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:

Telomeric TERB1–TRF1 interaction is crucial for male meiosis

Abstract

During meiotic prophase, the meiosis-specific telomere-binding protein TERB1 regulates chromosome movement required for homologous pairing and recombination by interacting with the telomeric shelterin subunit TRF1. Here, we report the crystal structure of the TRF1-binding motif of human TERB1 in complex with the TRFH domain of TRF1. Notably, specific disruption of the TERB1–TRF1 interaction by a point mutation in the mouse Terb1 gene results in infertility only in males. We find that this mutation causes an arrest in the zygotene–early pachytene stage and mild telomere abnormalities of autosomes but unpaired X and Y chromosomes in pachytene, leading to massive spermatocyte apoptosis. We propose that the loss of telomere structure mediated by the TERB1–TRF1 interaction significantly affects homologous pairing of the telomere-adjacent pseudoautosomal region (PAR) of the X and Y chromosomes in mouse spermatocytes. Our findings uncover a specific mechanism of telomeres that surmounts the unique challenges of mammalian X–Y pairing in meiosis.

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

Figure 1: Structural studies of the interaction between TRF1TRFH and TERB1PBM.
Figure 2: Mutational analysis of the TERB1–TRF1 interaction.
Figure 3: Disruption of the TRF1–TERB1 interaction results in failure of spermatogenesis.
Figure 4: Spermatogenesis process is arrested in Terb1AEA/AEA spermatocytes.
Figure 5: The 647AEA649 mutation of TERB1 results in telomere localization defect of TERB1, autosomal telomere aberrance and unpaired X and Y chromosomes.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Referenced accessions

NCBI Reference Sequence

Protein Data Bank

References

  1. Handel, M.A. & Schimenti, J.C. Genetics of mammalian meiosis: regulation, dynamics and impact on fertility. Nat. Rev. Genet. 11, 124–136 (2010).

    Article  CAS  Google Scholar 

  2. Petronczki, M., Siomos, M.F. & Nasmyth, K. Un ménage à quatre: the molecular biology of chromosome segregation in meiosis. Cell 112, 423–440 (2003).

    Article  CAS  Google Scholar 

  3. Kleckner, N. Meiosis: how could it work? Proc. Natl. Acad. Sci. USA 93, 8167–8174 (1996).

    Article  CAS  Google Scholar 

  4. Keeney, S., Giroux, C.N. & Kleckner, N. Meiosis-specific DNA double-strand breaks are catalyzed by Spo11, a member of a widely conserved protein family. Cell 88, 375–384 (1997).

    Article  CAS  Google Scholar 

  5. San Filippo, J., Sung, P. & Klein, H. Mechanism of eukaryotic homologous recombination. Annu. Rev. Biochem. 77, 229–257 (2008).

    Article  CAS  Google Scholar 

  6. Neale, M.J. & Keeney, S. Clarifying the mechanics of DNA strand exchange in meiotic recombination. Nature 442, 153–158 (2006).

    Article  CAS  Google Scholar 

  7. de Massy, B. Initiation of meiotic recombination: how and where? Conservation and specificities among eukaryotes. Annu. Rev. Genet. 47, 563–599 (2013).

    Article  CAS  Google Scholar 

  8. Baudat, F., Imai, Y. & de Massy, B. Meiotic recombination in mammals: localization and regulation. Nat. Rev. Genet. 14, 794–806 (2013).

    Article  CAS  Google Scholar 

  9. Syrjänen, J.L., Pellegrini, L. & Davies, O.R. A molecular model for the role of SYCP3 in meiotic chromosome organisation. eLife 3, e02963 (2014).

    Article  Google Scholar 

  10. Bolcun-Filas, E. & Schimenti, J.C. Genetics of meiosis and recombination in mice. Int. Rev. Cell Mol. Biol. 298, 179–227 (2012).

    Article  CAS  Google Scholar 

  11. Hirose, Y. et al. Chiasmata promote monopolar attachment of sister chromatids and their co-segregation toward the proper pole during meiosis I. PLoS Genet. 7, e1001329 (2011).

