate promoter epigenetic landscapes in mouse embryonic stem cells. Genome Biol. 19, 88 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  22. Manzo, M. et al. Isoform‐specific localization of DNMT3A regulates DNA methylation fidelity at bivalent CpG islands. EMBO J. 36, 3421–3434 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. 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 

  24. Radivojac, P. et al. Intrinsic disorder and functional proteomics. Biophys. J. 92, 1439–1456 (2007).

    Article  CAS  PubMed  Google Scholar 

  25. Shin, Y. et al. Spatiotemporal control of intracellular phase transitions using light-activated optodroplets. Cell 168, 159–171.e14 (2017).

    Article  CAS  PubMed  Google Scholar 

  26. Li, C. H. et al. MeCP2 links heterochromatin condensates and neurodevelopmental disease. Nature 586, 440–444 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jeong, M. et al. Large conserved domains of low DNA methylation maintained by Dnmt3a. Nat. Genet. 46, 17–23 (2014).

    Article  CAS  PubMed  Google Scholar 

  28. Stroud, H. et al. Early-life gene expression in neurons modulates lasting epigenetic states. Cell 171, 1151–1164.e16 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Li, J. et al. Dnmt3a knockout in excitatory neurons impairs postnatal synapse maturation and is partly compensated by repressive histone modification H3K27me3. Preprint at bioRxiv https://doi.org/10.1101/2019.12.20.883694 (2021).

  30. Sendžikaitė, G., Hanna, C. W., Stewart-Morgan, K. R., Ivanova, E. & Kelsey, G. A DNMT3A PWWP mutation leads to methylation of bivalent chromatin and growth retardation in mice. Nat. Commun. 10, 1884 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Zhang, X. et al. Large DNA methylation nadirs anchor chromatin loops maintaining hematopoietic stem cell identity. Mol. Cell 78, 506–521 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Endoh, M. et al. Polycomb group proteins Ring1A/B are functionally linked to the core transcriptional regulatory circuitry to maintain ES cell identity. Development 135, 1513–1524 (2008).

    Article  CAS  PubMed  Google Scholar 

  33. Weinberg, D. N. et al. Two competing mechanisms of DNMT3A recruitment regulate the dynamics of de novo DNA methylation at PRC1-targeted CpG islands. Nat. Genet. 53, 794–800 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Viré, E. et al. The Polycomb group protein EZH2 directly controls DNA methylation. Nature 439, 871–874 (2006).

    Article  PubMed  CAS  Google Scholar 

  35. Neri, F. et al. Dnmt3L antagonizes DNA methylation at bivalent promoters and favors DNA methylation at gene bodies in ESCs. Cell 155, 121–134 (2013).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Brunetti, L., Gundry, M. C. & Goodell, M. A. DNMT3A in leukemia. Cold Spring Harb. Perspect. Med. 7, a030320 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Nishiyama, A. et al. Uhrf1-dependent H3K23 ubiquitylation couples maintenance DNA methylation and replication. Nature 502, 249–253 (2013).

    Article  CAS  PubMed  Google Scholar 

  39. Qin, W. et al. DNA methylation requires a DNMT1 ubiquitin interacting motif (UIM) and histone ubiquitination. Cell Res. 25, 911–929 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Fuks, F., Burgers, W. A., Godin, N., Kasai, M. & Kouzarides, T. Dnmt3a binds deacetylases and is recruited by a sequence-specific repressor to silence transcription. EMBO J. 20, 2536–2544 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Hata, K., Okano, M., Lei, H. & Li, E. Dnmt3L cooperates with the Dnmt3 family of de novo DNA methyltransferases to establish maternal imprints in mice. Development 129, 1983–1993 (2002).

    Article  CAS  PubMed  Google Scholar 

  42. Li, J.-Y. et al. Synergistic function of DNA methyltransferases Dnmt3a and Dnmt3b in the methylation of Oct4 and Nanog. Mol. Cell. Biol. 27, 8748–8759 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Chang, Y. et al. MPP8 mediates the interactions between DNA methyltransferase Dnmt3a and H3K9 methyltransferase GLP/G9a. Nat. Commun. 2, 533 (2011).

