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:

DNA cross-link repair safeguards genomic stability during premeiotic germ cell development

Abstract

Germline de novo mutations are the basis of evolutionary diversity but also of genetic disease. However, the molecular origin, mechanisms and timing of germline mutagenesis are not fully understood. Here, we define a fundamental role for DNA interstrand cross-link repair in the germline. This repair process is essential for primordial germ cell (PGC) maturation during embryonic development. Inactivation of cross-link repair leads to genetic instability that is restricted to PGCs within the genital ridge during a narrow temporal window. Having successfully activated the PGC transcriptional program, a potent quality control mechanism detects and drives damaged PGCs into apoptosis. Therefore, these findings define a source of DNA damage and the nature of the subsequent DNA repair response in germ cells, which ensures faithful transmission of the genome between generations.

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: Ercc1−/− mice have reduced numbers of germ cells.
Fig. 2: ERCC1 is critical for PGC development.
Fig. 3: ERCC1 acts together with the Fanconi anemia pathway to preserve PGC development.
Fig. 4: Cross-link repair-deficient PGCs undergo normal PGC transcriptional activation and epigenetic reprogramming.
Fig. 5: PGCs accumulate unrepaired DNA breaks.
Fig. 6: Fetal aldehyde catabolism protects developing PGCs.
Fig. 7: Damaged PGCs are eliminated by apoptosis.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available from the corresponding author upon request.

References

  1. Veltman, J. A. & Brunner, H. G. De novo mutations in human genetic disease. Nat. Rev. Genet. 13, 565–575 (2012).

    Article  CAS  Google Scholar 

  2. Tang, W. W., Kobayashi, T., Irie, N., Dietmann, S. & Surani, M. A. Specification and epigenetic programming of the human germ line. Nat. Rev. Genet. 17, 585–600 (2016).

    Article  CAS  Google Scholar 

  3. Hajkova, P. et al. Genome-wide reprogramming in the mouse germ line entails the base excision repair pathway. Science 329, 78–82 (2010).

    Article  CAS  Google Scholar 

  4. Surani, M. A., Durcova-Hills, G., Hajkova, P., Hayashi, K. & Tee, W. W. Germ line, stem cells, and epigenetic reprogramming. Cold Spring Harb. Symp. Quant. Biol. 73, 9–15 (2008).

    Article  CAS  Google Scholar 

  5. Hajkova, P. et al. Chromatin dynamics during epigenetic reprogramming in the mouse germ line. Nature 452, 877–881 (2008).

    Article  CAS  Google Scholar 

  6. Nik-Zainal, S. et al. Mutational processes molding the genomes of 21 breast cancers. Cell 149, 979–993 (2012).

    Article  CAS  Google Scholar 

  7. Alexandrov, L. B., Nik-Zainal, S., Wedge, D. C., Campbell, P. J. & Stratton, M. R. Deciphering signatures of mutational processes operative in human cancer. Cell Rep. 3, 246–259 (2013).

    Article  CAS  Google Scholar 

  8. Alexandrov, L. B. et al. Signatures of mutational processes in human cancer. Nature 500, 415–421 (2013).

    Article  CAS  Google Scholar 

  9. Sekelsky, J. J., McKim, K. S., Chin, G. M. & Hawley, R. S. The Drosophila meiotic recombination gene mei-9 encodes a homologue of the yeast excision repair protein Rad1. Genetics 141, 619–627 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Yildiz, O., Majumder, S., Kramer, B. & Sekelsky, J. J. Drosophila MUS312 interacts with the nucleotide excision repair endonuclease MEI-9 to generate meiotic crossovers. Mol. Cell 10, 1503–1509 (2002).

    Article  CAS  Google Scholar 

  11. Saito, T. T., Lui, D. Y., Kim, H. M., Meyer, K. & Colaiacovo, M. P. Interplay between structure-specific endonucleases for crossover control during Caenorhabditis elegans meiosis. PLoS Genet. 9, e1003586 (2013).

    Article  CAS  Google Scholar 

  12. Baker, B. S. & Carpenter, A. T. Genetic analysis of sex chromosomal meiotic mutants in Drosophila melanogaster. Genetics 71, 255–286 (1972).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Hsia, K. T. et al. DNA repair gene Ercc1 is essential for normal spermatogenesis and oogenesis and for functional integrity of germ cell DNA in the mouse. Development 130, 369–378 (2003).

    Article  CAS  Google Scholar 

  14. McWhir, J., Selfridge, J., Harrison, D. J., Squires, S. & Melton, D. W. Mice with DNA repair gene (ERCC-1) deficiency have elevated levels of p53, liver nuclear abnormalities and die before weaning. Nat. Genet. 5, 217–224 (1993).

    Article  CAS  Google Scholar 

  15. Niedernhofer, L. J. et al. A new progeroid syndrome reveals that genotoxic stress suppresses the somatotroph axis. Nature 444, 1038–1043 (2006).

    Article  CAS  Google Scholar 

  16. Buaas, F. W. et al. Plzf is required in adult male germ cells for stem cell self-renewal. Nat. Genet. 36, 647–652 (2004).

