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:

Discovery of a first-in-class reversible DNMT1-selective inhibitor with improved tolerability and efficacy in acute myeloid leukemia

Abstract

DNA methylation, a key epigenetic driver of transcriptional silencing, is universally dysregulated in cancer. Reversal of DNA methylation by hypomethylating agents, such as the cytidine analogs decitabine or azacytidine, has demonstrated clinical benefit in hematologic malignancies. These nucleoside analogs are incorporated into replicating DNA where they inhibit DNA cytosine methyltransferases DNMT1, DNMT3A and DNMT3B through irreversible covalent interactions. These agents induce notable toxicity to normal blood cells thus limiting their clinical doses. Herein we report the discovery of GSK3685032, a potent first-in-class DNMT1-selective inhibitor that was shown via crystallographic studies to compete with the active-site loop of DNMT1 for penetration into hemi-methylated DNA between two CpG base pairs. GSK3685032 induces robust loss of DNA methylation, transcriptional activation and cancer cell growth inhibition in vitro. Due to improved in vivo tolerability compared with decitabine, GSK3685032 yields superior tumor regression and survival mouse models of acute myeloid leukemia.

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: Biochemical and cellular engagement of DNMT1 by GSK3484862.
Fig. 2: Biochemical and biophysical characterization of DNMT1-selective inhibitors.
Fig. 3: Phenotypic response following treatment with DNMT inhibitors in hematologic cancer cell lines.
Fig. 4: GSK3685032 induces changes in DNA methylation and gene expression in treated cells.
Fig. 5: GSK3685032 activates immune response pathways.
Fig. 6: Mechanistic evaluation of GSK3685032 and DAC in AML.
Fig. 7: Comparison of nucleoside versus non-nucleoside DNMT inhibitors.
Fig. 8: GSK3685032 reveals improved in vivo efficacy and tolerability in AML models compared with DAC.

Similar content being viewed by others

Data availability

All data generated to support the findings of this study are available. The functional genomics data have been deposited in the NCBI Gene Expression Omnibus (GEO) and are accessible through the GEO SuperSeries accession number: GSE135207 (https://www.ncbi.nlm.nih.gov/geo). The atomic coordinates and structure factors of DNMT1–DNA (zebularine)–SAH (PDB 6X9I), DNMT1–DNA–GSK3830052 (PDB 6X9J) and DNMT1–DNA–GSK3685032 (PDB 6X9K) have been deposited in the Protein Data Bank (http://www.rcsb.org). Source data are provided with this paper. All other data are available from the corresponding authors upon reasonable request.

Code availability

The code generated to analyze Infinium Methylation EPIC array, RNA-seq gene expression and hERV expression data (Figs. 46 and Extended Data Fig. 5) can be found at https://github.com/ShawnWFoley-GSK/Pappalardi_et_al_2021. A detailed list of software and package versions can be found in the Methods section.

References

  1. 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 

  2. Pradhan, S., Bacolla, A., Wells, R. D. & Roberts, R. J. Recombinant human DNA (cytosine-5) methyltransferase. I. Expression, purification, and comparison of de novo and maintenance methylation. J. Biol. Chem. 274, 33002–33010 (1999).

    Article  CAS  PubMed  Google Scholar 

  3. Bird, A. DNA methylation patterns and epigenetic memory. Genes Dev. 16, 6–21 (2002).

    Article  CAS  PubMed  Google Scholar 

  4. Jones, P. A. & Laird, P. W. Cancer epigenetics comes of age. Nat. Genet. 21, 163–167 (1999).

    Article  CAS  PubMed  Google Scholar 

  5. Baylin, S. B. & Herman, J. G. DNA hypermethylation in tumorigenesis: epigenetics joins genetics. Trends Genet. 16, 168–174 (2000).

    Article  CAS  PubMed  Google Scholar 

  6. Ting, A. H., McGarvey, K. M. & Baylin, S. B. The cancer epigenome—components and functional correlates. Genes Dev. 20, 3215–3231 (2006).

    Article  CAS  PubMed  Google Scholar 

  7. Sorm, F., Piskala, A., Cihak, A. & Vesely, J. 5-Azacytidine, a new, highly effective cancerostatic. Experientia 20, 202–203 (1964).