    Article  PubMed  CAS  Google Scholar 

  44. Palamarchuk, A. et al. Tcl1 protein functions as an inhibitor of de novo DNA methylation in B-cell chronic lymphocytic leukemia (CLL). Proc. Natl Acad. Sci. USA 109, 2555–2560 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Alberti, S. et al. A user’s guide for phase separation assays with purified proteins. J. Mol. Biol. 430, 4806–4820 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  47. Anders, S., Pyl, P. T. & Huber, W. HTSeq – A Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    Article  CAS  PubMed  Google Scholar 

  48. 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 

  49. 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 

  50. 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  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  53. Jühling, F. et al. metilene: Fast and sensitive calling of differentially methylated regions from bisulfite sequencing data. Genome Res. 26, 256–262 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. 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 

Download references

Acknowledgements

We thank the Genetically Engineered Rodent Models Core at the Baylor College of Medicine (BCM) for help with generation of mouse models and the Cytometry and Cell Sorting Core Facility at BCM for help with flow cytometry and sorting. We thank the Protein and Monoclonal Antibody Production Core and Y. Zhu at BCM for help with recombinant protein purification. We also thank X. Cheng, X. Zhang and M. T. Bedford (The University of Texas MD Anderson Cancer Center) for helpful discussions. This work was supported by National Institutes of Health grants: R01HL134780 and R01HL146852 (to Y.H.), R01HG007538 and R01CA228140 (to W.L.), NS071153 (to B.D.), R01GM112003 (to Y.Z.), R01CA183252 and R01DK092883 (to M.A.G.) and P30CA125123. This work was also supported by American Cancer Society (RSG-18-043-01-LIB, to Y.H.) and Welch Foundation (BE-1913-20190330, to Y.Z.).

Author information

Authors and Affiliations

Authors

Contributions

M.A.G. and T.G. conceived the study. T.G. performed most experiments with help from M.J., T.-W.H., A.G.G., A.T. and C.R. D.H. carried out computational analysis of all the high-throughput sequencing data and X.L. carried out initial analysis of DNMT3A ChIP–seq data under supervision of W.L. J.W. performed the LTP experiment under supervision of B.D. L.G. performed the OptoIDR assay under supervision of Y.Z. and Y.H. M.A.G., T.G. and D.H. designed the figures. M.A.G. and T.G. wrote the manuscript with input from the other authors. All authors approved the final version of the manuscript.

Corresponding author

Correspondence to Margaret A. Goodell.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Genetics thanks Déborah Bourc’his, Wolf Reik, Aled Parry and the other, anonymous, reviewer(s) 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 Expression profiles of DNMT3A isoforms in mouse postnatal tissues and unaltered cell populations in the spleen and thymus of Dnmt3a1–/– mice.

a, RT-qPCR expression analysis of Dnmt3a and its transcript variants in WT mouse tissues at P18. n = 2 biological independent mice. Data are presented as mean with individual values. b, A diagram representing the generation of Dnmt3a-FLAG knock-in (KI) mouse model. 3× FLAG tag was added to the C terminus of the endogenous Dnmt3a by CRISPR-Cas9 mediated homologous recombination. Linker amino acid sequences: GGSG. c, Western blots for DNMT3A isoforms within P18 Dnmt3a-FLAG mouse tissues. The left panel shows specific detection of FLAG-tagged DNMT3A isoforms in the KI mouse thymus. The experiment was repeated two times independently with similar results. d, Alignment of the sequence of residual N-terminal fragment potentially left in cells after exon 4 deletion to that of DNMT3A1. The reading frame shifts after Proline 56, and the premature stop codon occurs 11 amino acids afterwards. e, Western blots to verify the loss of DNMT3A1 or DNMT3A2 protein respectively in P18 knockout mouse thymuses and cortices. Samples of each genotype were biological replicates. f, Quantification of DNMT3A1 expression in Western blot analysis of P18 mouse cortices (e, right). n = 2 biological independent samples. Data are shown as mean with individual values. g,h, Smaller spleens in P21 Dnmt3a1–/– mice. Representative images of KO and WT spleens are shown (g). Spleen weight was normalized to the body weight (h). Data are shown as mean ± s.e.m. ** P = 0.0014; ns: not significant, P = 0.9707 (two-sided unpaired t tests). i,j, Smaller thymuses in P21 Dnmt3a1–/– mice. Data are shown as mean ± s.e.m. (j). * P = 0.0142; ns: not significant, P = 0.4033 (two-sided unpaired t tests). k, Flow cytometry analysis of cell populations on viable splenic cells from a WT mouse: Gr1/Mac-1 for myeloid cells, CD4/CD8 for T cells and B220 for B cells. The sequential gating strategy is displayed in Supplementary Fig. 1. l, Quantification of frequencies for cell subsets as shown in k. Data are presented as mean ± s.d. m, A representative flow cytometry plot of CD4 and CD8 on viable thymocytes from a WT mouse. SP: single positive; DP: double positive, DN: double negative. The sequential gating strategy is provided in Supplementary Fig. 2. n, Quantification of frequencies for cell subsets as shown in m. Data are presented as mean ± s.d. The number of mice analyzed in h,l and j,n: WT littermates for Dnmt3a1 KO, n = 4; Dnmt3a1 KO, n = 5; WT littermates for Dnmt3a2 KO, n = 4; Dnmt3a2 KO, n = 5.