    Article  CAS  Google Scholar 

  17. Szabó, P. E., Hübner, K., Schöler, H. & Mann, J. R. Allele-specific expression of imprinted genes in mouse migratory primordial germ cells. Mech. Dev. 115, 157–160 (2002).

    Article  Google Scholar 

  18. Yeom, Y. I. et al. Germline regulatory element of Oct-4 specific for the totipotent cycle of embryonal cells. Development 122, 881–894 (1996).

    CAS  PubMed  Google Scholar 

  19. Payer, B. et al. Generation of stella-GFP transgenic mice: a novel tool to study germ cell development. Genesis 44, 75–83 (2006).

    Article  CAS  Google Scholar 

  20. Ohinata, Y. et al. Blimp1 is a critical determinant of the germ cell lineage in mice. Nature 436, 207–213 (2005).

    Article  CAS  Google Scholar 

  21. Ahmad, A. et al. ERCC1-XPF endonuclease facilitates DNA double-strand break repair. Mol. Cell. Biol. 28, 5082–5092 (2008).

    Article  CAS  Google Scholar 

  22. Bogliolo, M. et al. Mutations in ERCC4, encoding the DNA-repair endonuclease XPF, cause Fanconi anemia. Am. J. Hum. Genet. 92, 800–806 (2013).

    Article  CAS  Google Scholar 

  23. Crossan, G. P. et al. Disruption of mouse Slx4, a regulator of structure-specific nucleases, phenocopies Fanconi anemia. Nat. Genet. 43, 147–152 (2011).

    Article  CAS  Google Scholar 

  24. Kottemann, M. C. & Smogorzewska, A. Fanconi anaemia and the repair of Watson and Crick DNA crosslinks. Nature 493, 356–363 (2013).

    Article  CAS  Google Scholar 

  25. Wong, J. C. et al. Targeted disruption of exons 1 to 6 of the Fanconi Anemia group A gene leads to growth retardation, strain-specific microphthalmia, meiotic defects and primordial germ cell hypoplasia. Hum. Mol. Genet. 12, 2063–2076 (2003).

    Article  CAS  Google Scholar 

  26. Sklavos, M. M., Giri, N., Stratton, P., Alter, B. P. & Pinto, L. A. Anti-Müllerian hormone deficiency in females with Fanconi anemia. J. Clin. Endocrinol. Metab. 99, 1608–1614 (2014).

    Article  CAS  Google Scholar 

  27. Giri, N., Stratton, P., Savage, S. A. & Alter, B. P. Pregnancies in patients with inherited bone marrow failure syndromes in the NCI cohort. Blood 130, 1674–1676 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Alavattam, K. G. et al. Elucidation of the Fanconi anemia protein network in meiosis and its function in the regulation of histone modifications. Cell Rep. 17, 1141–1157 (2016).

    Article  CAS  Google Scholar 

  29. Agoulnik, A. I. et al. A novel gene, Pog, is necessary for primordial germ cell proliferation in the mouse and underlies the germ cell deficient mutation, gcd. Hum. Mol. Genet. 11, 3047–3053 (2002).

    Article  CAS  Google Scholar 

  30. Luo, Y. et al. Hypersensitivity of primordial germ cells to compromised replication-associated DNA repair involves ATM-p53-p21 signaling. PLoS Genet. 10, e1004471 (2014).

    Article  Google Scholar 

  31. Houghtaling, S. et al. Epithelial cancer in Fanconi anemia complementation group D2 (Fancd2) knockout mice. Genes Dev. 17, 2021–2035 (2003).

    Article  CAS  Google Scholar 

  32. Yang, Y. et al. Targeted disruption of the murine Fanconi anemia gene, Fancg/Xrcc9. Blood 98, 3435–3440 (2001).

    Article  CAS  Google Scholar 

  33. Newkirk, S. J. et al. Intact piRNA pathway prevents L1 mobilization in male meiosis. Proc. Natl Acad. Sci. USA 114, E5635–E5644 (2017).

    Article  CAS  Google Scholar 

  34. Long, J. et al. Telomeric TERB1–TRF1 interaction is crucial for male meiosis. Nat. Struct. Mol. Biol. 24, 1073–1080 (2017).

    Article  CAS  Google Scholar 

  35. Yuen, B. T., Bush, K. M., Barrilleaux, B. L., Cotterman, R. & Knoepfler, P. S. Histone H3.3 regulates dynamic chromatin states during spermatogenesis. Development 141, 3483–3494 (2014).

    Article  CAS  Google Scholar 

  36. Gaysinskaya, V., Soh, I. Y., van der Heijden, G. W. & Bortvin, A. Optimized flow cytometry isolation of murine spermatocytes. Cytometry A 85, 556–565 (2014).

    Article  Google Scholar 

  37. Larson, E. L., Keeble, S., Vanderpool, D., Dean, M. D. & Good, J. M. The composite regulatory basis of the large X-effect in mouse speciation. Mol. Biol. Evol. 34, 282–295 (2017).

    CAS  PubMed  Google Scholar 

  38. Hayashi, K., Ohta, H., Kurimoto, K., Aramaki, S. & Saitou, M. Reconstitution of the mouse germ cell specification pathway in culture by pluripotent stem cells. Cell 146, 519–532 (2011).

