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 CIP2A–TOPBP1 axis safeguards chromosome stability and is a synthetic lethal target for BRCA-mutated cancer

Abstract

BRCA1/2-mutated cancer cells adapt to the genome instability caused by their deficiency in homologous recombination (HR). Identification of these adaptive mechanisms may provide therapeutic strategies to target tumors caused by the loss of these genes. In the present study, we report genome-scale CRISPR–Cas9 synthetic lethality screens in isogenic pairs of BRCA1- and BRCA2-deficient cells and identify CIP2A as an essential gene in BRCA1- and BRCA2-mutated cells. CIP2A is cytoplasmic in interphase but, in mitosis, accumulates at DNA lesions as part of a complex with TOPBP1, a multifunctional genome stability factor. Unlike PARP inhibition, CIP2A deficiency does not cause accumulation of replication-associated DNA lesions that require HR for their repair. In BRCA-deficient cells, the CIP2A–TOPBP1 complex prevents lethal mis-segregation of acentric chromosomes that arises from impaired DNA synthesis. Finally, physical disruption of the CIP2A–TOPBP1 complex is highly deleterious in BRCA-deficient tumors, indicating that CIP2A represents an attractive synthetic lethal therapeutic target for BRCA1- and BRCA2-mutated cancers.

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: CIP2A loss is synthetically lethal with BRCA1 or BRCA2 deficiency.
Fig. 2: CIP2A-null cells do not have increased loads of HR-engaging DNA lesions.
Fig. 3: CIP2A prevents acentric chromosome segregation.
Fig. 4: CIP2A colocalizes with TOPBP1 on mitotic structures.
Fig. 5: The CIP2A–TOPBP1 interaction is essential in BRCA2−/− cells.
Fig. 6: Disruption of the CIP2A–TOPBP1 complex causes genome instability and lethality in BRCA-deficient cells.
Fig. 7: CIP2A prevents S phase-associated DNA lesions from causing acentric chromosome mis-segregation.
Fig. 8: Therapeutic proof of concept.

Similar content being viewed by others

Data availability

Source data for Figs. 18 and Extended Data Figs. 110 have been provided. Illumina sequencing data for the screens were deposited at the National Center for Biotechnology Information (accession no. PRJNA753451). Further information on research design is available in the Nature Research Reporting Summary linked to this article. All other data supporting the findings of the present study are available from the corresponding author on reasonable request.

Code availability

The CCA code is also available at Zenodo (https://doi.org/10.5281/zenodo.5149154).

References

  1. Roy, R., Chun, J. & Powell, S. N. BRCA1 and BRCA2: different roles in a common pathway of genome protection. Nat. Rev. Cancer 12, 68–78 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  2. Moynahan, M. E. & Jasin, M. Mitotic homologous recombination maintains genomic stability and suppresses tumorigenesis. Nat. Rev. Mol. Cell Biol. 11, 196–207 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Nik-Zainal, S. et al. Landscape of somatic mutations in 560 breast cancer whole-genome sequences. Nature 534, 47–54 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Setton, J. et al. Synthetic lethality in cancer therapeutics: the next generation. Cancer Discov. https://doi.org/10.1158/2159-8290.CD-20-1503 (2021).

  5. Lord, C. J. & Ashworth, A. PARP inhibitors: synthetic lethality in the clinic. Science 355, 1152–1158 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Feng, W. & Jasin, M. BRCA2 suppresses replication stress-induced mitotic and G1 abnormalities through homologous recombination. Nat. Commun. 8, 525 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Hakem, R. et al. The tumor suppressor gene Brca1 is required for embryonic cellular proliferation in the mouse. Cell 85, 1009–1023 (1996).

    Article  CAS  PubMed  Google Scholar 

  8. Bouwman, P. et al. 53BP1 loss rescues BRCA1 deficiency and is associated with triple-negative and BRCA-mutated breast cancers. Nat. Struct. Mol. Biol. 17, 688–695 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Hakem, R., de la Pompa, J. L., Elia, A., Potter, J. & Mak, T. W. Partial rescue of Brca1 (5–6) early embryonic lethality by p53 or p21 null mutation. Nat. Genet. 16, 298–302 (1997).

    Article  CAS  PubMed  Google Scholar 

  10. Ludwig, T., Chapman, D. L., Papaioannou, V. E. & Efstratiadis, A. Targeted mutations of breast cancer susceptibility gene homologs in mice: lethal phenotypes of Brca1, Brca2, Brca1/Brca2, Brca1/p53, and Brca2/p53 nullizygous embryos. Genes Dev. 11, 1226–1241 (1997).