Source data

Extended Data Fig. 2 General morphology and cell populations in Dnmt3a1–/– mouse brain.

ac, Double immunofluorescence (IF) staining for DNMT3A1 and the neuronal marker NEUN (a), astroglial marker SOX9 (b), or oligodendroglial marker OLIG2 (c) on sagittal brain sections of P23 WT mice. White arrows point to cells positive for SOX9 or OLIG2. IF was performed with an N-terminal anti-DNMT3A antibody which is DNMT3A1-specific. Section thickness: 15 μm. Scale bar = 100 μm. d, Quantification of DNMT3A1 signal intensity in DNMT3A1 positive and NEUN, SOX9, or OLIG2 positive cells in the same IF images. Positive cells were detected and measured using QuPath software. e, Dnmt3a1–/– mice had smaller brains, but the ratio of brain to body weight was greater than WT littermates at P21. WT, n = 8; Dnmt3a1 KO, n = 7. All data are presented as mean ± s.d. with individual values. P values for brain weight, *** P = 0.0005; body weight, **** P < 0.0001 and brain/body ratio, **** P < 0.0001 (two-sided unpaired t tests). f, Normal brain morphology of Dnmt3a1–/– mouse revealed by Nissl staining of sagittal brain sections. Scale bar = 1.0 mm. The experiment was repeated two times with independent biological samples. g, IF staining showing deficiency of DNMT3A1 protein and unaltered cell populations of neurons (NEUN+), astrocytes (SOX9+) and oligodendrocytes (OLIG2+) in the cerebral cortex of Dnmt3a1–/– mice at P23. Scale bar = 100 μm. The experiment was repeated three times and cell counts are provided below. h, Schematic representation of a sagittal section of the mouse brain. The black squares indicate regions selected for cell counting. i, Percentages of neurons, astrocytes, and oligodendrocytes in the indicated brain regions (h) of Dnmt3a1–/– (n = 3) and WT (n = 3) mice. Two sections from each mouse brain were stained for each cell marker. All data are presented as mean ± s.d. j, Representative IF staining for NEUN and the motoneuron-specific marker ChAT on coronal sections through the hypoglossal nucleus of P23 mouse brain. Motor neurons are positive for both ChAT and NEUN. Scale bar = 100 μm. The experiment was repeated two times. k, Total numbers of motor neurons in the hypoglossal nucleus of Dnmt3a1–/– (n = 3) and WT (n = 2) mice. Data are presented as mean with individual values.