    Article  CAS  Google Scholar 

  39. Hayashi, K. et al. Offspring from oocytes derived from in vitro primordial germ cell-like cells in mice. Science 338, 971–975 (2012).

    Article  CAS  Google Scholar 

  40. von Meyenn, F. et al. Comparative principles of DNA methylation reprogramming during human and mouse in vitro primordial germ cell specification. Dev. Cell 39, 104–115 (2016).

    Article  CAS  Google Scholar 

  41. Seisenberger, S. et al. The dynamics of genome-wide DNA methylation reprogramming in mouse primordial germ cells. Mol. Cell 48, 849–862 (2012).

    Article  CAS  Google Scholar 

  42. Hodskinson, M. R. et al. Mouse SLX4 is a tumor suppressor that stimulates the activity of the nuclease XPF-ERCC1 in DNA crosslink repair. Mol. Cell 54, 472–484 (2014).

    Article  CAS  Google Scholar 

  43. Trujillo, J. P. et al. On the role of FAN1 in Fanconi anemia. Blood 120, 86–89 (2012).

    Article  CAS  Google Scholar 

  44. Airik, R. et al. A FANCD2/FANCI-associated nuclease 1-knockout model develops karyomegalic interstitial nephritis. J. Am. Soc. Nephrol. 27, 3552–3559 (2016).

    Article  CAS  Google Scholar 

  45. Gillich, A. et al. Epiblast stem cell-based system reveals reprogramming synergy of germline factors. Cell Stem Cell 10, 425–439 (2012).

    Article  CAS  Google Scholar 

  46. Hackett, J. A. et al. Promoter DNA methylation couples genome-defence mechanisms to epigenetic reprogramming in the mouse germline. Development 139, 3623–3632 (2012).

    Article  CAS  Google Scholar 

  47. Francis, R. J. & Lo, C. W. Primordial germ cell deficiency in the connexin 43 knockout mouse arises from apoptosis associated with abnormal p53 activation. Development 133, 3451–3460 (2006).

    Article  CAS  Google Scholar 

  48. Takeuchi, Y., Molyneaux, K., Runyan, C., Schaible, K. & Wylie, C. The roles of FGF signaling in germ cell migration in the mouse. Development 132, 5399–5409 (2005).

    Article  CAS  Google Scholar 

  49. Molyneaux, K. A. et al. The chemokine SDF1/CXCL12 and its receptor CXCR4 regulate mouse germ cell migration and survival. Development 130, 4279–4286 (2003).

    Article  CAS  Google Scholar 

  50. Seki, Y. et al. Extensive and orderly reprogramming of genome-wide chromatin modifications associated with specification and early development of germ cells in mice. Dev. Biol. 278, 440–458 (2005).

    Article  CAS  Google Scholar 

  51. Seki, Y. et al. Cellular dynamics associated with the genome-wide epigenetic reprogramming in migrating primordial germ cells in mice. Development 134, 2627–2638 (2007).

    Article  CAS  Google Scholar 

  52. Hajkova, P. et al. Epigenetic reprogramming in mouse primordial germ cells. Mech. Dev. 117, 15–23 (2002).

    Article  CAS  Google Scholar 

  53. Olek, A. & Walter, J. The pre-implantation ontogeny of the H19 methylation imprint. Nat. Genet. 17, 275–276 (1997).

    Article  CAS  Google Scholar 

  54. Garaycoechea, J. I. et al. Genotoxic consequences of endogenous aldehydes on mouse haematopoietic stem cell function. Nature 489, 571–575 (2012).

    Article  CAS  Google Scholar 

  55. Pontel, L. B. et al. Endogenous formaldehyde is a hematopoietic stem cell genotoxin and metabolic carcinogen. Mol. Cell 60, 177–188 (2015).

    Article  CAS  Google Scholar 

  56. Langevin, F., Crossan, G. P., Rosado, I. V., Arends, M. J. & Patel, K. J. Fancd2 counteracts the toxic effects of naturally produced aldehydes in mice. Nature 475, 53–58 (2011).

    Article  CAS  Google Scholar 

  57. Rosado, I. V., Langevin, F., Crossan, G. P., Takata, M. & Patel, K. J. Formaldehyde catabolism is essential in cells deficient for the Fanconi anemia DNA-repair pathway. Nat. Struct. Mol. Biol. 18, 1432–1434 (2011).

    Article  CAS  Google Scholar 

  58. Garaycoechea, J. I. et al. Alcohol and endogenous aldehydes damage chromosomes and mutate stem cells. Nature 553, 171–177 (2018).

    Article  CAS  Google Scholar 

  59. Lai, C. L. et al. Dominance of the inactive Asian variant over activity and protein contents of mitochondrial aldehyde dehydrogenase 2 in human liver. Alcohol. Clin. Exp. Res. 38, 44–50 (2014).

    Article  CAS  Google Scholar 

  60. Heyer, B. S., MacAuley, A., Behrendtsen, O. & Werb, Z. Hypersensitivity to DNA damage leads to increased apoptosis during early mouse development. Genes Dev. 14, 2072–2084 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Raya, A. et al. Disease-corrected haematopoietic progenitors from Fanconi anaemia induced pluripotent stem cells. Nature 460, 53–59 (2009).