    Article  CAS  PubMed  Google Scholar 

  11. Gowen, L. C., Johnson, B. L., Latour, A. M., Sulik, K. K. & Koller, B. H. Brca1 deficiency results in early embryonic lethality characterized by neuroepithelial abnormalities. Nat. Genet. 12, 191–194 (1996).

    Article  CAS  PubMed  Google Scholar 

  12. Mengwasser, K. E. et al. Genetic screens reveal FEN1 and APEX2 as BRCA2 synthetic lethal targets. Mol. Cell https://doi.org/10.1016/j.molcel.2018.12.008 (2019).

  13. Alvarez-Quilon, A. et al. Endogenous DNA 3′ blocks are vulnerabilities for BRCA1 and BRCA2 deficiency and are reversed by the APE2 nuclease. Mol. Cell 78, 1152–1165 e1158 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Farmer, H. et al. Targeting the DNA repair defect in BRCA mutant cells as a therapeutic strategy. Nature 434, 917–921 (2005).

    Article  CAS  PubMed  Google Scholar 

  15. Bryant, H. E. et al. Specific killing of BRCA2-deficient tumours with inhibitors of poly(ADP-ribose) polymerase. Nature 434, 913–917 (2005).

    Article  CAS  PubMed  Google Scholar 

  16. Zimmermann, M. et al. CRISPR screens identify genomic ribonucleotides as a source of PARP-trapping lesions. Nature 559, 285–289 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Blessing, C. et al. The oncogenic helicase ALC1 regulates PARP inhibitor potency by trapping PARP2 at DNA breaks. Mol. Cell 80, 862–875 e866 (2020).

    Article  CAS  PubMed  Google Scholar 

  18. Verma, P. et al. ALC1 links chromatin accessibility to PARP inhibitor response in homologous recombination-deficient cells. Nat. Cell Biol. 23, 160–171 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Mateos-Gomez, P. A. et al. Mammalian polymerase theta promotes alternative NHEJ and suppresses recombination. Nature 518, 254–257 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ceccaldi, R. et al. Homologous-recombination-deficient tumours are dependent on Poltheta-mediated repair. Nature 518, 258–262 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Dempster, J. M. et al. Extracting biological insights from the project achilles genome-scale CRISPR screens in cancer cell lines. Preprint at bioRxiv https://doi.org/10.1101/720243 (2019).

  22. Behan, F. M. et al. Prioritization of cancer therapeutic targets using CRISPR–Cas9 screens. Nature https://doi.org/10.1038/s41586-019-1103-9 (2019).

  23. Hustedt, N. et al. A consensus set of genetic vulnerabilities to ATR inhibition. Open Biol. 9, 190156 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Kim, E. & Hart, T. Improved analysis of CRISPR fitness screens and reduced off-target effects with the BAGEL2 gene essentiality classifier. Genome Med. 13, 2 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Wang, J. et al. Oncoprotein CIP2A is stabilized via interaction with tumor suppressor PP2A/B56. EMBO Rep. 18, 437–450 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Junttila, M. R. et al. CIP2A inhibits PP2A in human malignancies. Cell 130, 51–62 (2007).

    Article  CAS  PubMed  Google Scholar 

  27. Khanna, A., Pimanda, J. E. & Westermarck, J. Cancerous inhibitor of protein phosphatase 2A, an emerging human oncoprotein and a potential cancer therapy target. Cancer Res. 73, 6548–6553 (2013).

    Article  CAS  PubMed  Google Scholar 

  28. Olivieri, M. et al. A genetic map of the response to DNA damage in human cells. Cell 182, 481–496 e421 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Wilson, D. M. 3rd & Thompson, L. H. Molecular mechanisms of sister-chromatid exchange. Mutat. Res. 616, 11–23 (2007).

    Article  CAS  PubMed  Google Scholar 

  30. Kim, J. S., Kim, E. J., Oh, J. S., Park, I. C. & Hwang, S. G. CIP2A modulates cell-cycle progression in human cancer cells by regulating the stability and activity of Plk1. Cancer Res. 73, 6667–6678 (2013).

    Article  CAS  PubMed  Google Scholar 

  31. Rizk, A. et al. Segmentation and quantification of subcellular structures in fluorescence microscopy images using Squassh. Nat. Protoc. 9, 586–596 (2014).