Source data

Extended Data Fig. 3 Correlation between DNMT3A1 and H3K27me3 across gene bodies and gene expression analysis for Dnmt3a1–/– mouse brain.

a, Heatmaps of H3K4me3, H3K27me3 and DNMT3A1 across genes (left) in P18 Dnmt3a-FLAG cortex. Genes were ranked by H3K4me3 occupancy, and the curves represent smoothing cubic splines (right) fitted by the median signals of H3K27me3 and DNMT3A1 across each gene. b, Unchanged expression levels of DNA methylation regulators in Dnmt3a1–/– cortex. TPM: transcripts per million. WT, n = 3; Dnmt3a1 KO, n = 3 biological replicates. Data are shown as mean ± s.d. *** P = 0.0002 (two-sided unpaired t test). c, RT-qPCR verification of depletion of Dnmt3a1 full-length transcripts and down-regulation of neural development-related genes in P21 Dnmt3a1–/– cerebral cortex. The expression level of each gene was normalized to that of Gapdh. WT, n = 4; Dnmt3a1 KO, n = 5 biological replicates. Data are shown as mean with s.d. P values for Dnmt3a1, **** P < 0.0001; Neurod6, * P = 0.0173; Bdnf, * P = 0.0280; Wnt7a, *** P = 0.0002 (two-sided unpaired t tests). d,e, Heatmaps highlighting the relative expression level of members in gene sets synapse assembly (d) and neuromuscular process (e). f, RT-qPCR verification of expression changes for selected members in gene sets synapse assembly and neuromuscular process in P21 Dnmt3a1–/– cerebral cortex. WT, n = 4; Dnmt3a1 KO, n = 5 biological replicates. Data are shown as mean ± s.d. P values for Wnt5a, ** P = 0.0081; Homer1, * P = 0.0215; Shank1, *** P = 0.0006; Shank2, ** P = 0.0053 (two-sided unpaired t tests). g, GSEA of Dnmt3a1–/– hippocampal transcriptome. Top dysregulated gene sets were listed and ranked by NES. h, Enrichment plots for down-regulated gene set regulation of synaptic transmission in Dnmt3a1–/– hippocampus. i, Heatmap and average density of wildtype DNMT3A1 binding across DEGs (up- or down-DEGs) and non-DEGs (other) of Dnmt3a1–/– cortex. Input was subtracted from the signal of DNMT3A1 ChIP-seq. Genes are ordered by P values, with up-DEGs and down-DEGs on the top and bottom ends, respectively.

Source data

Extended Data Fig. 4 DNMT3A1 N-terminal domain is an intrinsically disordered region.

a, Prediction of intrinsic disorder for DNMT3A1 by PONDR (Predictor of Natural Disordered Regions) online. Amino acid positions are shown on the x axis. The cyan bars designate the N-terminal regions investigated. b, Fluorescence images of living Hela cells expressing Cry2-mCherry or N-terminal IDR fusion proteins (optoN219 and optoN278). All cells were subjected to 488 nm blue light stimulation under identical conditions. The experiment was repeated three times independently with similar results. Representative images after 80-second stimulation are presented. Scale bar = 10 μm. c, Purification of 6×His tagged GFP and N-terminal IDR-GFP fusion proteins for in vitro droplet formation assay. d, Visualization of turbidity of indicated protein solutions (20 µM) in droplet formation buffer in the absence (–) or presence (+) of 8% PEG-8000. e, Representative images of GFP-positive spherical protein droplets formed at concentrations of 5 µM and 20 µM. Proteins were diluted to the final concentrations with droplet formation buffer in the presence of 8% PEG-8000. The experiment was repeated four times with similar results. Scale bar = 20 μm. f, Fusion events between proximal droplets of N219-GFP (top) or N278-GFP (bottom). g, Live-cell images of FRAP analysis on GFP-DNMT3A1 expressed in NIH 3T3 cells. Scale bar = 10 μm. h, Average fluorescence recovery trace in GFP-DNMT3A1 FRAP experiments (n = 12 cells). All data are presented as mean ± s.d. i, Live-cell images of FRAP analysis on GFP-MeCP2. Scale bar = 10 μm. j, Average fluorescence recovery trace in GFP-MeCP2 FRAP experiments (n = 6 cells).

