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:

Break-induced telomere synthesis underlies alternative telomere maintenance

Abstract

Homology-directed DNA repair is essential for genome maintenance through templated DNA synthesis. Alternative lengthening of telomeres (ALT) necessitates homology-directed DNA repair to maintain telomeres in about 10–15% of human cancers. How DNA damage induces assembly and execution of a DNA replication complex (break-induced replisome) at telomeres or elsewhere in the mammalian genome is poorly understood. Here we define break-induced telomere synthesis and demonstrate that it utilizes a specialized replisome, which underlies ALT telomere maintenance. DNA double-strand breaks enact nascent telomere synthesis by long-tract unidirectional replication. Proliferating cell nuclear antigen (PCNA) loading by replication factor C (RFC) acts as the initial sensor of telomere damage to establish predominance of DNA polymerase δ (Pol δ) through its POLD3 subunit. Break-induced telomere synthesis requires the RFC–PCNA–Pol δ axis, but is independent of other canonical replisome components, ATM and ATR, or the homologous recombination protein Rad51. Thus, the inception of telomere damage recognition by the break-induced replisome orchestrates homology-directed telomere maintenance.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Break-induced telomere synthesis occurs by long-tract unidirectional telomeric recombination.
Figure 2: Break-induced telomere synthesis occurs by alternative HDR and utilizes a non-canonical replisome defined by Pol δ.
Figure 3: Rapid loading of PCNA acts as the initial sensor of telomere damage.
Figure 4: POLD3 is critical for telomere maintenance in ALT-dependent cells.

Similar content being viewed by others

References

  1. Ciccia, A. & Elledge, S. J. The DNA damage response: making it safe to play with knives. Mol. Cell 40, 179–204 (2010)

    Article  CAS  Google Scholar 

  2. Bryan, T. M., Englezou, A., Gupta, J., Bacchetti, S. & Reddel, R. R. Telomere elongation in immortal human cells without detectable telomerase activity. EMBO J. 14, 4240–4248 (1995)

    Article  CAS  Google Scholar 

  3. Cesare, A. J. & Reddel, R. R. Alternative lengthening of telomeres: models, mechanisms and implications. Nat. Rev. Genet. 11, 319–330 (2010)

    Article  CAS  Google Scholar 

  4. Norio, P. & Schildkraut, C. Visualization of DNA replication on individual Epstein-Barr virus episomes. Science 294, 2361–2365 (2001)

    Article  ADS  CAS  Google Scholar 

  5. Sfeir, A. et al. Mammalian telomeres resemble fragile sites and require TRF1 for efficient replication. Cell 138, 90–103 (2009)

    Article  CAS  Google Scholar 

  6. Nabetani, A., Yokoyama, O. & Ishikawa, F. Localization of hRad9, hHus1, hRad1, and hRad17 and caffeine-sensitive DNA replication at the alternative lengthening of telomeres-associated promyelocytic leukemia body. J. Biol. Chem. 279, 25849–25857 (2004)

    Article  CAS  Google Scholar 

  7. 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  Google Scholar 

  8. 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  Google Scholar 

  9. Lundblad, V. & Blackburn, E. H. An alternative pathway for yeast telomere maintenance rescues est1- senescence. Cell 73, 347–360 (1993)

    Article  CAS  Google Scholar 

  10. Teng, S. C. & Zakian, V. A. Telomere-telomere recombination is an efficient bypass pathway for telomere maintenance in Saccharomyces cerevisiae. Mol. Cell. Biol. 19, 8083–8093 (1999)

    Article  CAS  Google Scholar 

  11. Le, S., Moore, J. K., Haber, J. E. & Greider, C. W. RAD50 and RAD51 define two pathways that collaborate to maintain telomeres in the absence of telomerase. Genetics 152, 143–152 (1999)

    Article  CAS  Google Scholar 

  12. Lydeard, J. R., Jain, S., Yamaguchi, M. & Haber, J. E. Break-induced replication and telomerase-independent telomere maintenance require Pol32. Nature 448, 820–823 (2007)

    Article  ADS  CAS  Google Scholar 

  13. Malkova, A., Ivanov, E. L. & Haber, J. E. Double-strand break repair in the absence of RAD51 in yeast: a possible role for break-induced DNA replication. Proc. Natl Acad. Sci. USA 93, 7131–7136 (1996)

    Article  ADS  CAS  Google Scholar 

  14. Chen, Q., Ijpma, A. & Greider, C. W. Two survivor pathways that allow growth in the absence of telomerase are generated by distinct telomere recombination events. Mol. Cell. Biol. 21, 1819–1827 (2001)