    Article  CAS  PubMed  Google Scholar 

  32. Bunting, S. F. et al. 53BP1 inhibits homologous recombination in Brca1-deficient cells by blocking resection of DNA breaks. Cell 141, 243–254 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Leimbacher, P. A. et al. MDC1 Interacts with TOPBP1 to maintain chromosomal stability during mitosis. Mol. Cell 74, 571–583 e578 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Laine, A. et al. CIP2A interacts with TopBP1 and drives basal-like breast cancer tumorigenesis. Cancer Res. https://doi.org/10.1158/0008-5472.CAN-20-3651 (2021).

  35. Bagge, J., Oestergaard, V. H. & Lisby, M. Functions of TopBP1 in preserving genome integrity during mitosis. Semin. Cell Dev. Biol. https://doi.org/10.1016/j.semcdb.2020.08.009 (2020).

  36. Broderick, R., Nieminuszczy, J., Blackford, A. N., Winczura, A. & Niedzwiedz, W. TOPBP1 recruits TOP2A to ultra-fine anaphase bridges to aid in their resolution. Nat. Commun. 6, 6572 (2015).

    Article  CAS  PubMed  Google Scholar 

  37. Ozer, O. & Hickson, I. D. Pathways for maintenance of telomeres and common fragile sites during DNA replication stress. Open Biol. https://doi.org/10.1098/rsob.180018 (2018).

  38. Bang, S. W. et al. Human TopBP1 localization to the mitotic centrosome mediates mitotic progression. Exp. Cell. Res. 317, 994–1004 (2011).

    Article  CAS  PubMed  Google Scholar 

  39. Jeong, A. L. et al. Cancerous inhibitor of protein phosphatase 2A (CIP2A) protein is involved in centrosome separation through the regulation of NIMA (never in mitosis gene A)-related kinase 2 (NEK2) protein activity. J. Biol. Chem. 289, 28–40 (2014).

    Article  CAS  PubMed  Google Scholar 

  40. Bhowmick, R., Minocherhomji, S. & Hickson, I. D. RAD52 facilitates mitotic DNA synthesis following replication stress. Mol. Cell 64, 1117–1126 (2016).

    Article  CAS  PubMed  Google Scholar 

  41. Elstrodt, F. et al. BRCA1 mutation analysis of 41 human breast cancer cell lines reveals three new deleterious mutants. Cancer Res. 66, 41–45 (2006).

    Article  CAS  PubMed  Google Scholar 

  42. Banaszynski, L. A., Chen, L. C., Maynard-Smith, L. A., Ooi, A. G. & Wandless, T. J. A rapid, reversible, and tunable method to regulate protein function in living cells using synthetic small molecules. Cell 126, 995–1004 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Janysek, D. C., Kim, J., Duijf, P. H. G. & Dray, E. Clinical use and mechanisms of resistance for PARP inhibitors in homologous recombination-deficient cancers. Transl. Oncol. 14, 101012 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Liu, Z. et al. Cancerous inhibitor of PP2A is targeted by natural compound celastrol for degradation in non-small-cell lung cancer. Carcinogenesis 35, 905–914 (2014).

    Article  CAS  PubMed  Google Scholar 

  45. Yu, H. C. et al. Erlotinib derivative inhibits hepatocellular carcinoma by targeting CIP2A to reactivate protein phosphatase 2A. Cell Death Dis. 5, e1359 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Blackford, A. N. & Stucki, M. How cells respond to DNA breaks in mitosis. Trends Biochem. Sci. 45, 321–331 (2020).

    Article  CAS  PubMed  Google Scholar 

  47. Lezaja, A. & Altmeyer, M. Dealing with DNA lesions: when one cell cycle is not enough. Curr. Opin. Cell Biol. 70, 27–36 (2020).

    Article  PubMed  Google Scholar 

  48. De Marco Zompit, M. et al. The CIP2A–TOPBP1 complex safeguards chromosomal stability during mitosis. Preprint at bioRxiv https://doi.org/10.1101/2021.02.08.430274 (2021).

  49. Ventela, S. et al. CIP2A promotes proliferation of spermatogonial progenitor cells and spermatogenesis in mice. PLoS ONE 7, e33209 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Shanbhag, N. M., Rafalska-Metcalf, I. U., Balane-Bolivar, C., Janicki, S. M. & Greenberg, R. A. ATM-dependent chromatin changes silence transcription in cis to DNA double-strand breaks. Cell 141, 970–981 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Noordermeer, S. M. et al. The shieldin complex mediates 53BP1-dependent DNA repair. Nature 560, 117–121 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Hart, T. et al. High-resolution CRISPR screens reveal fitness genes and genotype-specific cancer liabilities. Cell 163, 1515–1526 (2015).