Source data

Extended Data Fig. 5 Dnmt3a1ΔN mice showed a reduced rate of weight gain and impaired behaviors.

a, Body weights of female Dnmt3a1ΔN (n = 5) mice and WT littermates (n = 5) at the age of 2, 4 and 6 months. Data are shown as mean ± s.d. ** P = 0.003166; *** P = 0.000360; **** P < 0.0001 (multiple t tests). b, Representative images of Dnmt3a1ΔN and WT females at 6 months. c, Dnmt3a1ΔN mice exhibited an increase in fecal boli deposits during 30-min open field test. **** P < 0.0001 (two-sided unpaired t tests). d, Dnmt3a1ΔN mice showed a similar level of anxiety with WT littermates in a Light-dark exploration test. ns: not significant (two-sided unpaired t tests), dark duration, P = 0.5814; light duration, P = 0.5810; entries, P = 0.4214. The number of mice analyzed in c,d: Dnmt3a1ΔN, n = 19; WT, n = 16. Data are shown as mean ± s.e.m. with individual values.

Source data

Extended Data Fig. 6 DNA methylation and gene expression analyses for Dnmt3a1–/–, Dnmt3a2–/– and Dnmt3a1ΔN P21 cortices.

a, Average levels, and changes of CpG methylation genome-wide (overall), and at various genomic features in the cerebral cortex of WT, Dnmt3a1–/–, Dnmt3a2–/– and Dnmt3a1ΔN P21 mice. Promoters: 2,000 bp upstream to 500 bp downstream of TSSs. Canyons were defined as long unmethylated regions over 3.5 kb with an average methylation level < 0.1 in WT cortices (n = 855). CGI or Canyon shores: ± 2 kb regions flanking CGI or Canyons. Data are shown as mean with individual values. b, Genome-wide average CpH (non-CpG) methylation levels in WT and KO cerebral cortices (n = 2 biological replicates each genotype). Data are shown as mean with individual values. c, DNA methylation levels across bivalent genes and two other clusters of genes defined in Fig. 2a. d, Scatterplot of DMR methylation levels in KOs and WT. Each dot represents a DMR. Regions with methylation difference less than 0.1 are not included, giving rise to the gaps. e, Smooth curves of the percentage of each gene covered by DMRs in KO cortices. Genes were ordered by wildtype DNMT3A signal (Fig. 5h). f, Venn diagram of overlapping DMRs (FDR < 0.05) between KO cortices. g, Venn diagram representations of overlapping up-regulated DEGs (left, P < 0.01) and down-regulated DEGs (right) between KOs. h, GSEA analysis of KO cortex transcriptomes. GO terms (biological processes) with FDR < 0.05 in any of the differential expression analysis (Dnmt3a1 KO versus WT, Dnmt3a2 KO versus WT, or Dnmt3a1ΔN versus WT) were included (n = 49).

Source data

Extended Data Fig. 7 Neuron nuclei sorting from the cerebral cortex.

a, Flow cytometry analysis of purified cortical cell nuclei by sucrose ultracentrifugation. NEUN-positive single nuclei were sorted and used for DNA/RNA extraction and ChIP experiments. b, Representative IF images of presort and sorted nuclei. The purity of sorted neuron nuclei was above 95%, and it was checked for every sample after sorting. Scale bar = 50 μm.

Extended Data Fig. 8 DNA methylation and gene expression analyses in sorted neuron nuclei.