    Article  CAS  Google Scholar 

  62. Müller, L. U. et al. Overcoming reprogramming resistance of Fanconi anemia cells. Blood 119, 5449–5457 (2012).

    Article  Google Scholar 

  63. Takahashi, K. & Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 126, 663–676 (2006).

    Article  CAS  Google Scholar 

  64. Kim, S. et al. PRMT5 protects genomic integrity during global DNA demethylation in primordial germ cells and preimplantation embryos. Mol. Cell 56, 564–579 (2014).

    Article  CAS  Google Scholar 

  65. Popp, I. et al. Fanconi anemia with sun-sensitivity caused by a Xeroderma pigmentosum-associated missense mutation in XPF. BMC Med. Genet. 19, 7 (2018).

    Article  Google Scholar 

  66. Auerbach, A. D. Fanconi anemia and its diagnosis. Mutat. Res. 668, 4–10 (2009).

    Article  CAS  Google Scholar 

  67. Norris, P. G., Hawk, J. L., Avery, J. A. & Giannelli, F. Xeroderma pigmentosum complementation group F in a non-Japanese patient. J. Am. Acad. Dermatol. 18, 1185–1188 (1988).

    Article  CAS  Google Scholar 

  68. Fujiwara, Y. et al. Xeroderma pigmentosum groups C and F: additional assignments and a review of the subjects in Japan. J. Radiat. Res. 26, 443–449 (1985).

    Article  CAS  Google Scholar 

  69. van der Horst, G. T. et al. Defective transcription-coupled repair in Cockayne syndrome B mice is associated with skin cancer predisposition. Cell 89, 425–435 (1997).

    Article  CAS  Google Scholar 

  70. Jaarsma, D. et al. Age-related neuronal degeneration: complementary roles of nucleotide excision repair and transcription-coupled repair in preventing neuropathology. PLoS Genet. 7, e1002405 (2011).

    Article  CAS  Google Scholar 

  71. Farley, F. W., Soriano, P., Steffen, L. S. & Dymecki, S. M. Widespread recombinase expression using FLPeR (Flipper) mice. Genesis 28, 106–110 (2000).

    Article  CAS  Google Scholar 

  72. 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 

  73. Kawasaki, Y. et al. Active DNA demethylation is required for complete imprint erasure in primordial germ cells. Sci. Rep. 4, 3658 (2014).

    Article  Google Scholar 

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

    Google Scholar 

Download references

Acknowledgements

We would like to thank the Wellcome Trust Sanger Institute, the Knockout Mouse Project Consortium and the Mouse Biology Program, University of California, Davis for the Ercc1-targeted mESCs. We thank A. Surani for the gift of Stella-GFP mice, F. Hildebrandt for Fan1-deficient mice, G.T. van der Horst for the gift of Xpa- and Csb-deficient mice. We would like to thank K.J. Patel for the ALDH2- and ADH5-deficient mice. We thank the Human Research Tissue Bank (National Institute for Health Research Cambridge Biomedical Research Centre) for processing the histology. We would like to thank R. Pannell, C. Knox, C. Watson, K. Kemp, R. Higginson, J. Clark, the Ares and Biomed staff for assistance with animal procedures and experiments. We thank M. Daly, F. Zhang and the Flow Cytometry Core staff for their technical assistance. We would like to thank G. Oliveira for assistance with the PGCLC assays. We would like to thank V. Sacalean for assistance with apoptosis assays. We would like to thank J. Garaycoechea for his criticism, for screening the Ercc1 founder mice, for collecting samples from NER-deficient mice and for critically reading the manuscript. We would like to thank J. Sale for critically reading the manuscript and for useful discussions. R.J.H. and G.P.C. are supported by the Medical Research Council. This work was supported by the Medical Research Council as part of UK Research and Innovation (file reference no. MC_UP_1201/18). G.P.C. would like to thank K.J. Patel for scientific discussion and support.

Author information

Authors and Affiliations

Authors

Contributions

G.P.C. and R.J.H. conceived the study, designed experiments and wrote the paper. R.J.H. performed all the experiments.

Corresponding author

Correspondence to Gerry P. Crossan.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Ercc1 knockout mice recapitulate key features of ERCC1 deficiency.