    Article  CAS  PubMed  Google Scholar 

  8. Jones, P. A. & Taylor, S. M. Cellular differentiation, cytidine analogs and DNA methylation. Cell 20, 85–93 (1980).

    Article  CAS  PubMed  Google Scholar 

  9. Silverman, L. R. et al. Randomized controlled trial of azacitidine in patients with the myelodysplastic syndrome: a study of the cancer and leukemia group B. J. Clin. Oncol. 20, 2429–2440 (2002).

    Article  CAS  PubMed  Google Scholar 

  10. Oki, Y., Jelinek, J., Shen, L., Kantarjian, H. M. & Issa, J. P. Induction of hypomethylation and molecular response after decitabine therapy in patients with chronic myelomonocytic leukemia. Blood 111, 2382–2384 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Fenaux, P. et al. Efficacy of azacitidine compared with that of conventional care regimens in the treatment of higher-risk myelodysplastic syndromes: a randomised, open-label, phase III study. Lancet Oncol. 10, 223–232 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Lubbert, M. et al. Low-dose decitabine versus best supportive care in elderly patients with intermediate- or high-risk myelodysplastic syndrome (MDS) ineligible for intensive chemotherapy: final results of the randomized phase III study of the European Organisation for Research and Treatment of Cancer Leukemia Group and the German MDS Study Group. J. Clin. Oncol. 29, 1987–1996 (2011).

    Article  PubMed  CAS  Google Scholar 

  13. Agrawal, K., Das, V., Vyas, P. & Hajduch, M. Nucleosidic DNA demethylating epigenetic drugs—a comprehensive review from discovery to clinic. Pharmacol. Ther. 188, 45–79 (2018).

    Article  CAS  PubMed  Google Scholar 

  14. Gnyszka, A., Jastrzebski, Z. & Flis, S. DNA methyltransferase inhibitors and their emerging role in epigenetic therapy of cancer. Anticancer Res. 33, 2989–2996 (2013).

    CAS  PubMed  Google Scholar 

  15. Issa, J. P. & Kantarjian, H. M. Targeting DNA methylation. Clin. Cancer Res. 15, 3938–3946 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Stresemann, C. & Lyko, F. Modes of action of the DNA methyltransferase inhibitors azacytidine and decitabine. Int. J. Cancer 123, 8–13 (2008).

    Article  CAS  PubMed  Google Scholar 

  17. Chabot, G. G., Bouchard, J. & Momparler, R. L. Kinetics of deamination of 5-aza-2ʹ-deoxycytidine and cytosine arabinoside by human liver cytidine deaminase and its inhibition by 3-deazauridine, thymidine or uracil arabinoside. Biochem. Pharmacol. 32, 1327–1328 (1983).

    Article  CAS  PubMed  Google Scholar 

  18. Brueckner, B. et al. Epigenetic reactivation of tumor suppressor genes by a novel small-molecule inhibitor of human DNA methyltransferases. Cancer Res. 65, 6305–6311 (2005).

    Article  CAS  PubMed  Google Scholar 

  19. Stresemann, C., Brueckner, B., Musch, T., Stopper, H. & Lyko, F. Functional diversity of DNA methyltransferase inhibitors in human cancer cell lines. Cancer Res. 66, 2794–2800 (2006).

    Article  CAS  PubMed  Google Scholar 

  20. Manara, M. C. et al. A quinoline-based DNA methyltransferase inhibitor as a possible adjuvant in osteosarcoma therapy. Mol. Cancer Ther. 17, 1881–1892 (2018).

    Article  CAS  PubMed  Google Scholar 

  21. Datta, J. et al. A new class of quinoline-based DNA hypomethylating agents reactivates tumor suppressor genes by blocking DNA methyltransferase 1 activity and inducing its degradation. Cancer Res. 69, 4277–4285 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Valente, S. et al. Selective non-nucleoside inhibitors of human DNA methyltransferases active in cancer including in cancer stem cells. J. Med. Chem. 57, 701–713 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Gros, C. et al. New insights on the mechanism of quinoline-based DNA methyltransferase inhibitors. J. Biol. Chem. 290, 6293–6302 (2015).

    Article  CAS  PubMed  Google Scholar 

  24. Shirahata, A. et al. Vimentin methylation as a marker for advanced colorectal carcinoma. Anticancer Res. 29, 279–281 (2009).

    CAS  PubMed  Google Scholar 

  25. Cong, H. et al. DNA hypermethylation of the vimentin gene inversely correlates with vimentin expression in intestinal- and diffuse-type gastric cancer. Oncol. Lett. 11, 842–848 (2016).

