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:

Regulation of ALT-associated homology-directed repair by polyADP-ribosylation

Abstract

The synthesis of poly(ADP-ribose) (PAR) reconfigures the local chromatin environment and recruits DNA-repair complexes to damaged chromatin. PAR degradation by poly(ADP-ribose) glycohydrolase (PARG) is essential for progression and completion of DNA repair. Here, we show that inhibition of PARG disrupts homology-directed repair (HDR) mechanisms that underpin alternative lengthening of telomeres (ALT). Proteomic analyses uncover a new role for poly(ADP-ribosyl)ation (PARylation) in regulating the chromatin-assembly factor HIRA in ALT cancer cells. We show that HIRA is enriched at telomeres during the G2 phase and is required for histone H3.3 deposition and telomere DNA synthesis. Depletion of HIRA elicits systemic death of ALT cancer cells that is mitigated by re-expression of ATRX, a protein that is frequently inactivated in ALT tumors. We propose that PARylation enables HIRA to fulfill its essential role in the adaptive response to ATRX deficiency that pervades ALT 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: Deregulation of PAR metabolism alters ALT activity.
Fig. 2: Disruption of PAR metabolism interferes with telomeric HDR.
Fig. 3: Identification of the PAR-regulated ALT proteome.
Fig. 4: Consequences of HIRA depletion on ALT activity.
Fig. 5: HIRA compensates for ATRX loss, and its depletion is synthetic lethal with ATRX loss.
Fig. 6: PAR-dependent regulation of HIRA-dependent chromatin assembly.

Similar content being viewed by others

Data availability

Original immunofluorescence and colony-formation images are separately available in the Figshare data depository (https://figshare.com/s/17f00c7faa765b329c22). Proteomics data has been deposited at ProteomeXchange Consortium via the PRIDE partner repository with the dataset identifier PXD020243. Source data are provided with this paper.

References

  1. Gupte, R., Liu, Z. & Kraus, W. L. PARPs and ADP-ribosylation: recent advances linking molecular functions to biological outcomes. Genes Dev. 31, 101–126 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Slade, D. et al. The structure and catalytic mechanism of a poly(ADP-ribose) glycohydrolase. Nature 477, 616–620 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Koh, D. W. et al. Failure to degrade poly(ADP-ribose) causes increased sensitivity to cytotoxicity and early embryonic lethality. Proc. Natl Acad. Sci. USA 101, 17699–17704 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Hanzlikova, H. et al. The importance of poly(ADP-ribose) polymerase as a sensor of unligated Okazaki fragments during DNA replication. Mol. Cell 71, 319–331.e3 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Maya-Mendoza, A. et al. High speed of fork progression induces DNA replication stress and genomic instability. Nature 559, 279–284 (2018).

    Article  CAS  PubMed  Google Scholar 

  6. García-Expósito, L. et al. Proteomic profiling reveals a specific role for translesion DNA polymerase η in the alternative lengthening of telomeres. Cell Rep. 17, 1858–1871 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  7. Schmutz, I., Timashev, L., Xie, W., Patel, D. J. & de Lange, T. TRF2 binds branched DNA to safeguard telomere integrity. Nat. Struct. Mol. Biol. 24, 734–742 (2017).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Doksani, Y. & de Lange, T. Telomere-internal double-strand breaks are repaired by homologous recombination and PARP1/Lig3-dependent end-joining. Cell Rep. 17, 1646–1656 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Dunham, M. A., Neumann, A. A., Fasching, C. L. & Reddel, R. R. Telomere maintenance by recombination in human cells. Nat. Genet. 26, 447–450 (2000).

    Article  CAS  PubMed  Google Scholar 

  11. Dilley, R. L. et al. Break-induced telomere synthesis underlies alternative telomere maintenance. Nature 539, 54–58 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Roumelioti, F.-M. et al. Alternative lengthening of human telomeres is a conservative DNA replication process with features of break-induced replication. EMBO Rep. 17, 1731–1737 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. James, D. I. et al. First-in-class chemical probes against poly(ADP-ribose) glycohydrolase (PARG) inhibit DNA repair with differential pharmacology to olaparib. ACS Chem. Biol. 11, 3179–3190 (2016).