    Article  CAS  Google Scholar 

  15. Teng, S. C., Chang, J., McCowan, B. & Zakian, V. A. Telomerase-independent lengthening of yeast telomeres occurs by an abrupt Rad50p-dependent, Rif-inhibited recombinational process. Mol. Cell 6, 947–952 (2000)

    Article  CAS  Google Scholar 

  16. Cox, K. E., Maréchal, A. & Flynn, R. L. SMARCAL1 resolves replication stress at ALT telomeres. Cell Reports 14, 1032–1040 (2016)

    Article  CAS  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  18. Flynn, R. L. et al. Alternative lengthening of telomeres renders cancer cells hypersensitive to ATR inhibitors. Science 347, 273–277 (2015)

    Article  ADS  CAS  Google Scholar 

  19. Lydeard, J. R. et al. Break-induced replication requires all essential DNA replication factors except those specific for pre-RC assembly. Genes Dev. 24, 1133–1144 (2010)

    Article  CAS  Google Scholar 

  20. Johnson, R. E., Prakash, L. & Prakash, S. Pol31 and Pol32 subunits of yeast DNA polymerase δ are also essential subunits of DNA polymerase ζ. Proc. Natl Acad. Sci. USA 109, 12455–12460 (2012)

    Article  ADS  CAS  Google Scholar 

  21. Baranovskiy, A. G. et al. DNA polymerase δ and ζ switch by sharing accessory subunits of DNA polymerase δ. J. Biol. Chem. 287, 17281–17287 (2012)

    Article  CAS  Google Scholar 

  22. Makarova, A. V., Stodola, J. L. & Burgers, P. M. A four-subunit DNA polymerase ζ complex containing Pol δ accessory subunits is essential for PCNA-mediated mutagenesis. Nucleic Acids Res. 40, 11618–11626 (2012)

    Article  CAS  Google Scholar 

  23. Ducoux, M. et al. Mediation of proliferating cell nuclear antigen (PCNA)-dependent DNA replication through a conserved p21Cip1-like PCNA-binding motif present in the third subunit of human DNA polymerase δ. J. Biol. Chem. 276, 49258–49266 (2001)

    Article  CAS  Google Scholar 

  24. Chilkova, O. et al. The eukaryotic leading and lagging strand DNA polymerases are loaded onto primer-ends via separate mechanisms but have comparable processivity in the presence of PCNA. Nucleic Acids Res. 35, 6588–6597 (2007)

    Article  CAS  Google Scholar 

  25. Karras, G. I. & Jentsch, S. The RAD6 DNA damage tolerance pathway operates uncoupled from the replication fork and is functional beyond S phase. Cell 141, 255–267 (2010)

    Article  CAS  Google Scholar 

  26. Murga, M. et al. POLD3 is haploinsufficient for DNA replication in mice. Mol. Cell 63, 877–883 (2016)

    Article  CAS  Google Scholar 

  27. Zhong, Z.-H. et al. Disruption of telomere maintenance by depletion of the MRE11/RAD50/NBS1 complex in cells that use alternative lengthening of telomeres. J. Biol. Chem. 282, 29314–29322 (2007)

    Article  CAS  Google Scholar 

  28. Henson, J. D. et al. DNA C-circles are specific and quantifiable markers of alternative-lengthening-of-telomeres activity. Nat. Biotechnol. 27, 1181–1185 (2009)

    Article  CAS  Google Scholar 

  29. Costantino, L. et al. Break-induced replication repair of damaged forks induces genomic duplications in human cells. Science 343, 88–91 (2014)

    Article  ADS  CAS  Google Scholar 

  30. Mayle, R. et al. Mus81 and converging forks limit the mutagenicity of replication fork breakage. Science 349, 742–747 (2015)

    Article  CAS  Google Scholar 

  31. Minocherhomji, S. et al. Replication stress activates DNA repair synthesis in mitosis. Nature 528, 286–290 (2015)

    Article  ADS  CAS  Google Scholar 

  32. Miyabe, I. et al. Polymerase δ replicates both strands after homologous recombination-dependent fork restart. Nat. Struct. Mol. Biol. 22, 932–938 (2015)