    Article  CAS  PubMed  Google Scholar 

  26. Constantinides, P. G., Jones, P. A. & Gevers, W. Functional striated muscle cells from non-myoblast precursors following 5-azacytidine treatment. Nature 267, 364–366 (1977).

    Article  CAS  PubMed  Google Scholar 

  27. Song, J., Rechkoblit, O., Bestor, T. H. & Patel, D. J. Structure of DNMT1-DNA complex reveals a role for autoinhibition in maintenance DNA methylation. Science 331, 1036–1040 (2011).

    Article  CAS  PubMed  Google Scholar 

  28. Song, J., Teplova, M., Ishibe-Murakami, S. & Patel, D. J. Structure-based mechanistic insights into DNMT1-mediated maintenance DNA methylation. Science 335, 709–712 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Joshi, M., Rajpathak, S. N., Narwade, S. C. & Deobagkar, D. Ensemble-based virtual screening and experimental validation of inhibitors targeting a novel site of human DNMT1. Chem. Biol. Drug Des. 88, 5–16 (2016).

    Article  CAS  PubMed  Google Scholar 

  30. Zhou, L. et al. Zebularine: a novel DNA methylation inhibitor that forms a covalent complex with DNA methyltransferases. J. Mol. Biol. 321, 591–599 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Afonine, P. V. et al. Towards automated crystallographic structure refinement with phenix.refine. Acta Crystallogr. D 68, 352–367 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Jones, P. A. & Baylin, S. B. The fundamental role of epigenetic events in cancer. Nat. Rev. Genet. 3, 415–428 (2002).

    Article  CAS  PubMed  Google Scholar 

  33. Siebenkas, C. et al. Inhibiting DNA methylation activates cancer testis antigens and expression of the antigen processing and presentation machinery in colon and ovarian cancer cells. PLoS ONE 12, e0179501 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Li, H. et al. Immune regulation by low doses of the DNA methyltransferase inhibitor 5-azacitidine in common human epithelial cancers. Oncotarget 5, 587–598 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Chiappinelli, K. B. et al. Inhibiting DNA methylation causes an interferon response in cancer via dsRNA including endogenous retroviruses. Cell 162, 974–986 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Roulois, D. et al. DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts. Cell 162, 961–973 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Saba, H. I. Decitabine in the treatment of myelodysplastic syndromes. Ther. Clin. Risk Manag. 3, 807–817 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Hashimoto, H. et al. Recognition and potential mechanisms for replication and erasure of cytosine hydroxymethylation. Nucleic Acids Res. 40, 4841–4849 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Gilmartin, A. G. et al. In vitro and in vivo induction of fetal hemoglobin with a reversible and selective DNMT1 inhibitor. Haematologica https://doi.org/10.3324/haematol.2020.248658 (2020).

  40. Ariazi, J. L. et al. Discovery and preclinical characterization of GSK1278863 (daprodustat), a small molecule hypoxia inducible factor-prolyl hydroxylase inhibitor for anemia. J. Pharmacol. Exp. Ther. 363, 336–347 (2017).

    Article  CAS  PubMed  Google Scholar 

  41. Dyachenko, V. D. K., Krivokolysko, S. G. & Litvinov, V. P. Synthesis and transformations of 6-amino-3,5-dicyano-4ethylpyridine-2(1H)-thione. Chem. Heterocycl. Compd. 32, 942–946 (1996).

    Article  Google Scholar 

  42. Stromgaard, K. et al. Ginkgolide derivatives for photolabeling studies: preparation and pharmacological evaluation. J. Med. Chem. 45, 4038–4046 (2002).

    Article  CAS  PubMed  Google Scholar 

  43. Sou, T. & Bergstrom, C. A. S. Automated assays for thermodynamic (equilibrium) solubility determination. Drug Discov. Today Technol. 27, 11–19 (2018).

    Article  PubMed  Google Scholar 

  44. Valko, K., Bevan, C. & Reynolds, D. Chromatographic hydrophobicity index by fast-gradient RP-HPLC: a high-throughput alternative to log P/log D. Anal. Chem. 69, 2022–2029 (1997).

    Article  CAS  PubMed  Google Scholar 

  45. Bunally, S. Y. & Robert, J. Y. The role and impact of high throughput biomimetic measurements in drug discovery. ADMET DMPK 6, 74–84 (2018).