    Article  CAS  PubMed  Google Scholar 

  14. Yeager, T. R. et al. Telomerase-negative immortalized human cells contain a novel type of promyelocytic leukemia (PML) body. Cancer Res. 59, 4175–4179 (1999).

    CAS  PubMed  Google Scholar 

  15. Londoño-Vallejo, J. A., Der-Sarkissian, H., Cazes, L., Bacchetti, S. & Reddel, R. R. Alternative lengthening of telomeres is characterized by high rates of telomeric exchange. Cancer Res. 64, 2324–2327 (2004).

    Article  PubMed  Google Scholar 

  16. O’Sullivan, R. J. et al. Rapid induction of alternative lengthening of telomeres by depletion of the histone chaperone ASF1. Nat. Struct. Mol. Biol. 21, 167–174 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Berti, M. et al. Human RECQ1 promotes restart of replication forks reversed by DNA topoisomerase I inhibition. Nat. Struct. Mol. Biol. 20, 347–354 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Clynes, D. et al. Suppression of the alternative lengthening of telomere pathway by the chromatin remodelling factor ATRX. Nat. Commun. 6, 7538–11 (2015).

    Article  PubMed  Google Scholar 

  19. Cho, N. W., Dilley, R. L., Lampson, M. A. & Greenberg, R. A. Interchromosomal homology searches drive directional ALT telomere movement and synapsis. Cell 159, 108–121 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Mortusewicz, O., Fouquerel, E., Amé, J.-C., Leonhardt, H. & Schreiber, V. PARG is recruited to DNA damage sites through poly(ADP-ribose)- and PCNA-dependent mechanisms. Nucleic Acids Res. 39, 5045–5056 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Kaufmann, T. et al. A novel non-canonical PIP-box mediates PARG interaction with PCNA. Nucleic Acids Res. 45, 9741–9759 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Zhang, J.-M., Yadav, T., Ouyang, J., Lan, L. & Zou, L. Alternative lengthening of telomeres through two distinct break-induced replication pathways. Cell Rep. 26, 955–968.e3 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Illuzzi, G. et al. PARG is dispensable for recovery from transient replicative stress but required to prevent detrimental accumulation of poly(ADP-ribose) upon prolonged replicative stress. Nucleic Acids Res. 42, 7776–7792 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Gagné, J.-P. et al. Proteome-wide identification of poly(ADP-ribose) binding proteins and poly(ADP-ribose)-associated protein complexes. Nucleic Acids Res. 36, 6959–6976 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Gagné, J.-P. et al. Quantitative proteomics profiling of the poly(ADP-ribose)-related response to genotoxic stress. Nucleic Acids Res. 40, 7788–7805 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Jungmichel, S. et al. Proteome-wide identification of poly(ADP-ribosyl)ation targets in different genotoxic stress responses. Mol. Cell 52, 272–285 (2013).

    Article  CAS  PubMed  Google Scholar 

  27. Karras, G. I. et al. The macro domain is an ADP-ribose binding module. EMBO J. 24, 1911–1920 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ahel, D. et al. Poly(ADP-ribose)-dependent regulation of DNA repair by the chromatin remodeling enzyme ALC1. Science 325, 1240–1243 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Gomez, M. et al. PARP1 is a TRF2-associated poly(ADP-ribose)polymerase and protects eroded telomeres. Mol. Biol. Cell 17, 1686–1696 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Masson, M. et al. XRCC1 is specifically associated with poly(ADP-ribose) polymerase and negatively regulates its activity following DNA damage. Mol. Cell. Biol. 18, 3563–3571 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Altmeyer, M. et al. The chromatin scaffold protein SAFB1 renders chromatin permissive for DNA damage signaling. Mol. Cell 52, 206–220 (2013).

    Article  CAS  PubMed  Google Scholar 

  32. Acs, K. et al. The AAA-ATPase VCP/p97 promotes 53BP1 recruitment by removing L3MBTL1 from DNA double-strand breaks. Nat. Struct. Mol. Biol. 18, 1345–1350 (2011).