a, Schematic representation of the wildtype (Ercc1+) and disrupted Ercc1tm1a(KOMP)Wtsi allele (Ercc1), when crossed with Flp-e recombinase this generates the conditional allele (Ercc1f). This can produce a null allele by crossing with Cre-recombinase, deleting the floxed exon 5 (Ercc1Δ). b, Verification of gene targeting in mESCs by long-range PCR. Oligonucleotide pairs were designed so that one oligonucleotide binds within the targeting construct and the other binds within the genomic sequence beyond the homology arm. c, Mice were genotyped using two PCR strategies to allow the discrimination between the four different alleles. Firstly, oligonucleotides ER1F2, ER1R1 and En2A would amplify a 594 bp product from the Ercc1+ allele, a 477 bp product from the Ercc1 allele and a 965 bp product from the Ercc1f allele (top panel). Secondly, oligonucleotides ER1NF and ER1NR would amplify a 776 bp product from the Ercc1+ allele, a 947 bp product from the Ercc1f allele and a 339 bp product from the Ercc1Δ allele (bottom panel). d, Immunoblot showing the absence of the ERCC1 protein in whole cell lysates of MEFs derived from Ercc1-/- embryos (representative data from 2 independent experiments). Uncropped images in Supplementary Fig. 17 e, Ercc1-/- MEFs are hypersensitive to mitomycin c (MMC) and ultraviolet (UV) irradiation (data represent mean and s.e.m.; data represent 3 independent experiments each carried out in triplicate). f, Kaplan-Meier curve showing the survival of cohorts of Ercc1-/- and congenic control mice. g, Ercc1-/- embryos are observed at expected ratios throughout development (E11.5, E12.5 and E18.5). However, Ercc1-/- mice are observed at a significantly reduced frequency in 14-day old (P14) mice, obtained from Ercc1+/- males intercrossed with Ercc1+/- females (p-value calculated by Chi-square test). h, Ercc1-/- mice have a moderate reduction in ovary mass (p-value calculated by 2-tailed Mann-Whitney test; data shown represent mean and s.e.m.; each point represents one ovary, n= 21 and 6, left to right).

Supplementary Figure 2 Ercc1-/- embryos have reduced numbers of germ cells.

a, Quantification of MVH+ cells per seminiferous tubule at E18.5 (p-value calculated by 2-tailed Mann-Whitney test; data shown represent mean and s.e.m.; a total of 150 tubules per genotype were scored, 50 per mouse). b, Distribution of the number of MVH+ cells per tubule in E18.5 wildtype and Ercc1-/- embryos (250 tubules per genotype were scored, 50 per mouse). c, Representative images of E18.5 ovarian sections stained with the marker of apoptosis cleaved-Caspase 3 (CC3) and quantification of the frequency of CC3+ cells (p-value calculated by 2-tailed Mann-Whitney test; data represent mean and s.e.m.; each point represents data from one embryo, n= 4 per genotype). d, Representative images of E18.5 testis sections stained with the marker of apoptosis cleaved-Caspase 3 (CC3) and quantification of the frequency of CC3+ cells (p-value calculated by 2-tailed Mann-Whitney test; data represent mean and s.e.m.; each point represents data from one embryo, n= 5 per genotype).

Supplementary Figure 3 ERCC1 is required for PGC development.

a, Quantification of PGCs (SSEA1+GFP+) per gonad at E12.5 from wildtype and Ercc1-/- embryos carrying either the GOF18-GFP or Stella-GFP reporters show equivalent defects with either reporter (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=11, 7, 10 and 6, left to right). b, Quantification of PGCs (SSEA1+GOF18-GFP+) per gonad by flow cytometry from embryos carrying the GOF18-GFP reporter at E11.5 separated by sex (p-value calculated by Mann-Whitney test; data shown represents mean and s.e.m.; each point represents data from one embryo, n=2, 2, 2 and 2, left to right). c, Quantification of PLZF+ cells per seminiferous tubule in adult Ercc1-/fBlimp1-Cre+ and control mice (p-value calculated by 2-tailed Mann-Whitney test; data represent median and interquartile range ; n=150 tubules per genotype from n=3 mice). d, Frequency of PLZF+ cells per seminiferous tubule excluding tubules with no PLZF+ cells in adult Ercc1-/fBlimp1-Cre+ and control mice (p-value calculated by 2-tailed Mann-Whitney test; data shown represent mean and s.e.m.; n=150 tubules per genotype, 50 per mouse). e-f, Cumulative number of offspring over successive litters from male (e) and female (f) mice carrying the conditional Ercc1 allele and the PGC Cre-recombinase (Ercc1-/fBlimp1-Cre+) and congenic controls, bred with wildtype mates shows that expression of Ercc1 in PGCs is essential for fertility (data shown represent mean and s.e.m.; n=3 mice per genotype.

Supplementary Figure 4 Fanconi-mediated DNA crosslink repair is essential for fertility in mice.

a, Microscopic analysis of H&E-stained ovaries, testes and epididymis of 8-12-week old mice with constitutive deletions of Xpa, Csb, Fanca and mice that lack Ercc1 in PGCs (Ercc1-/fBlimp1-Cre+) (representative data, 3 independent animals per assessed per genotype and gender). b, Cumulative number of offspring over successive litters from male and female mice with constitutive deletions of Xpa, Csb, Fanca and mice that lack Ercc1 in PGCs (Ercc1-/fBlimp1-Cre+) when bred with wildtype mates (data shown represent mean and s.e.m.; n=3 mice per genotype and gender, mice were checked for evidence of copulation). c, Quantification of PLZF+ cells per tubule in 8-12-week old wildtype and Fanca-/- mice (p-value calculated by 2-tailed Mann-Whitney test; data represent median and the interquartile range; n=150 tubules per genotype, 50 per mouse). d, Distribution of the number of PLZF+ cells per tubule in 8-12-week old wildtype and Fanca-/- mice (n=150 tubules per genotype, 50 per mouse). e, Frequency of PLZF+ cells per seminiferous tubule excluding tubules with no PLZF+ cells in 8- to 12-week old mice (p-value calculated by 2-tailed Mann-Whitney test; data shown represent mean and s.e.m.; n=150 tubules per genotype, 50 per mouse).