    Article  CAS  Google Scholar 

  12. Bascom-Slack, C.A., Ross, L.O. & Dawson, D.S. Chiasmata, crossovers, and meiotic chromosome segregation. Adv. Genet. 35, 253–284 (1997).

    Article  CAS  Google Scholar 

  13. Boateng, K.A., Bellani, M.A., Gregoretti, I.V., Pratto, F. & Camerini-Otero, R.D. Homologous pairing preceding SPO11-mediated double-strand breaks in mice. Dev. Cell 24, 196–205 (2013).

    Article  CAS  Google Scholar 

  14. Brown, P.W. et al. Meiotic synapsis proceeds from a limited number of subtelomeric sites in the human male. Am. J. Hum. Genet. 77, 556–566 (2005).

    Article  CAS  Google Scholar 

  15. Koszul, R., Kim, K.P., Prentiss, M., Kleckner, N. & Kameoka, S. Meiotic chromosomes move by linkage to dynamic actin cables with transduction of force through the nuclear envelope. Cell 133, 1188–1201 (2008).

    Article  CAS  Google Scholar 

  16. Koszul, R. & Kleckner, N. Dynamic chromosome movements during meiosis: a way to eliminate unwanted connections? Trends Cell Biol. 19, 716–724 (2009).

    Article  CAS  Google Scholar 

  17. Lee, C.Y. et al. Mechanism and regulation of rapid telomere prophase movements in mouse meiotic chromosomes. Cell Rep. 11, 551–563 (2015).

    Article  CAS  Google Scholar 

  18. Hiraoka, Y. & Dernburg, A.F. The SUN rises on meiotic chromosome dynamics. Dev. Cell 17, 598–605 (2009).

    Article  CAS  Google Scholar 

  19. Ding, X. et al. SUN1 is required for telomere attachment to nuclear envelope and gametogenesis in mice. Dev. Cell 12, 863–872 (2007).

    Article  CAS  Google Scholar 

  20. Daniel, K. et al. Mouse CCDC79 (TERB1) is a meiosis-specific telomere associated protein. BMC Cell Biol. 15, 17 (2014).

    Article  Google Scholar 

  21. Shibuya, H., Ishiguro, K. & Watanabe, Y. The TRF1-binding protein TERB1 promotes chromosome movement and telomere rigidity in meiosis. Nat. Cell Biol. 16, 145–156 (2014).

    Article  CAS  Google Scholar 

  22. Shibuya, H. et al. MAJIN links telomeric DNA to the nuclear membrane by exchanging telomere cap. Cell 163, 1252–1266 (2015).

    Article  CAS  Google Scholar 

  23. Viera, A. et al. CDK2 regulates nuclear envelope protein dynamics and telomere attachment in mouse meiotic prophase. J. Cell Sci. 128, 88–99 (2015).

    Article  CAS  Google Scholar 

  24. Tu, Z. et al. Speedy A-Cdk2 binding mediates initial telomere-nuclear envelope attachment during meiotic prophase I independent of Cdk2 activation. Proc. Natl. Acad. Sci. USA 114, 592–597 (2017).

    Article  Google Scholar 

  25. Perry, J., Palmer, S., Gabriel, A. & Ashworth, A. A short pseudoautosomal region in laboratory mice. Genome Res. 11, 1826–1832 (2001).

    Article  CAS  Google Scholar 

  26. Kauppi, L., Jasin, M. & Keeney, S. The tricky path to recombining X and Y chromosomes in meiosis. Ann. NY Acad. Sci. 1267, 18–23 (2012).

    Article  CAS  Google Scholar 

  27. Kauppi, L. et al. Distinct properties of the XY pseudoautosomal region crucial for male meiosis. Science 331, 916–920 (2011).

    Article  CAS  Google Scholar 

  28. Chen, Y. et al. A shared docking motif in TRF1 and TRF2 used for differential recruitment of telomeric proteins. Science 319, 1092–1096 (2008).

    Article  CAS  Google Scholar 

  29. Wan, B. et al. SLX4 assembles a telomere maintenance toolkit by bridging multiple endonucleases with telomeres. Cell Rep. 4, 861–869 (2013).