    Article  CAS  PubMed  Google Scholar 

  33. Flynn, R. L. et al. TERRA and hnRNPA1 orchestrate an RPA-to-POT1 switch on telomeric single-stranded DNA. Nature 471, 532–536 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Gagné, J.-P., Hunter, J. M., Labrecque, B., Chabot, B. & Poirier, G. G. A proteomic approach to the identification of heterogeneous nuclear ribonucleoproteins as a new family of poly(ADP-ribose)-binding proteins. Biochem. J. 371, 331–340 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Polo, S. E. et al. Regulation of DNA-end resection by hnRNPU-like proteins promotes DNA double-strand break signaling and repair. Mol. Cell 45, 505–516 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Adamson, B., Smogorzewska, A., Sigoillot, F. D., King, R. W. & Elledge, S. J. A genome-wide homologous recombination screen identifies the RNA-binding protein RBMX as a component of the DNA-damage response. Nat. Cell Biol. 14, 318–328 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Altmeyer, M. et al. Liquid demixing of intrinsically disordered proteins is seeded by poly(ADP-ribose). Nat. Commun. 6, 8088–12 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Singatulina, A. S. et al. PARP-1 activation directs FUS to DNA damage sites to form PARG-reversible compartments enriched in damaged DNA. Cell Rep. 27, 1809–1821.e5 (2019).

    Article  CAS  PubMed  Google Scholar 

  39. Schrank, B. R. et al. Nuclear ARP2/3 drives DNA break clustering for homology-directed repair. Nature 559, 61–66 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Tagami, H., Ray-Gallet, D., Almouzni, G. & Nakatani, Y. Histone H3.1 and H3.3 complexes mediate nucleosome assembly pathways dependent or independent of DNA synthesis. Cell 116, 51–61 (2004).

    Article  CAS  PubMed  Google Scholar 

  41. Goldberg, A. D. et al. Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Pchelintsev, N. A. et al. Placing the HIRA histone chaperone complex in the chromatin landscape. Cell Rep. 3, 1012–1019 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Adam, S., Polo, S. E. & Almouzni, G. Transcription recovery after DNA damage requires chromatin priming by the H3.3 histone chaperone HIRA. Cell 155, 94–106 (2013).

    Article  CAS  PubMed  Google Scholar 

  44. Lewis, P. W., Elsaesser, S. J., Noh, K.-M., Stadler, S. C. & Allis, C. D. Daxx is an H3.3-specific histone chaperone and cooperates with ATRX in replication-independent chromatin assembly at telomeres. Proc. Natl Acad. Sci. USA 107, 14075–14080 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Juhász, S., Elbakry, A., Mathes, A. & Löbrich, M. ATRX promotes DNA repair synthesis and sister chromatid exchange during homologous recombination. Mol. Cell 71, 11–24.e7 (2018).

    Article  PubMed  CAS  Google Scholar 

  46. Heaphy, C. M. et al. Altered telomeres in tumors with ATRX and DAXX mutations. Science 333, 425–425 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Lovejoy, C. A. et al. Loss of ATRX, genome instability, and an altered DNA damage response are hallmarks of the alternative lengthening of telomeres pathway. PLoS Genet. 8, e1002772 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Zhang, H. et al. RPA interacts with HIRA and regulates H3.3 deposition at gene regulatory elements in mammalian cells. Mol. Cell 65, 272–284 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Ray-Gallet, D. et al. Dynamics of histone H3 deposition in vivo reveal a nucleosome gap-filling mechanism for H3.3 to maintain chromatin integrity. Mol. Cell 44, 928–941 (2011).