    Article  CAS  PubMed  Google Scholar 

  53. Olivieri, M. & Durocher, D. Genome-scale chemogenomic CRISPR screens in human cells using the TKOv3 library. STAR Protoc. 2, 100321 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  54. Hart, T. et al. Evaluation and design of genome-wide CRISPR/SpCas9 knockout screens. G3 (Bethesda) 7, 2719–2727 (2017).

    Article  CAS  Google Scholar 

  55. Kim, E. & Hart, T. Improved analysis of CRISPR fitness screens and reduced off-target effects with the BAGEL2 gene essentiality classifier. Genome Med. 13, 2 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Ohsumi, T. K. & Borowsky, M. L. MolBioLib: a C++11 framework for rapid development and deployment of bioinformatics tasks. Bioinformatics 28, 2412–2416 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Hart, T. & Moffat, J. BAGEL: a computational framework for identifying essential genes from pooled library screens. BMC Bioinf. 17, 164 (2016).

    Article  Google Scholar 

  58. Delignette-Muller, M. L. & Dutang, C. fitdistrplus: an R package for fitting distributions. J. Statist. Software 64, 1–34 (2015).

  59. Becker, J. R. et al. BARD1 reads H2A lysine 15 ubiquitination to direct homologous recombination. Nature https://doi.org/10.1038/s41586-021-03776-w (2021).

  60. van der Weegen, Y. et al. ELOF1 is a transcription-coupled DNA repair factor that directs RNA polymerase II ubiquitylation. Nat. Cell Biol. 23, 595–607 (2021).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We dedicate this work to the memory of our friend and colleague Silvia Emma Rossi. We thank members of the Durocher lab for helpful discussions and A. Jeanrenaud for technical help. We also thank J. Moffat for his generous sharing of the TKO sgRNA libraries and R. Greenberg for the U2OS 2-6-5 cell line. We thank K. Chan at the NBCC (LTRI) for sequencing. We also thank the Repare in vivo pharmacology team (A. Roulston, A. Bonneau-Fortin, M.-È. Leclaire and S. Fournier) for their help with tumor xenograft assays. S.A. was supported by a Banting post-doctoral fellowship and S.E.R. was supported by a fellowship from Fondazione AIRC per la Ricerca sul Cancro. Y.X. holds, and A.A.Q. and D.S. held, post-doctoral fellowships from the Canadian Institutes for Health Research (CIHR). Work in M.S.’s lab was supported by grants from the Swiss National Foundation (grant no. 310030_189141) and the Groeber Foundation. D.D. is a Canada Research Chair (Tier I) and work in his lab was supported by grants from the CIHR (grant no. FDN143343) and Canadian Cancer Society (grant no. 705644), with additional support from the Krembil Foundation and Repare Therapeutics.

Author information

Authors and Affiliations

Authors

Contributions

S.A. and S.E.R. contributed equally to the present study and carried out the bulk of the experimental work, writing and revising drafts, as well as preparing figures. N.M., M.M.Z., Y.X., T.F.N., A.A.Q., D.S. and S.M.N. provided additional data and reagents. J.D. undertook the BRCA1 screen shown in the present study. V.B., G.M., S.Y.Y. and R.P. undertook the animal and pharmacokinetic studies. S.M.N. undertook the original BRCA1 synthetic lethal screen. T.K.O. provided CCA. N.H. provided the first validation of the BRCA–CIP2A synthetic lethality. R.K.S., N.C. and M.M. provided technical assistance that included key reagent generation and, for R.K.S., manuscript editing. A.V. and H.M. provided bioinformatics expertise. J.T.F.Y. supervised J.D. M.Z. supervised all authors based at Repare Therapeutics. M.S. supervised M.M.Z. D.D. conceived the project, supervised the project, wrote and revised the manuscript, prepared figures and acquired funding for the present study.

Corresponding author

Correspondence to Daniel Durocher.

Ethics declarations

Competing interests

D.D. is a shareholder and advisor of Repare Therapeutics. A.A.Q., J.D., V.B., G.M., S.Y.Y., R.P., J.T.F.Y., A.V. and M.Z. are employees of Repare Therapeutics. T.K.O. was an employee of Repare Therapeutics. The remaining authors declare no competing interests.