    Article  CAS  Google Scholar 

  30. Rai, R. et al. NBS1 phosphorylation status dictates repair choice of dysfunctional telomeres. Mol. Cell 65, 801–817 e4 (2017).

    Article  CAS  Google Scholar 

  31. Wang, H. et al. One-step generation of mice carrying mutations in multiple genes by CRISPR/Cas-mediated genome engineering. Cell 153, 910–918 (2013).

    Article  CAS  Google Scholar 

  32. de la Fuente, R. et al. Meiotic pairing and segregation of achiasmate sex chromosomes in eutherian mammals: the role of SYCP3 protein. PLoS Genet. 3, e198 (2007).

    Article  Google Scholar 

  33. de Vries, F.A.T. et al. Mouse Sycp1 functions in synaptonemal complex assembly, meiotic recombination, and XY body formation. Genes Dev. 19, 1376–1389 (2005).

    Article  CAS  Google Scholar 

  34. Klutstein, M. & Cooper, J.P. The chromosomal courtship dance-homolog pairing in early meiosis. Curr. Opin. Cell Biol. 26, 123–131 (2014).

    Article  CAS  Google Scholar 

  35. Scherthan, H. Analysis of telomere dynamics in mouse spermatogenesis. Methods Mol. Biol. 558, 383–399 (2009).

    Article  CAS  Google Scholar 

  36. Scherthan, H. et al. Centromere and telomere movements during early meiotic prophase of mouse and man are associated with the onset of chromosome pairing. J. Cell Biol. 134, 1109–1125 (1996).

    Article  CAS  Google Scholar 

  37. Hunter, N., Börner, G.V., Lichten, M. & Kleckner, N. Gamma-H2AX illuminates meiosis. Nat. Genet. 27, 236–238 (2001).

    Article  CAS  Google Scholar 

  38. Celli, G.B. & de Lange, T. DNA processing is not required for ATM-mediated telomere damage response after TRF2 deletion. Nat. Cell Biol. 7, 712–718 (2005).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  40. Kim, S.H., Kaminker, P. & Campisi, J. TIN2, a new regulator of telomere length in human cells. Nat. Genet. 23, 405–412 (1999).

    Article  CAS  Google Scholar 

  41. Ye, J.Z. et al. POT1-interacting protein PIP1: a telomere length regulator that recruits POT1 to the TIN2/TRF1 complex. Genes Dev. 18, 1649–1654 (2004).

    Article  CAS  Google Scholar 

  42. Frescas, D. & de Lange, T. TRF2-tethered TIN2 can mediate telomere protection by TPP1/POT1. Mol. Cell. Biol. 34, 1349–1362 (2014).

    Article  Google Scholar 

  43. Minor, W., Cymborowski, M., Otwinowski, Z. & Chruszcz, M. HKL-3000: the integration of data reduction and structure solution–from diffraction images to an initial model in minutes. Acta Crystallogr. D Biol. Crystallogr. 62, 859–866 (2006).

    Article  Google Scholar 

  44. Murshudov, G.N., Vagin, A.A. & Dodson, E.J. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr. D Biol. Crystallogr. 53, 240–255 (1997).

    Article  CAS  Google Scholar 

  45. Emsley, P., Lohkamp, B., Scott, W.G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).

    Article  CAS  Google Scholar 

  46. Schrodinger, L. The PyMOL Molecular Graphics System, Version 1.8. (2015).

  47. Moretti, P., Freeman, K., Coodly, L. & Shore, D. Evidence that a complex of SIR proteins interacts with the silencer and telomere-binding protein RAP1. Genes Dev. 8, 2257–2269 (1994).

    Article  CAS  Google Scholar 

  48. Bastos, H. et al. Flow cytometric characterization of viable meiotic and postmeiotic cells by Hoechst 33342 in mouse spermatogenesis. Cytometry A 65, 40–49 (2005).

    Article  Google Scholar 

  49. Getun, I.V., Torres, B. & Bois, P.R. Flow cytometry purification of mouse meiotic cells. J. Vis. Exp. 50, 2602 (2011).