Supplementary Figure 5 ERCC1 and FANCA are dispensable for entry and progression through meiosis.

a, Frequency of meiotic intermediates observed in wildtype and constitutive Ercc1-/- and Fanca-/- females at E16.5 and E18.5, and male mice at P21 and P50 (as Ercc1-/- mice die before P50 Ercc1-/fStra8-iCre+ males were employed) (data represent mean and s.e.m.; n=3 embryos or mice per genotype and gender). b, Representative immunofluorescence images of meiotic cells from P21 mice (representative data). c, Generation of offspring using a conditional Ercc1 or Fanca allele with Stra8-iCre. d, Flow cytometry plot showing the spermatogonia (Sp) used for isolation of cells from the testes of adult mice (all samples that were sorted were similar to this plot). Gating strategy shown in Supplementary Fig. 18. e, qRT-PCR expression analysis of the spermatogonial stem cell marker Plzf (data represent mean and s.e.m.; each point represents one mouse, n=4 biologically independent samples per genotype, expression was normalized to Gapdh and made relative to the sorted spermatogonia population). f, Immunoblot for PLZF on equivalent cell numbers of whole testis lysate and sorted populations. g,h, Immunoblot for ERCC1 and FANCA in whole cell spermatogonia lysates shows successful Stra8-iCre-mediated loss of ERCC1 or FANCA, respectively. Uncropped images in Supplementary Fig. 17 These blots are representative of 3 independent experiments i, Propidium iodide DNA content analysis of whole testis and sorted spermatogonia. This experiment was repeated on three separate occasions. Gating strategy shown in Supplementary Fig. 19.

Supplementary Figure 6 Generation of functional gametes in the absence of ERCC1 or FANCA.

a, Cumulative number of offspring over successive months from male mice carrying the conditional Ercc1 allele and the Cre-recombinase Stra8-iCre and congenic controls bred with wildtype mates (data shown represent mean and s.e.m.; n=3 mice per genotype). b, Representative genotyping gels from the offspring of Ercc1-/fStra8-iCre+ and congenic control showing the inheritance of recombined alleles in offspring when mated with wildtype mates. c, The frequency of Stra8-iCre-mediated recombination as determined by the inheritance of recombined alleles by offspring of Ercc1-/fStra8-iCre+ and control mice when mated with wildtype mates (p-value calculated by 2-tailed Mann-Whitney test; data shown represent mean and s.e.m.; each point represents the total offspring of one animal over 3 litters, n=4 and 6, left to right). d, Cumulative number of offspring from male mice carrying the conditional Fanca allele and Stra8-iCre and congenic controls bred with wildtype mates (data shown represent mean and s.e.m.; n=3 mice per genotype). e, Representative genotyping gels from the offspring of Fanca-/fStra8-iCre+ and congenic control showing the inheritance of recombined alleles in offspring when mated with wildtype mates. f, The frequency of Stra8-iCre-mediated recombination as determined by the inheritance of recombined alleles by offspring of Fanca-/fStra8-iCre+ and control mice when mated with wildtype mates (p-value calculated by 2-tailed Mann-Whitney test; data shown represent mean and s.e.m.; each point represents the total offspring of one animal over 3 litters, n=3 and 4, left to).

Supplementary Figure 7 FANCA and ERCC1 are required for PGC-like cell (PGCLC) development in vitro.

a, Immunoblot for ERCC1 and FANCA in whole cell lysates shows successful loss of ERCC1 and FANCA, respectively, in targeted mouse embryonic stem cells (mESCs) (ES clones were screened by immunoblot and the blot repeated for the clones used in the assays below, the blot was repeated on 3 occasions). Uncropped images in Supplementary Fig. 17 b, Ercc1-/- mESCs are hypersensitive to ultraviolet (UV) irradiation and mitomycin c (MMC) and Fanca-/- mESCs are hypersensitive to just mitomycin c (MMC) (data represent mean and s.e.m.; data represent 3 independent experiments, each carried out in triplicate). c,d, In vitro differentiation of mESCs carrying the PGC-specific GOF18-GFP reporter into primordial germ cell-like cells (PGCLCs). (c) Representative images of GOF18-GFP fluorescence and flow cytometry quantification of PGCLCs (SSEA1+GOF18-GFP+) formed from parental and ERCC1-deficient mESCs (p-value calculated by 2-tailed Mann-Whitney test; data represent mean and s.e.m.; each point represents data from one independent experiment, n= 9 and 5, left to right). (d) Representative images of GOF18-GFP fluorescence and flow cytometry quantification of PGCLCs (SSEA1+GOF18-GFP+) formed from FANCA-deficient and congenic control mESCs (p-value calculated by 2-tailed Mann-Whitney test; data represent mean and s.e.m.; each point represents data from one independent experiment, n= 9 and 8, left to right).

Supplementary Figure 8 FANCA is required for both male and female PGC development.

a, Quantification of PGCs (SSEA1+GOF18-GFP+) per gonad at E11.5, E12.5 and E13.5 from wildtype and Fanca-/- embryos (p-value calculated by 2-tailed Mann-Whitney test; data represent mean and s.e.m.; each point represents data from one embryo, E11.5 n=9 and 13, E12.5 n=13 and 22, E13.5 n=5 and 9, left to right). b, Quantification of PGCs (SSEA1+GOF18-GFP+) per gonad at E11.5, E12.5 and E13.5 from wildtype and Fanca-/- embryos separated by sex (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, E11.5 n=6, 2, 6 and 5, E12.5 n=9, 4, 7 and 4, E13.5 n=2, 2, 14 and 7, left to right).