    Article  Google Scholar 

  46. Young, R. J., Green, D. V., Luscombe, C. N. & Hill, A. P. Getting physical in drug discovery II: the impact of chromatographic hydrophobicity measurements and aromaticity. Drug Discov. Today 16, 822–830 (2011).

    Article  CAS  PubMed  Google Scholar 

  47. Horiuchi, K. Y. et al. Assay development for histone methyltransferases. Assay Drug Dev. Technol. 11, 227–236 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Anastassiadis, T., Deacon, S. W., Devarajan, K., Ma, H. & Peterson, J. R. Comprehensive assay of kinase catalytic activity reveals features of kinase inhibitor selectivity. Nat. Biotechnol. 29, 1039–1045 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Moriarty, N. W., Grosse-Kunstleve, R. W. & Adams, P. D. electronic Ligand Builder and Optimization Workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta Crystallogr. D 65, 1074–1080 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D 60, 2126–2132 (2004).

    Article  CAS  PubMed  Google Scholar 

  52. McCabe, M. T. et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 492, 108–112 (2012).

    Article  CAS  PubMed  Google Scholar 

  53. Fedoriw, A. et al. Anti-tumor activity of the type I PRMT inhibitor, GSK3368715, synergizes with PRMT5 inhibition through MTAP loss. Cancer Cell 36, 100–114 e125 (2019).

    Article  CAS  PubMed  Google Scholar 

  54. Savitski, M. M. et al. Multiplexed proteome dynamics profiling reveals mechanisms controlling protein homeostasis. Cell 173, 260–274 e225 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Franken, H. et al. Thermal proteome profiling for unbiased identification of direct and indirect drug targets using multiplexed quantitative mass spectrometry. Nat. Protoc. 10, 1567–1593 (2015).

    Article  CAS  PubMed  Google Scholar 

  56. Werner, T. et al. Ion coalescence of neutron encoded TMT 10-plex reporter ions. Anal. Chem. 86, 3594–3601 (2014).

    Article  CAS  PubMed  Google Scholar 

  57. Le, T., Kim, K. P., Fan, G. & Faull, K. F. A sensitive mass spectrometry method for simultaneous quantification of DNA methylation and hydroxymethylation levels in biological samples. Anal. Biochem. 412, 203–209 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  59. Jin, Y., Tam, O. H., Paniagua, E. & Hammell, M. TEtranscripts: a package for including transposable elements in differential expression analysis of RNA-seq datasets. Bioinformatics 31, 3593–3599 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Wang, L., Wang, S. & Li, W. RSeQC: quality control of RNA-seq experiments. Bioinformatics 28, 2184–2185 (2012).

    Article  CAS  PubMed  Google Scholar 

  61. Ewels, P., Magnusson, M., Lundin, S. & Kaller, M. MultiQC: summarize analysis results for multiple tools and samples in a single report. Bioinformatics 32, 3047–3048 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Patro, R., Duggal, G., Love, M. I., Irizarry, R. A. & Kingsford, C. Salmon provides fast and bias-aware quantification of transcript expression. Nat. Methods 14, 417–419 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Soneson, C., Love, M. I. & Robinson, M. D. Differential analyses for RNA-seq: transcript-level estimates improve gene-level inferences. F1000Res 4, 1521 (2015).

    Article  CAS  PubMed  Google Scholar 

  64. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Benjamini, Y., Drai, D., Elmer, G., Kafkafi, N. & Golani, I. Controlling the false discovery rate in behavior genetics research. Behav. Brain Res. 125, 279–284 (2001).

    Article  CAS  PubMed  Google Scholar 

  66. Hulsen, T., de Vlieg, J. & Alkema, W. BioVenn—a web application for the comparison and visualization of biological lists using area-proportional Venn diagrams. BMC Genomics 9, 488 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank T. Tomaszek and P. Keller for their contributions toward the high-throughput screen, K. Morasco and D. Depagnier for execution of the CBC panel, D. Cooper and H. Tran for input regarding statistical analysis, and S. Rajapurkar for assistance with genomics data analysis and visualization. We also thank A.C. Wong for original development of the DNA-binding and fluorescence-coupled breaklight assays, B. Waszkowycz for early computational design input and H. Hashimoto for his effort preparing initial human DNMT1 constructs and protein for the crystallization studies. Lastly, we thank S. Pessagno, D. Wilson and M. Bottomley for alliance oversight. Work conducted at the Cancer Research UK Manchester Institute was wholly funded by Cancer Research UK (grant nos. C480/A11411 and C5759/A17098). Work at the MD Anderson Cancer Center was supported by the Cancer Prevention Research Institute of Texas (CPRIT) grant no. RR160029 and National Institutes of Health (NIH) grant no. R35GM134744 to X.C., who is a CPRIT Scholar in Cancer Research. Work at the Van Andel Research Institute (P.A.J.) was supported by National Cancer Institute grant no. R35CA209859 and by the Van Andel Research Institute–Stand Up to Cancer Epigenetics Dream Team. Stand Up to Cancer is a division of the Entertainment Industry Foundation, administered by AACR.