Additional information

Peer review information Nature Cancer thanks Dipanjan Chowdhury, Joanna Loizou and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Data supporting CIP2A as a protein recognizing mitotic DNA lesions.

a, Immunoblotting of whole-cell extracts of RPE1-hTERT p53-/- Cas9 cells, parental (WT) or BRCA1-/-, expressing the indicated sgRNAs and either a virus expressing an sgRNA-resistant FLAG-CIP2A (CIP2A*) or an empty virus (EV). Lysates were probed for FLAG, CIP2A and tubulin (loading control). Representative of two immunoblots. b, Immunoblot examining CIP2A levels in isogenic and BRCA-deficient cells in the DLD1 and RPE1-hTERT p53-/- Cas9 (RPE1-hTERT) backgrounds. Tubulin was used as a loading control. Immunoblotting was performed once. c, Clonogenic survival assay of RPE1-hTERT p53-/- Cas9 cells of the indicated genotype following treatment with camptothecin (CPT). Data presented as the mean ± S.D. n = 4 independent clonogenic assays apart from the 15 nM dose where n = 3; analyzed with extra sum-of-squares F test and compared to WT. p = 1.27 × 10−7 (#1), 1.68 × 10−6 (#6). d, e, Analysis of sister chromatid exchanges (SCEs) in cells of the indicated genotype. Quantitation is shown in (d) as cell-based data; bars represent the mean ± S.D. n = 31, 30, 31, 30, 30 aggregated from 3 independent metaphase spreads; analyzed with Kruskall-Wallis non-parametric test, followed by Dunn’s multiple comparisons, compared to WT. n.s.= 0.4398, 0.6574, p = 1.97 × 10−13, 1.18 × 10−13. Representative micrographs of metaphase spreads are shown in (e). Arrowheads indicate an SCE event. f, Immunoblotting for CIP2A in DLD1 WT and CIP2A knockout clones. Tubulin was used as a loading control. Representative of 3 independent immunoblots. g, Analysis of RAD51 foci in EdU+ cells in RPE1-hTERT (WT) and indicated CIP2A-/- clones. Left, aggregate cell-based values. Right, mean of each independent experiment. Bars represent the mean ± S.D. n = 60 aggregated from 3 independent immunostainings; analyzed with Kruskall-Wallis non-parametric test, followed by Dunn’s multiple comparisons, compared to WT; n.s.= 0.5542, 1.0000, 1.0000. All scale bars = 10 μm. Statistical analyses and full blots are found in Source Data ED Fig. 1.

Source data

Extended Data Fig. 2 CIP2A promotes genome integrity of BRCA-deficient cells.

a, b, Colocalization analysis of CIP2A and γH2AX in DLD1 cells pre- and post-IR (2 Gy) using Squassh (size parameter). Quantitation is shown in (a). Threshold for colocalization is 0.5 (dashed line). Data points represent cells analysed and bars represent the mean ± 95% C.I. n = 15 aggregated from 3 independent immunostainings. Representative micrographs are shown in (b). UT, untreated. c, Representative micrographs of the experiment shown in Fig. 3f. d, Quantitation of radial chromosomes (left) and chromatid breaks (right) in metaphase spreads from DLD1 cells upon transduction of virus expressing the indicated sgRNAs along with Cas9. The bars represent the mean ± S.D. n = 30 aggregated from 3 independent experiments with 10 metaphases scored per experiment; analyzed with Kruskall-Wallis non-parametric test, followed by Dunn’s multiple comparisons, compared to sgAAVS1; n.s.=1.0000, p = 0.0107 (radials); p = 0.0001, 0.0044 (breaks). e, Representative micrographs of the experiment presented in Fig. 3g. Arrowheads indicate chromosome aberrations. f, Representative micrographs of the experiment shown in Fig. 3h. MNi are indicated with white triangles. Scale bar = 10 μm. Statistical analyses are found in Source Data ED Fig. 2.

Source data

Extended Data Fig. 3 CIP2A acts in mitosis with TOPBP1.

a, Correlation network based on Pearson’s correlation of gene-level NormZ derived from the genotoxic dataset shown in ref. 28. b, Colocalization analysis by Squassh of the experiment shown in Fig. 4b using size as parameter. Each point represents a mitotic cell analysed, bars represent the mean ± 95% C.I. n = 15 aggregated from 3 independent immunostainings. c, Representative micrographs of the experiment quantitated in Fig. 4c. Nocodazole-treated U2OS cells previously transfected with the indicated siRNAs were fixed 1 h post-X-irradiation (2 Gy) and processed for immunofluorescence with the indicated antibodies. Scale bar = 10 μm. d, Competitive growth assays in DLD1 Cas9 cells (WT) or an isogenic BRCA2-/- counterpart transduced with virus expressing the indicated sgRNAs. Data presented as mean ± S.D. n = 3 independent growth assays; analyzed with two-way ANOVA with repeated measures and Šídák’s multiple comparisons testing. p = 0.0067 (d18), n.s.=0.46359. Statistical analyses are found in Source Data ED Fig. 3.