    Article  CAS  PubMed  Google Scholar 

  50. Elsässer, S. J. et al. DAXX envelops a histone H3.3-H4 dimer for H3.3-specific recognition. Nature 491, 560–565 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Episkopou, H. et al. Alternative lengthening of telomeres is characterized by reduced compaction of telomeric chromatin. Nucleic Acids Res. 42, 4391–4405 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Tang, Y. et al. Structure of a human ASF1a–HIRA complex and insights into specificity of histone chaperone complex assembly. Nat. Struct. Mol. Biol. 13, 921–929 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Ask, K. et al. Codanin-1, mutated in the anaemic disease CDAI, regulates Asf1 function in S-phase histone supply. EMBO J. 31, 2013–2023 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We are indebted to M. Lund Nielsen for sharing detailed protocols and advice in relation to enrichment of PARylated proteins with AF1521. We are grateful to M. Altmeyer, D. Clynes, R. Greenberg, B. Johnson, K. Miller, S. Oesterreich, S. Polo, D. Slade, H. van Attikum and S. Smith for sharing high quality reagents. We thank G. Delgoffe, A. Menk and J. Stewart-Ornstein for guidance with experiments and sharing resources. Resources and facilities at the UPMC Hillman Cancer Center were supported by Comprehensive Cancer Center Support Grant NCI/no. P30CA047904. Research funding was provided to individual investigators from the following agencies; R.J.O. NCI/no. 5R01CA207209-02 and American Cancer Society no. RSG-18-038-01-DMC; S.C.W. NIH/no. 1S10OD019973-O1; A.I.N. NIH U24CA210967 and R01 GM094231; R.W.S. NIH/no. R01CA148629 and NIEHS/no. R01ES014811. N.K. is supported by a T32 training grant, NIGMS/no. T32GM008424-25. This work was also supported by grants C480/A11411 and C5759/A17098 from Cancer Research UK to D.I.J., I.D.W., K.M.S and D.O.; and ERC-2015-ADG- 694694 ‘ChromADICT’, ANR-16-CE15-0018 ‘CHRODYT’, ANR-16-CE11-0028 ‘REPLICAF’, ANR-16-CE12-0024 ‘CHIFT’, ITN-765966 Curie ‘EPISYSTEM’ and ITN-813327 Curie ‘CHROMDESIGN’ to G.A. and D.R.G.

Author information

Authors and Affiliations

Authors

Contributions

R.J.O. conceived and designed the study; S.M.H. designed and conducted all experimentation; R.J.O. and S.M.H analyzed the data. N.K., R.B., M.L.L., J.L.R., J.B.G., L.G.-E., A.R.W. and R.J.O. assisted with experimentation and analyzed data. C.T.W. and S.C.W. assisted with optical-imaging experiment design, optimization and analyzes. D.I.J., I.D.W., K.M.S. and D.O. produced, validated and provided the PARG inhibitors. F.V.L., D.M. and A.I.N. generated and analyzed proteomic data. J.L. and R.W.S. provided essential reagents. D.R.-G. and G.A. provided essential reagents, assisted with experimental design and interpretation of results. S.M.H. and R.J.O. wrote the manuscript with input from coauthors.

Corresponding author

Correspondence to Roderick J. O’Sullivan.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Beth Moorefield was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 Disruption of PAR turnover perturbs recombinogenic activity at ALT telomeres.

a, Western blot of PARP1 or PARG knockdown and PAR accumulation in U2OS cells expressing shRNAs. Extracts from 2 mM hydroxyurea (HU)-PARGi (5μM/24hrs) treated U2OS cells serve as a control for PAR induction. b, Representative IF images of APBs (PML-TTAGGG) in U2OS cells expressing the indicated shRNAs. c, Quantification of APBs (% positive cells) in U2OS, Saos2 and HeLa LT cells expressing the indicated shRNAs. d, Quantification of telomere sister chromatid exchanges (t-SCE) (% per metaphase) in U2OS, Saos2 and HeLa LT cells expressing the indicated shRNAs. n refers to the number of metaphase spreads analyzed from N = 3 independent assays. e, Left: Western blot of U2OS cells transfected with non-targeting and/or RECQ1 siRNAs treated with PARPi (5μM/72hrs). Right: Quantification of t-SCEs in U2OS and HeLa LT cells transfected with NT siRNA, and/or RECQ1 siRNA. DMSO/PARPi (5μM/24hrs). n refers to the number of metaphase spreads analyzed from N = 4 independent assays. f, Cell cycle profiles and (g) cellular [ADP/ATP] ratio of U2OS and Hela LT cells treated with inhibitors for 12 days. h, PFGE of DMSO, PARPi (100 nM) or PARGi (1 μM) treated VA13 cells. The red dot indicates mean telomere length (kb). i, Representative images and quantification from clonogenic survival assays in ALT + cells, TEL + cells and (j) U2OS cells expressing ATRX treated with DMSO, PARPi (100μM) or PARGi (1μM) for 7 days. All scale bars in IF panels=5μm. All graphed data in the figure are mean ± s.e.m except d, e which are mean ± s.d. Unless otherwise stated, (n) is the number of cells analyzed and the number of independent assays (N) conducted is represented by black circles. Statistical significance was determined using one-way ANOVA except except (d) where the Mann-Whitney test was used and (i-j) where Students t test was used. Uncropped blots for a, e and digital images are deposited on Figshare. Graphed data is available as Source Data.