    Article  CAS  Google Scholar 

  33. Verma, P. & Greenberg, R. A. Noncanonical views of homology-directed DNA repair. Genes Dev. 30, 1138–1154 (2016)

    Article  CAS  Google Scholar 

  34. Viggiani, C. J., Knott, S. R. V. & Aparicio, O. M. Genome-wide analysis of DNA synthesis by BrdU immunoprecipitation on tiling microarrays (BrdU-IP-chip) in Saccharomyces cerevisiae. Cold Spring Harb. Protoc. 2010, pdb.prot5385 (2010)

    Article  Google Scholar 

  35. Göhring, J., Fulcher, N., Jacak, J. & Riha, K. TeloTool: a new tool for telomere length measurement from terminal restriction fragment analysis with improved probe intensity correction. Nucleic Acids Res. 42, e21 (2014)

    Article  Google Scholar 

  36. Tang, J. et al. Acetylation limits 53BP1 association with damaged chromatin to promote homologous recombination. Nat. Struct. Mol. Biol. 20, 317–325 (2013)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank A. Sfeir and A. Phillips (NYU) for guidance on telomere SMARD experiments and members of the Greenberg laboratory for critical discussion. This work was supported by NIH grants GM101149, CA138835, and CA17494 to R.A.G., who is also supported by funds from the Abramson Family Cancer Research Institute and Basser Research Center for BRCA. R.L.D. was supported by NIH grants T32GM007170 and T32GM008216.

Author information

Authors and Affiliations

Authors

Contributions

R.L.D., P.V., N.W.C., and R.A.G. designed the study. R.L.D. performed most of the experiments, with assistance from H.D.W. and A.R.W. P.V. conducted SMARD experiments. N.W.C. conducted ATR and Hop2 experiments. R.L.D., P.V., and R.A.G. wrote the manuscript.

Corresponding author

Correspondence to Roger A. Greenberg.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 An inducible system for studying break-induced telomere synthesis.

a, Schematic of inducible TRF1–FokI system. bd, Characterization of U2OS inducible TRF1–FokI system by western blot (b), immunofluorescence (c), and telomere ChIP (d). e, Agarose gel of sonicated DNA prepared for BrdU pulldown. f, g, BrdU pulldown dot blot for telomere content (f) from asynchronous or G2-enriched U2OS cells induced (Ind) with TRF1–FokI for 2 h, with cell-cycle profiles by propidium iodide staining (g). Images were captured at 60× magnification. Dox, doxycycline; S, Shield-1; T, 4-hydroxytamoxifen; DD, destabilization domain; ER, oestrogen receptor; rtTA, reverse tetracycline transactivator; TRE3G, tetracycline response element; WT, wild-type; D450A, nuclease-null mutant.

Extended Data Figure 2 Visualization of spontaneous ALT telomere synthesis.

ac, BrdU immunofluorescence assay to visualize spontaneous ALT telomere synthesis, with representative images of VA13 cells (a) and quantification of a panel of ALT and ALT+ cell lines (b) and U2OS cells induced with TRF1–FokI for 2 h (c). d, e, Representative images of BrdU immunofluorescence of metaphases from spontaneous GM847 cells (d) and U2OS induced (+Ind) with TRF1–FokI for 2 h upon release from RO-3306 (e). Images were captured at 60× magnification. Data represent mean ± s.e.m. of three independent experiments. ***P ≤ 0.001.

Extended Data Figure 3 Break-induced telomere synthesis occurs independently of telomere maintenance mechanism.

a, A panel of ALT and ALT+ inducible TRF1–FokI cell lines tested for TRF1–FokI and ATRX expression by western blot and nascent telomere synthesis by BrdU pulldown dot blot for telomere content after induction (Ind) with TRF1–FokI for 2 h. b, c, BrdU pulldown dot blot for telomere content (b) from HeLa 1.3 cells induced (Ind) with TRF1–FokI for 2 h, with quantification (c). Data represent mean ± s.e.m. of two independent experiments. *P ≤ 0.05.

Extended Data Figure 4 Hop2 contributes to telomere clustering but is dispensable for telomere length maintenance.

a–h, CRISPR/Cas9-mediated excision of HOP2 (sgHOP2) in VA13 cells, with western blot of populations (a). Analysis of Hop2 co-localization with telomere foci by IF-FISH (b), telomere focus size by FISH (c), APBs by PML co-localization with telomere foci (d, e), and telomere exchanges by CO-FISH (f) from sgHOP2 #2 population. Analysis of clones (c1–c6) by western blot (g) and TRF pulsed-field gel at ~PD 25 (h). Peak intensity of telomere length is indicated by red dot. Images were captured at 60× magnification. Data represent mean ± s.e.m. of at least two independent experiments. ***P < 0.0005, **P < 0.005, *P < 0.05.