Source data

Extended Data Fig. 4 CIP2A and TOPBP1 localize to under-replicated DNA in mitotic cells.

a, Additional micrographs of CIP2A/TOPBP1 structures observed in mitotic cells after treatment with low-dose aphidicolin. Relates to Fig. 4e. Shown are curved (upper panels) and straight (lower panels) filaments. Representative of n = 3 independent immunostainings. b, Micrographs of U2OS parental (WT) and MDC1-/- anaphase cells that were either treated with aphidicolin (400 nM) for 16 h or left untreated. Quantitation of this experiment is shown in Fig. 4f. c, d, Colocalization analysis of TOPBP1 and CIP2A foci in U2OS parental (WT) or MDC1-/- cells using Squassh, size parameter. Representative micrographs (of two independent immunostainings) are shown in (c) quantitation shown in (d). Threshold for colocalization is 0.5 (dashed line). Bar represents the mean ± 95% C.I. n = 24 (WT) or 23 (MDC1-/-). e, Colocalization analysis of RPA and CIP2A in early prophase DLD1 BRCA2-/- cells by Squassh, size parameter. Threshold for colocalization is 0.5 (dashed line). Bar represents the mean ± 95% C.I. n = 15 aggregated from 3 independent immunostainings. Refers to Fig. 4h. f, Quantitation of spontaneous RPA and CIP2A foci in early prophase DLD1 BRCA2-/- cells. Data points are cell-based data, bar represents the mean ± S.D. n = 32 cells aggregated from 3 independent immunostainings. Analyzed with a Mann-Whitney test; U value (normal approximation)=-6.49202, p = 1.30 × 10−14. g, Representative micrographs of the experiment shown in Fig. 4i with additional MDA-MB-436 cells showing elongation of CIP2A-TOPBP1 filaments during mitosis. h, Immunoblot of lysates from cells used in the experiment shown in Fig. 4j. Lysates were probed for GFP, CIP2A and tubulin (loading control). Representative of n = 2 independent immunoblots. Full blots are provided in Blot Source Data ED Fig. 4. All scale bars = 10 µm. Statistical analyses are found in Source Data ED Fig. 4.

Source data

Extended Data Fig. 5 CIP2A interacts with TOPBP1.

a, Co-immunoprecipitation of CIP2A with TOPBP1. Whole-cell extracts from U2OS parental (WT) or MDC1-/- (KO) cells were subjected to immunoprecipitation with normal mouse IgG or a CIP2A antibody and were then immunoblotted with the indicated antibodies. Representative of two independent experiments. b, Micrographs of the LacR/LacO assay assessing the interaction between endogenous CIP2A and TOPBP1 variants fused to FLAG-LacR shown in Fig. 5f. Representative of 3 independent immunostainings. Scale bar = 10 µm. c, LacR/LacO assay assessing the interaction between endogenous CIP2A and TOPBP1 variants fused to FLAG-LacR. Data presented as the mean values ± S.D. n = 3 independent immunostainings; analyzed with one-way ANOVA, followed by multiple comparisons Dunnet test, all compared to FLAG-LacR; p = 1 × 10−15. d, Alanine scanning of TOPBP1 (830-851) residues by yeast two-hybrid assay with CIP2A (1-560). Five residues that abolish the TOPBP1-CIP2A interaction when mutated to alanine were identified. AD, activation domain; BD, Gal4 DNA binding domain. Expression of proteins was verified by immunblotting but not shown. Representative of n = 2 independent sets of transformations. e, CIP2A and TOPBP1 interact directly. Upper: 1 μg of GST or GST-CIP2A (1-560) were separated by SDS-PAGE and stained with Coomassie. Lower blot: GST pull-down experiment with either GST or GST-CIP2A (1-560) incubated with either MBP, MBP-TOPBP1756–999 or MBP-TOPBP1756–999-3A. Bound proteins were processed for immunoblotting with an anti-MBP antibody. The CIP2A/TOPBP1 interaction has been observed > 5 times; the loss of interaction by the 3 A mutation is representative of two independent pulldown assays. Full blots are provided in Blot Source Data ED Fig. 5. Statistical analyses are found in Source Data ED Fig. 5.