Author information

Authors and Affiliations

Authors

Contributions

M.P. and A.J.J. contributed to the high-throughput screen. A.N.T. and T.H. generated DNMT proteins. M.B.P., A.S., K.W. and C.B. designed, executed and/or analyzed enzymatic studies. M.C. conducted the DNMT3B −/− HCT-116 vimentin assays. M. Mebrahtu and J.-P.J. executed the wild-type HCT-116 vimentin assay. A.S. and E.F. executed the DNA binding assay. M.S., T.W., H.C.E., A.Rutkowska, M.B. and P.G. contributed to thermal shift assays. J.B. and N.C. conducted the covalent modification and photoaffinity labeling studies. A.K.W. executed the mouse embryonic fibroblast studies under guidance from P.A.J. A.P.G. performed sequence alignment. A.G. executed the mutagenesis studies under guidance from A.G.G. and A.B.B. S.P. performed protein purification, complex formation and crystallization under guidance from X.Z. J.R.H. collected X-ray diffraction data and determined structures. K.K. and S.M. executed the AML proliferation studies. M.T.M., K.K., J.L.H., C.F.M. and M.B.P. designed, executed and/or analyzed the AML in vitro studies. K.K. executed the western blot studies with advice from S.B.B.; D.E.M. developed the LC–MS/MS methodology for the 5-methylcytosine assay. W.A.K. and S.W.F. analyzed the genomics data. C.S., C.F.M., E.M. and M.B.P. designed, executed and/or analyzed the in vivo studies. M. Muliaditan modeled the pharmacokinetic data. A.Raoof, R.J.B., L.R., D.T.F., C.Z., M.L., S.P.R., K.G., C.S.K., A.B.B., D.H., B.W.K., J.I.L. and A.M.J. designed and/or synthesized compounds. X.C. organized and designed the scope of the crystallography study. M.C., P.C., H.P.M., R.K.P., D.O., C.C., R.G.K., I.W., M.T.M. and M.B.P. provided conceptual advice. M.B.P. and M.T.M. wrote the manuscript.

Corresponding authors

Correspondence to Melissa B. Pappalardi or Michael T. McCabe.

Ethics declarations

Competing interests

M.B.P., K.K., W.A.K., C.S., K.W., J.B., M.S., A.G., C.F.M., N.C., A.P.G., T.W., L.R., D.T.F., C.Z., J.L.H., M.Muliaditan, M.Mebrahtu, J.P.J., D.E.M., H.C.E., A.N.T., T.H., S.M., S.W.F., A. Rutkowska, M.L., S.P.R., M.B., A.J.J., E.M., P.G., M.P., A.B.B., H.P.M., A.G.G., R.K.P., C.C., D.H., B.W.K., J.I.L., R.G.K. and M.T.M. are/were employees and/or shareholders of GlaxoSmithKline (GSK). The remaining authors declare no competing interests.

Additional information

Peer review information Nature Cancer thanks the anonymous reviewers 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 Characterization of GSK3484862 and GSK3685032.