    Google Scholar 

  50. Peters, A.H., Plug, A.W., van Vugt, M.J. & de Boer, P. A drying-down technique for the spreading of mammalian meiocytes from the male and female germline. Chromosome Res. 5, 66–68 (1997).

    Article  CAS  Google Scholar 

  51. Scherthan, H. et al. Mammalian meiotic telomeres: protein composition and redistribution in relation to nuclear pores. Mol. Biol. Cell 11, 4189–4203 (2000).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank H. Shi, X. Huang, Q. Shi, M. Tong, and J. Li for discussion. We thank M. Han (Fudan University) for the SUN1 antibody. We thank L. Wu, D. Yao, and R. Zhang from BL18U1 and BL19U1 beamlines at NCPSS and Shanghai Synchrotron Radiation Facility (SSRF) for help with crystal data collection and processing. We thank Y. Yu, S. He, Y. Wang and S. Li from NCPSS for help with confocal microscopy and FACS. This work was supported by grants from the Ministry of Science and Technology of China (2013CB910402 to M.L.), the National Natural Science Foundation of China (31330040 and 31525007 to M.L., 31500625 to J.W.), the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB08010201 to M.L.), the Outstanding Academic Leader Program of Science and Technology Commission of Shanghai Municipality (16XD1405000 to M.L. and C.H.), and the Youth Innovation Promotion Association of the Chinese Academy of Sciences (to J.W.). Y.L. acknowledges support from the Intramural Research Program of the National Institutes of Health, National Institute on Aging.

Author information

Authors and Affiliations

Authors

Contributions

J.L., C.H., J.W., and M.L. designed the study. J.L. crystallized the TERB1TBM–TRF1TRFH complex. J.W. collected X-ray data and determined the crystal structure. J.L., C.H., Y.C., and S.S. carried out biochemical assays. J.L., Y.L., C.L., Y.Z., and L.W. participated in the generation of the knock-in mice. J.L., C.H., and Y.C. performed the experiments to characterize the meiotic phenotype. C.H., J.W., and L.M. drafted the manuscript. All authors reviewed and edited the manuscript.

Corresponding authors

Correspondence to Jian Wu or Ming Lei.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Characterization of the interaction between TRF1TRFH and TERB1TBM.

Yeast two-hybrid analysis of the interaction between TRF1TRFH with different fragments of TERB1. Interaction between LexA-TRF1TRFH and different GAD-TERB1 fragments was analyzed by measuring the β-galactosidase activity produced by the reporter gene. Data are averages of three independent β-galactosidase measurements normalized to the interaction of TRF1TRFH with TERB1522-622, arbitrarily set to 100. Error bars: standard deviation (s.d.); n=3.

Supplementary Figure 2 Crystal structure of the TRF1TRFH and TERB1TBM complex.

(a) Stereo view of the TRF1TRFH-TERB1TBM complex. The electron density (2Fo-Fc) map of TERB1TBM, contoured at 1.0 σ, is shown in the complex. (b) The electrostatic surface potential of TRF1TRFH responsible for TERB1TBM binding (positive potential, blue; negative potential, red). TERB1TBM is in stick representation and colored in yellow.

Supplementary Figure 3 TERB1TBM and TIN2TBM occupy the same binding pocket on TRF1TRFH.

Electrostatic surface potential of superimposed complex structures, TRF1TRFH-TERB1TBM and TRF1TRFH-TIN2TBM are shown in (a), (b), and (c), respectively. TERB1TBM and TIN2TBM are in stick representation and colored in yellow and cyan.

Supplementary Figure 4 In vitro ITC measurement between TRF1TRFH and mutant TERB1TBM.

All the mutant TERB1TBM disrupt the interaction with TRF1TRFH, consistent with the Co-IP and co-localization data shown in Figure 2.

Supplementary Figure 5 The TERB1 647AEA649 mutant mice were generated using the CRISPR/Cas9 method.