Source data

Extended Data Fig. 6 Data supporting that the CIP2A-TOPBP1 interaction promotes HR-deficient cell viability.

a, Immunoblots of whole-cell extracts from DLD1 parental (WT) or BRCA2-/- cells transduced with the indicated HA-tagged TOPBP1-expressing lentivirus or an empty virus that expresses an HA epitope (EV(HA)). Lysates were probed with antibodies to HA or tubulin (loading control). FL, full-length. Representative of n = 3 independent transductions. b, Micrographs of DLD1 cells transduced with the indicated virus that were arrested in mitosis with a 16 h treatment with nocodazole, exposed to a 2 Gy IR dose and processed for immunofluorescence with the indicated antibodies 1 h later. Relates to the experiment quantitated in Fig. 5h. Representative of three immunostainings. c, Representative images of crystal violet stains of the clonogenic survival experiment presented in Fig. 5i. Representative of three clonogenic assays. d, Representative micrographs of the experiment presented in Fig. 5j showing DAPI-stained cells to monitor micronucleation (labeled with arrowheads). Representative of three immunostainings. Scale bars =10 μm. Full blots are provided in Source Data ED Fig. 6.

Source data

Extended Data Fig. 7 Disruption of the TOPBP1-CIP2A interaction is lethal in BRCA-deficient cells.

a, DLD1 BRCA2-/- cells transduced with either an empty virus containing only the destabilization domain (DD; EV, left) or a virus encoding B6L (right) were treated with AS1 (1 μM) for the indicated periods or left untreated (UT). Shown are micrographs of mitotic cells stained with antibodies to CIP2A, γH2AX or FLAG (labeling the DD). Quantitation is shown in Fig. 6b. Representative of three independent immunostainings. b, Anti-FLAG immunoblots of whole-cell extracts derived from DLD1 parental (WT) or BRCA2-/- cells untreated or treated with either Shield-1 (S1) or Aqua-Shield-1 (AS1) for 72 h. KAP1 is used as a loading control. Representative of three independent immunostainings. c, Representative images of the clonogenic survival experiment presented in Fig. 6f. Representative of three independent clonogenic experiments. d, Anti-FLAG immunoblots of whole-cell extracts from RPE1-hTERT p53-/- Cas9 parental (WT) or BRCA1-/- cells treated with AS1 for 72 h. Histone H2A is used as a loading control. Representative of two independent immunoblots. e, Clonogenic survival of RPE1-hTERT p53-/- Cas9 parental (WT) or BRCA1-/- cells following expression of B6L for the indicated periods of time prior to replating without AS1. Data presented as mean ± S.D. n = 3 clonogenic assays; analyzed with two-way ANOVA with repeated measures and Šídák’s multiple comparisons. n.s.=0.99907, p = 0.0076 (d4) for WT and BRCA1-/- respectively. f, Additional representative micrographs from the experiment shown in Fig. 7c. Left panel shows a normal anaphase cell; right panel shows a cell with a lagging centromere-positive chromosome. All scale bars = 10 µm. Statistical analyses and full blots are found in Source Data ED Fig. 7.

Source data

Extended Data Fig. 8 The CIP2A-TOPBP1 interaction does not impact ATR activation but promotes viability of MDA-MB-436 cells.

a, Immunoblots assessing CHK1 S345 phosphorylation in DLD1 cells transduced with either an empty virus (EV) that expresses the DD domain alone or a virus expressing DD-tagged B6L following induction with 1 μM AS1. Cells were treated with hydroxyurea (HU) at the indicated doses for 24 h before harvesting. Representative of two independent transductions. b, Immunoblots assessing CHK1 S345 and RPA2 pS33 phosphorylation in whole-cell lysates of RPE1-hTERT p53-/- Cas9 parental (WT) or independent CIP2A-/- clones. Cells were treated with camptothecin (CPT; 1 μM) for the indicated times prior to harvesting. Representative of two independent immunoblots. c, Anti-FLAG immunoblots of whole-cell extracts from MDA-MB-436 cells treated with Aqua-Shield-1 (AS1) for 72 h. KAP1 is used as loading control. Representative of three independent immunoblots. d, Proliferation curves for MDA-MB-436 cells upon B6L stabilization by AS1 treatment (1 μM). Cells were transduced with an empty virus (EV) as control. Representative of three proliferation assays, analyzed with a Mann-Whitney test; U value (normal approximation)= -5.612486, -5.163978. e, Quantitation of independent proliferation experiments described in (d). Data presented as the mean values at endpoint ± S.D. n = 3 proliferation assays; analyzed with a t-test, assuming data is normally distributed, p = 0.0038. f, Quantitation of micronuclei (MNi)-positive cells in MDA-MB-436 cells transduced with an either empty virus (EV) or B6L-expressing virus following addition of AS1. Data presented as the mean ± S.D. n = 3 independent MNi counts; analyzed with two-way ANOVA with repeated measures and Šídák’s multiple comparisons testing. p = 0.00015. Statistical analyses and full blots are found in Source Data ED Fig. 8.