a, Violin plots for the distribution of average CpG methylation ratios of 5 kb bins (top) and CpH methylation ratio of 50 kb bins (bottom) over the genome in WT, Dnmt3a1–/–, Dnmt3a2–/– and Dnmt3a1ΔN neuron nuclei (n = 2 biological replicates each genotype). The lower and upper hinges of boxplots correspond to the first and third quartiles. The lower or upper whisker extends from the hinge to the smallest or largest value within 1.5× interquartile range of the hinge respectively. b, DNA methylation levels at Canyons (left), CpG islands (right) and flanking regions. c, IGV displays of DNMT3A1 (cortex), H3K4me3, H3K27me3 and H2AK119ub enrichment in Dnmt3a-FLAG neuron nuclei, and CpG methylation in WT and KO neuron nuclei at Bmp7 gene locus. The differentially methylated region is highlighted in light green. d, Heatmaps for relative CpG methylation levels at Dnmt3a1–/– DMRs (FDR < 0.05) in WT, Dnmt3a1–/–, Dnmt3a2–/– and Dnmt3a1ΔN neuron nuclei. Overlapping DMRs in Dnmt3a2–/– or Dnmt3a1ΔN neuron genome are displayed on the right. e, Identification of enhancers (poised and active) by histone modifications H3K4me1, H3K4me3 and H3K27ac. H3K27me3 and DNMT3A1 were plotted accordingly. f, Average DNA methylation levels at poised and active enhancers in in WT, Dnmt3a1–/–, Dnmt3a2–/– and Dnmt3a1ΔN neuron nuclei. g, Volcano plot of the distribution of differentially expressed genes (P < 0.01) in Dnmt3a1–/– neuron nuclei. h, Heatmap for wildtype DNMT3A1 binding across Dnmt3a1–/–-neuron nuclei DEGs (up- or down-DEGs) and non-DEGs (other). Density plot was presented in Fig. 6c. i, Average DNA methylation levels across Dnmt3a1–/–-neuron nuclei DEGs and other genes in WT and Dnmt3a1–/– neuron nuclei. j, Heatmaps of DMR distribution across up- and down-regulated DEGs in Dnmt3a1–/– neuron nuclei. The gene body was scaled to a 10 kb region. Each red line represents a DMR. k, Fraction of genes covered by DMRs in the gene body and ± 5 kb flanking regions in Dnmt3a1–/– neuron nuclei. l, Average DNA methylation levels across Dnmt3a1–/–-cortex DEGs and other genes in WT and Dnmt3a1–/– cortices. m, Changes of CpG methylation levels (Dnmt3a1 KO–WT) across Dnmt3a1–/– cortical DEGs and other genes.

Extended Data Fig. 9 The N-terminus is required for DNMT3A1 enrichment around bivalent promoters via binding to H2AK119ub.

a, Schematic diagram of DNMT3A constructs that were re-expressed in Dnmt3a–/– mouse ESCs. A 3× FLAG tag and a nuclear localization signal (NLS) were added at the N-terminus of each protein. b, Western blots showing re-expression of DNMT3A variants in established stable ESC lines. The experiment was repeated two times with similar results. c, Density plots of H3K4me3, H3K27me3 and H3K36me3 ChIP-seq signals at each group of TSSs indicated in Fig. 7a. d, Heatmaps for DNMT3A1 (WT) and DNMT3A1ΔN binding profiles at regions flanking TSSs in the cerebral cortex. A C-terminus antibody was used in DNMT3A ChIP-seq. The same heatmaps for H3K4me3 and H3K27me3 marks are also shown in Fig. 2a. e, Coomassie blue staining of purified GFP-fused DNMT3A1 N-terminal fragments: N121 and N122–219 (left), N219 and N219 mutants (right). f, DNMT3A1 protein structure predicted by AlphaFold (https://alphafold.ebi.ac.uk). The UIM is a part of an α-helix in the flexible N-terminus of DNMT3A1. Met 220 is the first amino acid of DNMT3A2. g, Coomassie blue staining of purified GFP-fused DNMT3A1 N278 and DNMT3B1 N-terminus (aa 1–222). h, Western blot for DNMT3A1 N278 and DNMT3B1 N-terminus after pulldown assays with the indicated nucleosomes. The experiment was repeated three times independently with similar results. i, Heatmap showing genome-wide Spearman correlation between H3K4me3, H3K27me3, H2AK119ub marks and DNMT3A1 (cortex) binding in neuron nuclei. j, IGV displays of H3K4me3, H3K27me3, H2AK119ub and DNMT3A1 enrichment at a region on chromosome 2 in the cerebral cortex and neuron nuclei.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1 and 2.

Reporting Summary

Supplementary Tables

Supplementary Tables 1 and 2.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 4

Unprocessed western blots.

Source Data Fig. 7

Unprocessed western blots.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed gel.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 9

Unprocessed western blots and gels.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gu, T., Hao, D., Woo, J. et al. The disordered N-terminal domain of DNMT3A recognizes H2AK119ub and is required for postnatal development. Nat Genet 54, 625–636 (2022). https://doi.org/10.1038/s41588-022-01063-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-022-01063-6

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