Source data

Extended Data Fig. 2 PARylation is an early and direct mediator of TRF1-FokI DSB formation.

a, [ADP/ATP] ratio in DMSO/PARGi treated WT-TRF1-FokI U2OS cells. Cells were treated with 1.5 mM/1 hr MMS. (b) Representative IF images and quantification of PAR at WT-TRF1-FokI DSBs after PARGi, PARGi-PARPi or TNKS1 knockdown. c, Representative IF images and quantification showing GFP-PARP1 localization in WT-TRF1-FokI cells treated with PARPi, PARGi or both. d, Representative IF images and quantification showing GFP-PARG localization in WT-TRF1-FokI U2OS cells. e, Left: Representative IF images and quantification of telomere foci size per cell in VA13 and Hela LT cells transfected with WT-TRF1-FokI from N = 2 independent assays. f, Representative stills of telomere (eGFP-TRF1) movement in U2OS cells treated with DMSO, PARPi or PARGi. Graph displays the cumulative Mean Squared Displacement (MSD) of 100 telomeres. g, Top: Schematic of DNA combing in G2-synchronized WT-TRF1-FokI cells treated with DMSO, PARPi, PARGi, or co-treated with PARPi and PARGi. Left: Quantification of telomeric fiber length of combined pulses. Right: Violin plot analysis of fork velocity. h, Graphs of CldU/IdU tract distribution of telomeric fibers in inhibitor treated U2OS-TRF1-FokI cells. n refers to the number of fibers containing TTAGGG signals analyzed from N = 2 independent assays. i, Representative IF images and quantification of BrdU synthesis at telomeres in the indicated cell lines after transfection with WT-TRF1-FokI and treated with PARGi or PARPi. j, Representative IF images and quantification of PCNA and (k) POLD3 localization at WT-TRF1-FokI telomeres treated with treated with DMSO, PARPi, PARGi, PARGi-Me or PARPi-PARGi. All inhibitor treatments, 5μM/4hrs unless otherwise indicated. All scale bars in IF panels=5μm. All graphed data in the figure are mean ± s.e.m. Unless otherwise stated, (n) is the number of cells analyzed and the number of independent assays (N) conducted is represented by black circles. Statistical significance was determined using one-way ANOVA. Digital images are deposited on Figshare. Graphed data is available as Source Data.

Source data

Extended Data Fig. 3 PAR-dependent recruitment of proteins to TRF1-FokI induced telomeric DSBs.

a, Spectral counts for the indicated proteins that were identified by Af1521-PAR proteomics. b, Western blot analysis illustrating the expression of GFP fusion proteins in U2OS WT-TRF1-FokI cells. GFP antibody was used to blot for protein expression in each treatment. c, Representative IF images showing the localization of the indicated GFP fusion proteins in WT-TRF1-FokI U2OS cells following treatment with DMSO, PARPi and combined PARGi/PARPi. d, In vivo PARylation assay with GFP tagged FUS, RBMX and ARP3. * indicates the band corresponding to the immunoprecipitated GFP-tagged target protein on PAR blots. e, Western blotting was performed with antibodies to validate siRNA knockdown of endogenous protein hits from Af1521-PAR proteomics in U2OS cells. f, Cell cycle profile of U2OS and VA13 cells after siRNA knockdown of protein hits from Af1521-PAR proteomics. All inhibitor treatments, 5μM/4hrs. All scale bars in IF panels=5μm. Uncropped blots for b, d-e and digital images are deposited on Figshare. Graphed data is available as Source Data.