Source data

Extended Data Fig. 9 Probing the role of PP2A inhibition by CIP2A in the maintenance of genome integrity in BRCA-deficient cells.

a, Immunoblots showing levels of B56α (left), B56γ (middle) and of B6L and the destabilization domain (DD) from the empty virus (EV) (right) after 72 h of siRNA treatment and 48 h of AS1 induction. Representative of two independent immunostainings. b, Quantitation of micronuclei (MNi) in DLD1 BRCA2-/- cells stably transduced with a B6L-encoding lentivirus and transfected with control (siCTRL) or B56α and B56γ-targeting siRNAs. B6L expression was induced by AS1 addition for 48 h. Data presented as the mean ± S.D. n = 3 independent MNi counts; analyzed with a two-tailed t-test, assuming data is normally distributed, **p = 0.0031, **p = 0.0072. c, Representative micrographs of the experiment in (b). Arrowheads indicate micronuclei. d, LacR/LacO assay assessing the interaction between endogenous CIP2A, HA-tagged B56α or B56γ and FLAG-LacR-TOPBP1. Data presented as the intensity of the signal at the LacR-TOPBP1 focus over background (dashed line). Bars represent the mean ± S.D. aggregated from two independent localization assays measuring 15 cells per experiment. e, Representative micrographs of the experiment in (d). Statistical analyses and full blots are found in Source Data ED Fig. 9.

Source data

Extended Data Fig. 10 Data supporting proof-of-concept disruption of CIP2A-TOPBP1 interaction.

a, Determination of the EC50 growth inhibitory concentration of AS1 in DLD1 BRCA2-/- cells transduced with lentivirus expressing either the DD domain alone (EV) or B6L. Data is presented as the mean ± SEM. n = 3 technical replicates; analyzed with extra sum-of-squares F test. ***p = 1 × 10−15. b, Pharmacokinetic analysis of AS1 free plasma concentration over a 24 h period in mice. Data is presented as the mean of values from n = 3 animals. c, Analysis of cell viability by CellTiter GLo of the indicated DLD1 cell lines following celastrol (left) or TD-52 (right) treatments at the indicated concentrations. Data presented as the mean values ± S.D. n = 3 independent drug sensitivity assays; analyzed with extra sum-of-squares F test. Statistical analyses are found in Source Data ED Fig. 10.

Source data

Supplementary information

Source data

Source data Fig. 1

Numerical and statistical source data.

Source data Fig. 2

Numerical and statistical source data.

Source data Fig. 3

Numerical and statistical source data.

Source data Fig. 4

Numerical and statistical source data.

Source data Fig. 5

Numerical and statistical source data.

Source data Fig. 5

Unprocessed western blots.

Source data Fig. 6

Numerical and statistical source data.

Source data Fig. 7

Numerical and statistical source data.

Source data Fig. 8

Numerical and statistical source data.

Source data Extended data Fig. 1

Numerical and statistical source data.

Source data Extended data Fig. 1

Unprocessed western blots.

Source data Extended data Fig. 2

Numerical and statistical source data.

Source data Extended data Fig. 3

Numerical and statistical source data.

Source data Extended data Fig. 4

Numerical and statistical source data.

Source data Extended data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 5

Numerical and statistical source data.

Source data Extended data Fig. 5

Unprocessed western blots.

Source data Extended data Fig. 6

Unprocessed western blots.

Source data Extended data Fig. 7

Numerical and statistical source data.

Source data Extended data Fig. 7

Unprocessed western blots.

Source data Extended data Fig. 8

Numerical and statistical source data.

Source data Extended data Fig. 8

Unprocessed western blots.

Source data Extended data Fig. 9

Numerical and statistical source data.

Source data Extended data Fig. 9

Unprocessed western blots.

Source data Extended data Fig. 10

Numerical and statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adam, S., Rossi, S.E., Moatti, N. et al. The CIP2A–TOPBP1 axis safeguards chromosome stability and is a synthetic lethal target for BRCA-mutated cancer. Nat Cancer 2, 1357–1371 (2021). https://doi.org/10.1038/s43018-021-00266-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s43018-021-00266-w

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