(a) The Cas9/sgRNA-targeting sites in mouse Terb1 gene. The sgRNA-targeting sequence and the protospacer-adjacent motif (PAM) are underlined. The desired mutations on genomic DNA are indicated by lower-case letters. (b) Transgenic mouse offspring were genotyped by tail biopsy with PCR and DNA sequencing of the mutation loci. (c) Relative mRNA expression was determined in testes of 3-week old littermates by RT-qPCR using GAPDH as the loading control. The averaged value for WT was normalized to 1. Results were from three independent experiments. Error bars: standard deviation; n=3. (d) Western blot of TERB1 in testes of two-week-old littermates. β-actin was used as the loading control. The indicated numbers below the blot images represent the relative signals of anti-TERB1 normalized by the signals of anti-Actin. Band intensity of blots was quantified by ImageJ. The number for WT sample was set as 1. Uncropped blot images are shown in Supplementary Data Set 1.

Supplementary Figure 6 Hoechst 33342 fluorescence flow analysis of testicular cells from 6-week-old WT and Terb1AEA/AEA littermates.

(a) Left panel: Schematic diagram summarizes the Hoechst 33342 fluorescent profile of male germinal cells in adult spermatogenesis. Right panel: Hoechst 33342 fluorescence flow analysis of testicular cells from 6-week-old WT and Terb1AEA/AEA littermates. L-Z: subpopulation that mainly comprises leptotene and zygotene spermatocytes. P-D: subpopulation that mainly comprises pachytene and diplotene spermatocytes. Subpopulation was sorted and analyzed by immunofluorescence staining of SYCP3 and γ-H2AX or propidium iodide staining to determine the cell types (data not shown). (b) Quantification of spermatid population in all testicular cells. Statistical significances (*** p<0.01) were assessed by two-tailed t-tests. Error bars: standard deviation; n=4. (c) Quantification of L-Z and P-D subpopulation in all spermatocyte I (4N DNA content). Statistical significances (** p<0.05) were assessed by two-tailed t-tests. Error bars: standard deviation; n=4.

Supplementary Figure 7 Programmed DSB repair and chiasma formation autosomes are largely unaffected in Terb1AEA/AEA spermatocytes.

(a) WT and Terb1AEA/AEA pachytene spermatocytes stained for MLH1 (red), SYCP3 (green) and DAPI (blue). Scale bar: 10 μm. (b) Quantification of the number of MLH1 foci per cell from (a). The mean values were indicated. (c) WT and Terb1AEA/AEA spermatocytes stained for γ-H2AX (red), SYCP3 (green) and DAPI (blue). Scale bar: 10 μm.

Supplementary Figure 8 The 647AEA649 mutation of TERB1 results in telomere localization defect of SUN1.

(a) Cell lysates were prepared with testes from 3-week-old WT, heterozygous and Terb1AEA/AEA mice and immunoprecipitated with TERB1 antibody. Lysate samples and immunoprecipitates were blotted using antibodies as indicated. The indicated numbers below the blot images represent the relative signals of Co-IPed TRF1 normalized by the signals of IPed TERB1. Band intensity of blots was quantified by ImageJ. The number for WT sample was set as 1. Uncropped blot images are shown in Supplementary Data Set 1. (b) Localization of SUN1 (red) and SYCP3 (green) on pachytene spermatocyte spreads from WT and Terb1AEA/AEA mice. Scale bar: 10 μm. (c) Quantification of the relative intensity of SUN1 foci at the termini of synaptic chromosomes in WT and Terb1AEA/AEA pachytene spermatocytes. Three independent experiments with 6-week-old littermates showed similar results. More than 10 spermatocyte spreads were used for quantification for each group. (a.u., arbitrary units)

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8 (PDF 1810 kb)

Life Sciences Reporting Summary

Life Sciences Reporting Summary (PDF 176 kb)

Supplementary Data Set 1

Uncropped gels of supplementary figures. (PDF 1457 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Long, J., Huang, C., Chen, Y. et al. Telomeric TERB1–TRF1 interaction is crucial for male meiosis. Nat Struct Mol Biol 24, 1073–1080 (2017). https://doi.org/10.1038/nsmb.3496

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.3496

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