Source data

Extended Data Fig. 4 Selectivity of HIRA for localization to telomeres in ALT cancer cells telomeric DSBs is independent of RPA and is necessary for telomere DNA synthesis.

a, Representative IF images of HIRA-YFP localization in ALT + and TEL + cell lines treated with DMSO/PARGi. b, Representative IF images of HIRA-YFP localization in U2OS cells after exposure to 30 J/m2ultra-violet C (UV-C) and 10 Gy ionizing irradiation (γIR). 5 μM PARGi was added for 30 mins following irradiation. c, Left: Western blot validation of RPA70 knockdown in U2OS cells. Middle: Representative IF images of HIRA-YFP localization at telomeres in U2OS cells after RPA70 knockdown. Right: Quantification of HIRA-YFP localization to telomeres in indicated conditions from N = 2 independent assays. d, Western blot validation of HIRA, CABIN1 and UBN1 siRNA knockdown in U2OS cells. e, Graphs of CldU/IdU tract distribution of >30 telomeric fibers in NT siRNA and HIRA siRNA transfected U2OS-TRF1-FokI cells. f, Representative IF images and quantification of BrdU synthesis at telomeres in indicated cell lines that are transfected with WT-TRF1-FokI and HIRA siRNA. All inhibitor treatments, 5 μM/4 hrs. All scale bars in IF panels=5 μm. Unless otherwise stated, (n) is the number of cells analyzed and the number of independent assays (N) conducted is represented by black circles. Uncropped blots for c-d and digital images are deposited on Figshare. Graphed data is available as Source Data.

Source data

Extended Data Fig. 5 HIRA compensates for and forms a synthetic lethal interaction with ATRX loss.

a, Representative IF images and quantification of HIRA-YFP localization in Hela LT cells transfected with ATRX siRNA and WT-TRF1-FokI, as well as treated with PARGi (5 μM, from N = 2 independent assays. b, Western blot validation of HIRA knockdown in ALT + and TEL + cell lines using two different shRNA sequences (#1 and #2). c, Western blot validation of HIRA and ATRX knockdown in TEL + HeLa LT cell lines using the indicated shRNAs. d, Representative images and quantification of clonogenic survival assays with Hela LT cell line stably expressing scrambled non-targeting (NT), ATRX (#A and #B), and HIRA (#1 and #2) shRNAs for 5 days. All inhibitor treatments, 5μM/4hrs. All scale bars in IF panels=5μm. Unless otherwise stated, (n) is the number of cells analyzed and the number of independent assays (N) conducted is represented by black circles. Uncropped blots for c-d and digital images are deposited on Figshare. Graphed data is available as Source Data.

Source data

Extended Data Fig. 6 Analysis of potential binding of PAR by HIRA.

a, GFP-PARP1, YFP-HIRA or GFP-FUS were transiently transfected in U2OS cells. Immunoprecipitated and blotted GFP-fusion proteins were incubated with biotinylated PAR and detected using an anti-PAR (10H) antibody. GFP fusion proteins were detected with HRP conjugated GFP antibody. b, Western blot showing depletion of endogenous HIRA and complementation with the indicated HIRA constructs in U2OS cells. Uncropped blot images for panel a,b are shown in Supplementary Fig. 1.

Supplementary information

Supplementary Information

Supplementary Figure 1.

Reporting Summary

Supplementary Video 1

Telomere mobility in ALT U2OS cells treated with DMSO.

Supplementary Video 2

Telomere mobility in ALT U2OS cells treated with PARPi.

Supplementary Video 3

Telomere mobility in ALT U2OS cells treated with PARGi.

Supplementary Table 1

Listing of proteins identified by mass spectrometry.

Source data

Source Data Fig. 1

Source data for graphs

Source Data Fig. 2

Source data for graphs

Source Data Fig. 3

Source data for graphs

Source Data Fig. 4

Source data for graphs

Source Data Fig. 5

Source data for graphs

Source Data Fig. 6

Source data for graphs

Source Data Extended Data Fig. 1

Source data for graphs

Source Data Extended Data Fig. 2

Source data for graphs

Source Data Extended Data Fig. 3

Source data for graphs

Source Data Extended Data Fig. 4

Source data for graphs

Source Data Extended Data Fig. 5

Source data for graphs

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hoang, S.M., Kaminski, N., Bhargava, R. et al. Regulation of ALT-associated homology-directed repair by polyADP-ribosylation. Nat Struct Mol Biol 27, 1152–1164 (2020). https://doi.org/10.1038/s41594-020-0512-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-020-0512-7

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