a, Stability of GSK3484862 (1,000 nM) as determined by LC-MS/MS in media without or with cells (MV4-11) at 37 °C versus decitabine (1,000 nM). b, Thermal profile (Tm, 77.5 °C, n = 2 biologically independent experiments with two technical replicates each) of the hemi-methylated hairpin oligonucleotide with DMSO or GSK3484862 (100 µM). c, DNMT1 activity (average of technical replicates, data is representative of two biologically independent experiments) for uninhibited reaction (DMSO, n = 2) and recovery of inhibited enzyme activity following rapid dilution (100-fold) of pre-complexed DNMT1:GSK3484862 (n = 3) or DNMT1:SAH (n = 3). d, Dose-dependent increase in vimentin expression (n = 2 biologically independent experiments with two technical replicates each) following treatment of DNMT3B -/- HCT-116 cells with GSK3484862 or decitabine. e, Table summarizing up-regulation of vimentin expression in wild-type or DNMT3B -/- HCT-116 cells after treatment with GSK3484862 or decitabine (average ± s.d.; n = independently fitted EC50 values). f, IC50 values (bar represents average, n = biologically independent determinations) following a 0-minute (n = 15), 60-minute DNMT1:Inhibitor (EI, n = 2), or 60-minute DNMT1:Inhibitor:hemi-methylated DNA (ESI, n = 2) preincubation. g, Intact protein mass spectrometry for mDNMT1 (731-1602) following incubation with hemi-methylated DNA in the absence or presence of GSK3685032 showed no covalent adduct. h, Inhibition of a kinase panel (n = 369) by 10 µM GSK3685032. i, Inhibition of a methyltransferase panel by 10 µM GSK3685032. j, Isothermal dose-response curves for DNMT1 following treatment with GSK3685032 in a recombinant system (DNMT1 601-1600 in the presence of 40-mer hemi-methylated DNA) or in a cellular system (HepG2). k, Dose response curves (average ± s.d., n = biologically independent samples) for full-length DNMT1 using a 40-mer hemi-methylated or poly(dIdC) DNA substrate in a radioactive SPA assay with GSK3685032 (n = 4 or 5, respectively) or SAH (n = 4 or 6), respectively).

Source data

Extended Data Fig. 2 DNMT1 residues important for compound binding and inhibition.

a, Analogues containing a photoreactive benzophenone or diazerine moiety. b, c, Murine DNMT1 (731-1602) spectra in the absence or presence of a 45-minute photolysis step with 14-mer hemi-methylated DNA and GSK3844831 (b) or GSK3901839 (c). d, Dose response curves for HEK293 cells expressing either wild-type or site-directed alanine mutant DNMT1 (n = 2; technical replicates) treated for 6 days with decitabine or GSK3685032. Dashed line represents starting cell number (T0). e, Dose response curves (n = 4 biologically independent samples; average ± s.d.) for full-length wild-type or H1507Y DNMT1 activity in a radioactive SPA assay.

Source data

Extended Data Fig. 3 Sequence alignment of the methyltransferase domains of human DNMT1, DNMT3A, and DNMT3B.

Identical residues are shaded blue while similar residues are shaded yellow. The boxes indicate the target recognition domain of DNMT1 (dashed, black) and the active-site loop (solid, red). Residues that were photoaffinity labeled*, residues that conferred resistance to GSK3685032 upon mutation to alanine (gIC50 > 10 µM)† and are reported to be involved in recognition of the methylated cytosine‡, or the catalytic cysteine (C1226)♯ are marked within the DNMT1 sequence.

Extended Data Fig. 4 Crystal structure of DNMT1-DNA in complex with DNMT1 inhibitor.

a, View of the active-site loop bound in the space left by the flipped-out zebularine in the DNMT1-DNA complex. b, The omit electron density map in mesh for GSK3685032 contoured at 4σ above the mean. c, d, Orthogonal views of DNMT1-DNA in the presence of GSK3685032. The active-site loop is colored brown and adopts an open conformation. e, The omit electron density map in mesh for GSK3830052 contoured at 4σ above the mean. f, Superimposition of inhibitor (pink) and the active-site loop in the native complex (cyan). g, The inhibitor intercalates into DNA between two G:C base pairs. h, Two hydrogen bonds formed between G1 and zebularine. i, Inhibitor interacts with 5-methylcytosine (5mC) of the parent DNA strand and Trp1510 of DNMT1. j, The end of the inhibitor N-methyl-N-phenylmethanesulfonamide moiety is close to the DNMT1 active-site Cys1226.