Supplementary Figure 9 Fanconi-mediate crosslink repair is essential for normal fertility in mice and crosslink repair-deficient PGCs express germ cell markers.

a, H&E staining of ovaries, testes and epididymis of 8-12-week old mice with constitutive deletions of Fan1 and Slx4 shows that SLX4 but not FAN1 is required for gamete production in mice (representative data from 3 independent animals per genotype and gender). b, Cumulative number of offspring over successive litters from male and female mice with constitutive deletions of Fan1 or Slx4 when mated with wildtype mates (data represent mean and s.e.m.; n=3 mice per genotype and gender, mice were checked for evidence of copulation). c, Quantification of the germ cell compartment by flow cytometry of gonads from wildtype, Fanca-/-, Ercc1-/- and Fan1-/- E12.5 embryos shows that Fanconi-mediated crosslink repair is required for germ cell development (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=14, 10, 7 and 4, left to right). d, qRT-PCR expression analysis of early genital ridge germ cell markers (Ap2γ, Prdm14, Blimp-1 and Stella) in PGCs (SSEA1+GOF18-GFP+) and somatic cells (SSEA1-GOF18-GFP-) purified from E10.5 wildtype embryos (data represent mean and s.e.m.; each point represents data from one embryo, Ap2γ n=6, 4, 2, and 2, Prdm14 n=6, 2, 2 and 2, Blimp-1 n=4, 3, 3, and 3, Stella n=6, 3, 3, and 3, left to right). For each marker, the expression of each sample was normalized to Gapdh and made relative to the somatic cell expression.

Supplementary Figure 10 Crosslink repair-deficient PGCs migrate to the genital ridge.

a, Representative GOF18-GFP fluorescence images of whole-mount E9.5 wildtype and Ercc1-/- embryos. b, Microscopic quantification of the number of PGCs (GFP+ for GOF18-GFP transgene) in E9.5 wildtype and Ercc1-/- embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=9 and 3, left to right). c, Quantification of the distribution of PGCs (GFP+ for GOF18-GFP transgene) within the developing E9.5 embryo using somites as landmarks (data represent mean and s.e.m.; n=3 embryos per genotype). d, Representative GOF18-GFP fluorescence images of whole-mount E10.5 wildtype and Ercc1-/- embryos, the dashed line marks the outline of the embryo. e, Microscopic quantification of the number of PGCs (GFP+ for GOF18-GFP transgene) in E10.5 wildtype and Ercc1-/- embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=6 per genotype). f, Quantification of the number of ectopically located PGCs (GFP+ for GOF18-GFP transgene) in E10.5 wildtype and Ercc1-/- embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=4 and 5, left to right). g, Quantification of the distribution of PGCs (GFP+ for GOF18-GFP transgene) within the developing E10.5 embryo using somites as physiological landmarks (data represent mean and s.e.m.; n=4 embryos per genotype).

Supplementary Figure 11 Crosslink repair-deficient PGCs are reduced in number but localize to the genital ridges.

a, Representative GOF18-GFP fluorescence images of whole-mount E11.5 wildtype and Ercc1-/- embryos, the dashed line marks the outline of the two gonads. b, Quantification of the number of PGCs (GFP+ for GOF18-GFP transgene) in E11.5 wildtype and Ercc1-/- embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=9 and 5, left to right). c, Microscopic quantification of the number of PGCs (GFP+ for GOF18-GFP transgene) confined within the genital ridges of E11.5 wildtype and Ercc1-/- embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=4 and 3, left to right).

Supplementary Figure 12 Ercc1-/- and Fanca-/- PGCs successfully undergo DNA demethylation at Tex19.1 and Mili.

a, Genomic bisulfite sequencing analysis demonstrates promoter demethylation at Tex19.1 in FACS-purified PGCs (SSEA1+GOF18-GFP+) compared to surrounding somatic cells (SSEA1-GOF18-GFP-) in E12.5 Wildtype, Ercc1-/- and Fanca-/- embryos. b, Genomic bisulfite sequencing analysis demonstrates promoter demethylation at Mili in FACS-purified PGCs (SSEA1+GOF18-GFP+) compared to surrounding somatic cells (SSEA1-GOF18-GFP-) in E12.5 Wildtype, Ercc1-/- and Fanca-/- embryos. All data shown represent individual sequencing reads (filled = methylated CpG, open = unmethylated CpG, n=3 independent embryos per genotype).

Supplementary Figure 13 Ercc1-/- and Fanca-/- PGCs successfully undergo DNA demethylation at H19 DMR and Line1.

a, Bisulfite analysis of the H19 DMR in FACS-purified PGCs (SSEA1+GOF18-GFP+) compared to surrounding somatic cells (SSEA1-GOF18-GFP-) in E12.5 Wildtype, Ercc1-/- and Fanca-/- embryos. b, Bisulfite analysis of the CpG-rich region of the transposable element Line1 in FACS-purified PGCs (SSEA1+GOF18-GFP+) compared to surrounding somatic cells (SSEA1-GOF18-GFP-) in E12.5 Wildtype, Ercc1-/- and Fanca-/- embryos. All data shown represent individual sequencing reads (filled = methylated CpG, open = unmethylated CpG, n=3 independent embryos per genotype).

Supplementary Figure 14 Mutant PGCs replicate and accumulate markers of DNA damage, but undergo normal changes in histone modifications.

a, Representative flow cytometry plot and quantification of EdU+ PGCs (SSEA1+GOF18-GFP+) from E11.5 embryos following a single intraperitoneal (IP) injection of EdU. This shows that equivalent numbers of Ercc1-/- and Fanca-/- PGCs incorporate EdU compared to controls (p-value calculated by unpaired t-test; data shown as mean and SD, each point represents data from one embryo, n=2 per genotype). Gating strategy shown in Supplementary Figure 19. b-c, Quantification of the number of PGCs that stain positively for H3K27me3 from wildtype and Fanca-/- E11.5 embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=5, 3 and 3, left to right) and (c) quantification of the number of γ−H2A.X foci per nucleus. Left bar illustrates the percentage of PGCs with <10 γ−H2A.X foci per nucleus, which stain positively (red) or negatively (grey) for H3K27me3. The central pie chart illustrates the percentage of PGCs with <10 γ−H2A.X foci (blue) or >10 γ−H2A.X foci per nucleus (orange). The right bar illustrates the percentage of PGCs with >10 γ−H2A.X foci per nucleus, which either stain positively (red) or negatively (gray) for H3K27me3 (representative data from 3 independent animals per genotype). d-e, Quantification of the number of PGCs that stain positively for H3K9me2 from wildtype and Fanca-/- E11.5 embryos (p-value calculated by 2-tailed Mann-Whitney test; data shown as mean and s.e.m.; each point represents data from one embryo, n=5, 3 and 2, left to right) and (e) quantification of the number of γ−H2A.X foci per nucleus. Left bar illustrates the percentage of PGCs with <10 γ−H2A.X foci per nucleus, which stain positively (red) or negatively (gray) for H3K9me2. The central pie chart illustrates the percentage of PGCs with <10 γ−H2A.X foci (blue) or >10 γ−H2A.X foci per nucleus (orange). The right bar illustrates the percentage of PGCs with >10 γ−H2A.X foci per nucleus, which either stain positively (red) or negatively (gray) for H3K9me2 (representative data from 3 independent animals per genotype).

Supplementary Figure 15 PGCs with markers of DNA damage have phosphorylated p53.

a, Quantification of the number of PGCs that stain positively for phosphorylated p53 (pSer15-P53) and the number of γ−H2A.X foci per nucleus from E11.5 wildtype, Ercc1-/- and Fanca-/- embryos. Left bar illustrates the percentage of PGCs with <10 γ−H2A.X foci per nucleus, which either stain positively (red) or negatively (gray) for pSer15-P53. The central pie chart illustrates the percentage of PGCs with <10 γ−H2A.X foci (blue) or >10 γ−H2A.X foci per nucleus (orange). The right bar illustrates the percentage of PGCs with >10 γ−H2A.X foci per nucleus, which either stain positively (red) or negatively (gray) for pSer15-P53 (representative data from 3 independent animals per genotype).

Supplementary Figure 16 FA Crosslink repair is crucial during a narrow temporal window of PGC development.

DNA crosslink repair plays a critical role during a narrow window of embryonic PGC development. Crosslink repair is critical to prevent the accumulation of DNA double strand breaks. Surprisingly, this accumulation of damaged DNA is specific to the PGCs and does not occur in surrounding somatic cells. Damaged PGCs accumulated phosphorylated p53 and enter into apoptosis, providing a mechanism through which corrupted PGCs are eliminated, preventing their mutated genome being passed on to the next generation. (Upper panel) schema showing formation of an early embryo following fertilisation, which gives rise to somatic tissue from which the initial primordial germ cell (PGC) population is specified. The PGCs migrate towards and seed the genital ridge where they undergo epigenetic reprogramming prior to entering meiosis (at E13.5 in females and postnatally in males) to generate mature gametes. (Middle panel) FA crosslink repair-deficient PGCs migrate towards the genital ridge but are subsequently lost. (Lower panel) Crosslink repair-deficient PGCs accumulate unrepaired DNA DSBs within the developing gonad, stabilize p53 and are lost by apoptosis whilst the surrounding tissue is preserved.

Supplementary Figure 17

Uncropped western blots images.

Supplementary Figure 18

Flow cytometry gating strategy of PGCs and Spermatogonia

Supplementary Figure 19

Flow cytometry gating strategy of DNA content analysis and EdU incorporation.

Supplementary information

Supplementary Information

Supplementary Figs. 1–19, Supplementary Table 1 and Supplementary Note

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hill, R.J., Crossan, G.P. DNA cross-link repair safeguards genomic stability during premeiotic germ cell development. Nat Genet 51, 1283–1294 (2019). https://doi.org/10.1038/s41588-019-0471-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-019-0471-2

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