Extended Data Figure 5 Requirements for break-induced and spontaneous ALT telomere synthesis.

a, BrdU pulldown dot blot timecourse for telomere content from U2OS induced (Ind) with TRF1-FokI for indicated times and treated with indicated siRNAs. b, c, BrdU pulldown dot blots for telomere content from U2OS induced (Ind) with TRF1–FokI for 2 h and treated with indicated siRNAs. d, BrdU pulldown dot blot for telomere content from HeLa 1.3 induced (Ind) with TRF1–FokI for 2 h and treated with indicated siRNAs, with quantification. ei, Analysis of spontaneous ALT telomere synthesis using BrdU immunofluorescence from VA13 (eh) and GM847 and LM216J (i) treated with indicated siRNAs. Images were captured at 60× magnification. Data represent mean ± s.e.m. of two (d) or three (ei) independent experiments. ****P ≤ 0.0001, **P ≤ 0.01, *P ≤ 0.05.

Extended Data Figure 6 Pol δ predominates at ALT telomeres.

a, Representative images of replisome components (green) and telomere foci (red) from U2OS induced with TRF1–FokI for 2 h. b, western blot of Pol δ complex from cell lines treated with TRF1–FokI. Asterisk denotes non-specific band. c, d, Quantification of co-localized POLD3 (c) or POLD1 (d) with telomere foci from U2OS induced with TRF1–FokI for 2 h. e, f, Representative images (e) of co-localized RFC1-PCNA-POLD3 (green) and telomere foci (red) from U2OS induced with TRF1–FokI for 2 h, with quantification (f). WT = wild-type, D450A = nuclease-null mutant. Images were captured at 60× magnification. Data represent mean ± s.e.m. of three independent experiments. ****P ≤ 0.0001, ***P ≤ 0.001.

Extended Data Figure 7 POLD3 is critical for telomere maintenance in ALT-dependent cells.

a, b, analysis of transient POLD3 depletion by C-circle dot blot (a) from U2OS and CO-FISH (b) from VA13 (n = 1780 ends for siCtrl, n = 1637 ends for siPOLD3). c, TRF analysis from U2OS populations at ~PD 25. Peak intensity of telomere length is indicated by red dot. d, schematic of U2OS POLD3 CRISPR (sgPOLD3) cloning strategy with western blot. eg, analysis of POLD3 expression from U2OS clones c1–c4 by qPCR (e), POLD1 Co–IP (f), and darker exposure of western blot from Fig. 4a (g). Asterisk denotes non-specific band. h, Quantification of relative telomere content by dot blot from U2OS clones c1–c4. i, Heat map summarizing decreases (blue), increases (red), or no change (white) in telomere maintenance from U2OS clones c1–c4 as compared to U2OS control. j, k, POLD3-reconstituted CRISPR clones analysed for C-circles by dot blot (j) and Pol δ expression by western blot (k). EV, empty vector; WT, reconstituted POLD3. Data represent mean ± s.e.m. of two independent experiments. **P ≤ 0.01, *P ≤ 0.05.

Extended Data Figure 8 Extended analysis of POLD3 CRISPR clones.

a, b, TRF analysis by pulsed-field gel (a) and C-circle dot blot (b) from U2OS POLD3 CRISPR (sgPOLD3) clones with normal POLD3 expression (c5–c9) at ~PD 25. c, U2OS POLD3 CRISPR clones from an independent guide RNA (sgPOLD3 #2) analysed by TRF and western blot at ~PD 25. d, TRF analysis by pulsed-field gel from HeLa 1.3 populations at ~PD 25. e, Schematic of HeLa 1.3 POLD3 CRISPR cloning strategy with western blot. f, TRF analysis by pulsed-field gel from HeLa 1.3 clones c1–c11 at ~PD 25. Peak intensity of telomere length is indicated by red dot.

Extended Data Figure 9 Knockdown efficiencies.

a–n, Western blots of U2OS or VA13 cells treated with indicated siRNAs. Asterisk denotes non-specific band.

Supplementary information

Supplementary Figure 1

This file contains source images of cropped Western blot gels. (PDF 1299 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dilley, R., Verma, P., Cho, N. et al. Break-induced telomere synthesis underlies alternative telomere maintenance. Nature 539, 54–58 (2016). https://doi.org/10.1038/nature20099

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature20099

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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