Extended Data Fig. 5 Biochemical, phenotypic, and mechanistic activity of DNMT inhibitors.

a, Table reporting GSK3685032 activity in a panel of AML cell lines (day 6, average ± s.d., n = biologically independent experiments). b, Heatmap showing induction of caspase-glo 3/7 activity (Promega, average log2 fold change, n = 2 biologically independent experiments) in MV4-11 cells following treatment with GSK3510477, GSK3484862 (with 2 technical replicates), or GSK3685032 at days 1, 2, 4 & 6 (0.06-7,340 nM). c, Compound structures for reported DNMT inhibitors. d, Table containing output parameters (average ± SEM, n = biologically independent experiments) following biochemical, phenotypic (MV4-11, day 6) or mechanistic (MV4-11, day 4) assessment using a panel of DNMT inhibitors. NA, not applicable to assay format. 5mC, 5-methylcytosine. e, Top, Venn diagram for significantly increased genes in MV4-11 (RNA-seq, FDR < 0.05, |log2 fold-change| > 1, day 4, 400 nM) following treatment with GSK3685032 or decitabine. Bottom, Heatmap of log2 fold change differential expression (RNA-seq, day 4) following treatment with GSK3685032 or decitabine (3.2-10,000 nM) for overlapping genes (n = 1,542) from the Venn diagram.

Source data

Extended Data Fig. 6 Pharmacokinetic evaluation of GSK3685032.

a, Summary of mouse pharmacokinetic parameters for GSK3685032. NA, not applicable to dosing route. ND, value not determined. IV, intravenous. SC, subcutaneous. b, Blood concentration of GSK3685032 at multiple timepoints following a single dose of 2 mg/kg IV (male CD-1 mice), 2 mg/kg SC (male C57/BL6 mice), or 30 mg/kg SC (female Nu/Nu mice). Individual data shown (n = 3 animals/group). c, Dose proportional blood concentration of GSK3685032 following twice daily subcutaneous dosing for 8.5 days in a SKM-1 subcutaneous xenograft model (NOD-scid) collected 6 hours post last dose. Individual concentrations (n = 3 animals/group) with linear regression (R square = 0.9780) fit to the mean concentration for each group. d, Simulated profile of GSK3685032 over a 24 hour time frame adjusted for unbound fraction (2.5%) in the blood following twice daily subcutaneous dosing. 5-mC, 5-methylcytosine.

Source data

Extended Data Fig. 7 Compound effect in subcutaneous MV4-11 and SKM-1 xenograft models.

a, b, Animal body weight measurements for MV4-11 (a) or SKM-1 (b) xenograft models spanning the dosing duration of the study (average ± s.d.; n = 10 animals/group, # represents day first animal came off study due to tumor volume). c-f, Individual tumor volume measurements for MV4-11 (c, day 35) or SKM-1 (d, day 20). Solid line represents the median for each group (n = 10 animals unless noted). Dotted line represents the median tumor volume for vehicle. Statistical significance* of treatment versus vehicle was calculated using one-way ANOVA, Dunnett’s multiple comparisons test. Table summarizes adjusted P values to account for multiple comparisons and corresponding tumor growth inhibition (TGI) values for each group within the MV4-11 (e) or SKM-1 (f) xenograft models. g, h, Individual tumor volume measurements for the 45 mg/kg GSK3685032 group in MV4-11 (g) or SKM-1 (h) xenograft models during the dosing segment (orange bar) and continuing for ≥ 27 days off drug (blue bar) to monitor durability (n = 10 animals unless noted). The minimum measurable tumor volume was set to 10 mm3.

Source data

Extended Data Fig. 8 Effects of GSK3685032 and decitabine on complete blood cell counts.

a, b, Complete blood cell count (a) at day 28 across all dose groups (n = 5 animals/group; mean ± s.d.). Statistical significance* of treatment versus vehicle was calculated using one-way ANOVA, Dunnett’s multiple comparisons test. Each P value was adjusted to account for multiple comparisons. Table (b) showing output parameters following statistical analysis. ^Used log10 transformed values due to unequal variance between groups. Ratio represents treatment group normalized to vehicle. c, Complete blood cell count (mean ± s.d.) at day 28 on treatment (n = 5 animals/group) followed by 27 days off treatment (n = 5 animals/group for 15 and 45 mg/kg or n = 8 animals for 30 mg/kg group) with GSK3685032.

Source data

Supplementary information

Supplementary Information

Supplementary Data Tables 1–5.

Reporting Summary

Supplementary Table 6

Directed mutagenesis constructs.

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. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 7

Unprocessed western blots.

Source Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pappalardi, M.B., Keenan, K., Cockerill, M. et al. Discovery of a first-in-class reversible DNMT1-selective inhibitor with improved tolerability and efficacy in acute myeloid leukemia. Nat Cancer 2, 1002–1017 (2021). https://doi.org/10.1038/s43018-021-00249-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43018-021-00249-x

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer