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:

Thymidine nucleotide metabolism controls human telomere length

A Publisher Correction to this article was published on 10 May 2023

This article has been updated

Abstract

Telomere length in humans is associated with lifespan and severe diseases, yet the genetic determinants of telomere length remain incompletely defined. Here we performed genome-wide CRISPR–Cas9 functional telomere length screening and identified thymidine (dT) nucleotide metabolism as a limiting factor in human telomere maintenance. Targeted genetic disruption using CRISPR–Cas9 revealed multiple telomere length control points across the thymidine nucleotide metabolism pathway: decreasing dT nucleotide salvage via deletion of the gene encoding nuclear thymidine kinase (TK1) or de novo production by knockout of the thymidylate synthase gene (TYMS) decreased telomere length, whereas inactivation of the deoxynucleoside triphosphohydrolase-encoding gene SAMHD1 lengthened telomeres. Remarkably, supplementation with dT alone drove robust telomere elongation by telomerase in cells, and thymidine triphosphate stimulated telomerase activity in a substrate-independent manner in vitro. In induced pluripotent stem cells derived from patients with genetic telomere biology disorders, dT supplementation or inhibition of SAMHD1 promoted telomere restoration. Our results demonstrate a critical role of thymidine metabolism in controlling human telomerase and telomere length, which may be therapeutically actionable in patients with fatal degenerative diseases.

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: Telomere length CRISPR–Cas9 screening reveals that dT nucleotide metabolism genes are required for telomere length control.
Fig. 2: dT treatment drives telomerase-dependent telomere lengthening in human cells.
Fig. 3: dT nucleotide metabolism genes control dT-mediated telomere lengthening.
Fig. 4: SAMHD1 restricts human telomere length and limits dT-mediated telomere elongation.
Fig. 5: Telomere lengthening from dT treatment occurs without inhibiting cell growth.
Fig. 6: Induction of replication stress is insufficient to explain telomere lengthening from dT treatment.
Fig. 7: dT nucleotides enhance human telomerase activity independent of dTTP’s role as a telomerase substrate.
Fig. 8: dT supplementation or SAMHD1 knockdown drives telomere lengthening in iPSCs from patients with TBDs.

Similar content being viewed by others

Data availability

The sgRNA library sequencing data and T-free TRAP sequencing data used for analysis have been deposited in the Sequence Read Archive and are available via BioProject accession code PRJNA851386. Source data are provided with this paper.

Code availability

The MATLAB script used to analyze the T-free TRAP sequencing data has been posted to a public repository76. Version 1.0 was used in this manuscript (https://doi.org/10.5281/zenodo.7607615).

Change history

References

  1. Fagagna, F. et al. A DNA damage checkpoint response in telomere-initiated senescence. Nature 426, 194–198 (2003).

    Article  Google Scholar 

  2. Takai, H., Smogorzewska, A. & de Lange, T. DNA damage foci at dysfunctional telomeres. Curr. Biol. 13, 1549–1556 (2003).

    Article  CAS  PubMed  Google Scholar 

  3. Olovnikov, A. M. A theory of marginotomy: the incomplete copying of template margin in enzymic synthesis of polynucleotides and biological significance of the phenomenon. J. Theor. Biol. 41, 181–190 (1973).

    Article  CAS  PubMed  Google Scholar 

  4. Watson, J. D. Origin of concatemeric T7 DNA. Nat. New Biol. 239, 197–201 (1972).

    Article  CAS  PubMed  Google Scholar 

  5. Harley, C. B., Futcher, A. B. & Greider, C. W. Telomeres shorten during ageing of human fibroblasts. Nature 345, 458–460 (1990).

    Article  CAS  PubMed  Google Scholar 

  6. Greider, C. W. & Blackburn, E. H. Identification of a specific telomere terminal transferase activity in tetrahymena extracts. Cell 43, 405–413 (1985).

    Article  CAS  PubMed  Google Scholar 

  7. Greider, C. W. & Blackburn, E. H. A telomeric sequence in the RNA of Tetrahymena telomerase required for telomere repeat synthesis. Nature 337, 331–337 (1989).

    Article  CAS  PubMed  Google Scholar 

  8. Feng, J. et al. The RNA component of human telomerase. Science 269, 1236–1241 (1995).

    Article  CAS  PubMed  Google Scholar 

  9. Codd, V. et al. Polygenic basis and biomedical consequences of telomere length variation. Nat. Genet. 53, 1425–1433 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Armanios, M. & Blackburn, E. H. The telomere syndromes. Nat. Rev. Genet. 13, 693–704 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Niewisch, M. R. & Savage, S. A. An update on the biology and management of dyskeratosis congenita and related telomere biology disorders. Expert Rev. Hematol. 12, 1037–1052 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Revy, P., Kannengiesser, C. & Bertuch, A. A. Genetics of human telomere biology disorders. Nat. Rev. Genet. 24, 86–108 (2023).

    Article  CAS  PubMed  Google Scholar 

  13. Mathews, C. K. Deoxyribonucleotide metabolism, mutagenesis and cancer. Nat. Rev. Cancer 15, 528–539 (2015).

    Article  CAS  PubMed  Google Scholar 

  14. Buj, R. & Aird, K. M. Deoxyribonucleotide triphosphate metabolism in cancer and metabolic disease. Front. Endocrinol. 9, 177 (2018).

    Article  Google Scholar 

  15. El-Hattab, A. W. & Scaglia, F. Mitochondrial DNA depletion syndromes: review and updates of genetic basis, manifestations, and therapeutic options. Neurotherapeutics 10, 186–198 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Meuth, M., L’Heureux-Huard, N. & Trudel, M. Characterization of a mutator gene in Chinese hamster ovary cells. Proc. Natl Acad. Sci. USA 76, 6505–6509 (1979).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Weinberg, G., Ullman, B. & Martin, D. W. Mutator phenotypes in mammalian cell mutants with distinct biochemical defects and abnormal deoxyribonucleoside triphosphate pools. Proc. Natl Acad. Sci. USA 78, 2447–2451 (1981).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Schaich, M. A. et al. Mechanisms of nucleotide selection by telomerase. eLife 9, e55438 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Maine, I. P., Chen, S. F. & Windle, B. Effect of dGTP concentration on human and CHO telomerase. Biochemistry 38, 15325–15332 (1999).

    Article  CAS  PubMed  Google Scholar 

  20. Sun, D., Lopez-Guajardo, C. C., Quada, J., Hurley, L. H. & Von Hoff, D. D. Regulation of catalytic activity and processivity of human telomerase. Biochemistry 38, 4037–4044 (1999).

    Article  CAS  PubMed  Google Scholar 

  21. Chen, Y., Podlevsky, J. D., Logeswaran, D. & Chen, J. J.-L. A single nucleotide incorporation step limits human telomerase repeat addition activity. EMBO J. 37, e97953 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Hwang, H., Opresko, P. & Myong, S. Single-molecule real-time detection of telomerase extension activity. Sci. Rep. 4, 6391 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Hardy, C. D., Schultz, C. S. & Collins, K. Requirements for the dGTP-dependent repeat addition processivity of recombinant Tetrahymena telomerase. J. Biol. Chem. 276, 4863–4871 (2001).

    Article  CAS  PubMed  Google Scholar 

  24. Hammond, P. W. & Cech, T. R. dGTP-dependent processivity and possible template switching of euplotes telomerase. Nucleic Acids Res. 25, 3698–3704 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Doench, J. G. et al. Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR–Cas9. Nat. Biotechnol. 34, 184–191 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Baerlocher, G. M., Vulto, I., de Jong, G. & Lansdorp, P. M. Flow cytometry and FISH to measure the average length of telomeres (flow FISH). Nat. Protoc. 1, 2365–2376 (2006).

    Article  CAS  PubMed  Google Scholar 

  27. Alder, J. K. et al. Diagnostic utility of telomere length testing in a hospital-based setting. Proc. Natl Acad. Sci. USA 115, E2358–E2365 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Li, W. et al. MAGeCK enables robust identification of essential genes from genome-scale CRISPR/Cas9 knockout screens. Genome Biol. 15, 554 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  29. De Lange, T. Shelterin: the protein complex that shapes and safeguards human telomeres. Genes Dev. 19, 2100–2110 (2005).

    Article  PubMed  Google Scholar 

  30. Loayza, D. & de Lange, T. POT1 as a terminal transducer of TRF1 telomere length control. Nature 423, 1013–1018 (2003).

    Article  CAS  PubMed  Google Scholar 

  31. Lee, T. H. et al. Essential role of Pin1 in the regulation of TRF1 stability and telomere maintenance. Nat. Cell Biol. 11, 97–105 (2009).

    Article  CAS  PubMed  Google Scholar 

  32. O’Connor, M. S., Safari, A., Liu, D., Qin, J. & Songyang, Z. The human Rap1 protein complex and modulation of telomere length. J. Biol. Chem. 279, 28585–28591 (2004).

    Article  PubMed  Google Scholar 

  33. Wang, B. et al. Integrative analysis of pooled CRISPR genetic screens using MAGeCKFlute. Nat. Protoc. 14, 756–780 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Li, C. et al. Genome-wide association analysis in humans links nucleotide metabolism to leukocyte telomere length. Am. J. Hum. Genet. 106, 389–404 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Tummala, H. et al. Germline thymidylate synthase deficiency impacts nucleotide metabolism and causes dyskeratosis congenita. Am. J. Hum. Genet. 109, 1472–1483 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Reichard, P. Interactions between deoxyribonucleotide and DNA synthesis. Ann. Rev. Biochem. 57, 349–374 (1988).

    Article  CAS  PubMed  Google Scholar 

  37. Martínez-Arribas, B. et al. DCTPP1 prevents a mutator phenotype through the modulation of dCTP, dTTP and dUTP pools. Cell. Mol. Life Sci. 77, 1645–1660 (2020).

    Article  PubMed  Google Scholar 

  38. Goldstone, D. C. et al. HIV-1 restriction factor SAMHD1 is a deoxynucleoside triphosphate triphosphohydrolase. Nature 480, 379–382 (2011).

    Article  CAS  PubMed  Google Scholar 

  39. Franzolin, E. et al. The deoxynucleotide triphosphohydrolase SAMHD1 is a major regulator of DNA precursor pools in mammalian cells. Proc. Natl Acad. Sci. USA 110, 14272–14277 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Powell, R. D., Holland, P. J., Hollis, T. & Perrino, F. W. Aicardi–Goutieres syndrome gene and HIV-1 restriction factor SAMHD1 is a dGTP-regulated deoxynucleotide triphosphohydrolase. J. Biol. Chem. 286, 43596–43600 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Jamburuthugoda, V. K., Chugh, P. & Kim, B. Modification of human immunodeficiency virus type 1 reverse transcriptase to target cells with elevated cellular dNTP concentrations. J. Biol. Chem. 281, 13388–13395 (2006).

    Article  CAS  PubMed  Google Scholar 

  42. Yuan, M., Breitkopf, S. B., Yang, X. & Asara, J. M. A positive/negative ion-switching, targeted mass spectrometry-based metabolomics platform for bodily fluids, cells, and fresh and fixed tissue. Nat. Protoc. 7, 872–881 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Lahouassa, H. et al. SAMHD1 restricts HIV-1 by reducing the intracellular pool of deoxynucleotide triphosphates. Nat. Immunol. 13, 223–228 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Majerska, J., Feretzaki, M., Glousker, G. & Lingner, J. Transformation-induced stress at telomeres is counteracted through changes in the telomeric proteome including SAMHD1. Life Sci. Alliance 1, e201800121 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Daddacha, W. et al. SAMHD1 promotes DNA end resection to facilitate DNA repair by homologous recombination. Cell Rep. 20, 1921–1935 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Coquel, F. et al. SAMHD1 acts at stalled replication forks to prevent interferon induction. Nature 557, 57–61 (2018).

    Article  CAS  PubMed  Google Scholar 

  47. Beloglazova, N. et al. Nuclease activity of the human SAMHD1 protein implicated in the Aicardi–Goutières syndrome and HIV-1 restriction. J. Biol. Chem. 288, 8101–8110 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Morris, E. R. et al. Crystal structures of SAMHD1 inhibitor complexes reveal the mechanism of water-mediated dNTP hydrolysis. Nat. Commun. 11, 3165 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Xeros, N. Deoxyriboside control and synchronization of mitosis. Nature 194, 682–683 (1962).

    Article  CAS  PubMed  Google Scholar 

  50. Bootsma, D., Budke, L. & Vos, O. Studies on synchronous division of tissue culture cells initiated by excess thymidine. Exp. Cell Res. 33, 301–309 (1964).

    Article  CAS  PubMed  Google Scholar 

  51. Lee, S. S., Bohrson, C., Pike, A. M., Wheelan, S. J. & Greider, C. W. ATM kinase is required for telomere elongation in mouse and human cells. Cell Rep. 13, 1623–1632 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Tong, A. S. et al. ATM and ATR signaling regulate the recruitment of human telomerase to telomeres. Cell Rep. 13, 1633–1646 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Cristofari, G. & Lingner, J. Telomere length homeostasis requires that telomerase levels are limiting. EMBO J. 25, 565–574 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Morin, G. B. The human telomere terminal transferase enzyme is a ribonucleoprotein that synthesizes TTAGGG repeats. Cell 59, 521–529 (1989).

    Article  CAS  PubMed  Google Scholar 

  56. Yamaguchi, M., Hendrickson, E. A. & DePamphilis, M. L. DNA primase-DNA polymerase alpha from simian cells: sequence specificity of initiation sites on simian virus 40 DNA. Mol. Cell. Biol. 5, 1170–1183 (1985).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Roth, Y.-F. Eucaryotic primase. Eur. J. Biochem. 165, 473–481 (1987).

    Article  CAS  PubMed  Google Scholar 

  58. Desmarais, J. A., Unger, C., Damjanov, I., Meuth, M. & Andrews, P. Apoptosis and failure of checkpoint kinase 1 activation in human induced pluripotent stem cells under replication stress. Stem Cell Res. Ther. 7, 17 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Tummala, H., Walne, A. & Dokal, I. The biology and management of dyskeratosis congenita and related disorders of telomeres. Expert Rev. Hematol. 15, 685–696 (2022).

    Article  CAS  PubMed  Google Scholar 

  60. Gupta, A. et al. Telomere length homeostasis responds to changes in intracellular dNTP pools. Genetics 193, 1095–1105 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Maicher, A. et al. Rnr1, but not Rnr3, facilitates the sustained telomerase-dependent elongation of telomeres. PLoS Genet. 13, e1007082 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Jordheim, L. P., Durantel, D., Zoulim, F. & Dumontet, C. Advances in the development of nucleoside and nucleotide analogues for cancer and viral diseases. Nat. Rev. Drug Discov. 12, 447–464 (2013).

    Article  CAS  PubMed  Google Scholar 

  63. Broen, J. C. A. & van Laar, J. M. Mycophenolate mofetil, azathioprine and tacrolimus: mechanisms in rheumatology. Nat. Rev. Rheumatol. 16, 167–178 (2020).

    Article  CAS  PubMed  Google Scholar 

  64. Domínguez-González, C. et al. Deoxynucleoside therapy for thymidine kinase 2-deficient myopathy. Ann. Neurol. 86, 293–303 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  65. Meyerhans, A. et al. Restriction and enhancement of human immunodeficiency virus type 1 replication by modulation of intracellular deoxynucleoside triphosphate pools. J. Virol. 68, 535–540 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Moon, D. H. et al. Poly(A)-specific ribonuclease (PARN) mediates 3′-end maturation of the telomerase RNA component. Nat. Genet. 47, 1482–1488 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Agarwal, S. et al. Telomere elongation in induced pluripotent stem cells from dyskeratosis congenita patients. Nature 464, 292–296 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Paulsen, B. S. et al. Ectopic expression of RAD52 and dn53BP1 improves homology-directed repair during CRISPR–Cas9 genome editing. Nat. Biomed. Eng. 1, 878–888 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Warlich, E. et al. Lentiviral vector design and imaging approaches to visualize the early stages of cellular reprogramming. Mol. Ther. 19, 782–789 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Sanson, K. R. et al. Optimized libraries for CRISPR–Cas9 genetic screens with multiple modalities. Nat. Commun. 9, 5416 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Nagpal, N. et al. Small-molecule PAPD5 inhibitors restore telomerase activity in patient stem cells. Cell Stem Cell 26, 896–909.e8 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Lai, T.-P., Wright, W. E. & Shay, J. W. Generation of digoxigenin-incorporated probes to enhance DNA detection sensitivity. Biotechniques 60, 306–309 (2016).

    Article  CAS  PubMed  Google Scholar 

  73. Lyčka, M. et al. WALTER: an easy way to online evaluate telomere lengths from terminal restriction fragment analysis. BMC Bioinformatics 22, 145 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  74. Herbert, B.-S., Hochreiter, A. E., Wright, W. E. & Shay, J. W. Nonradioactive detection of telomerase activity using the telomeric repeat amplification protocol. Nat. Protoc. 1, 1583–1590 (2006).

    Article  CAS  PubMed  Google Scholar 

  75. Traut, T. W. Physiological concentrations of purines and pyrimidines. Mol. Cell. Biochem. 140, 1–22 (1994).

    Article  CAS  PubMed  Google Scholar 

  76. Mannherz, W. TRAP nanopore analysis. Zenodo https://doi.org/10.5281/zenodo.7607615 (2023).

  77. Wagih, O. ggseqlogo: a versatile R package for drawing sequence logos. Bioinformatics 33, 3645–3647 (2017).

    Article  CAS  PubMed  Google Scholar 

  78. Kimura, M. & Aviv, A. Measurement of telomere DNA content by dot blot analysis. Nucleic Acids Res. 39, e84 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the patients and families for research participation. We thank R. Mathieu and the Harvard Stem Cell Institute (HSCI)-Boston Children’s Hospital (BCH) Flow Cytometry Research Lab; the Molecular Biology Core Facilities at the Dana-Farber Cancer institute (for high-throughput sequencing support); J. Asara and the Beth Israel Deaconess Medical Center Mass Spectrometry Core Facility; and A. Shimamura, M. Fleming and the BCH Bone Marrow Failure/Myelodysplastic Syndrome Registry (National Institutes of Health (NIH) grant R21DK099808). We thank L. Zon and Y. Fong for critical input. We thank A. Gutierrez and K. Bodaar for guidance and support with the CRISPR–Cas9 screening. We thank D. Moon for generating the TERC-null 293T cell line. We acknowledge the following funding sources: NIH grants T32GM007226, T32GM007753 and T32GM144273 (to W.M.), NIH grants R01DK107716 and R33HL154133 (to S.A.), the BCH Translational Research Program, HSCI, Team Telomere, the Million Dollar Bike Ride/Penn Medicine Orphan Disease Center and philanthropic gifts (to S.A.). This content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.

Author information

Authors and Affiliations

Authors

Contributions

S.A. and W.M. conceived of the study and designed the experiments. W.M. performed the experiments and analyzed the data. S.A. and W.M. wrote the manuscript.

Corresponding author

Correspondence to Suneet Agarwal.

Ethics declarations

Competing interests

S.A. and W.M. are named as inventors on provisional patent application 63/394,588 relating to the data shown.

Peer review

Peer review information

Nature Genetics thank Jens Schmidt, Tracy Bryan and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Telomere length CRISPR/Cas9 screening using flow-FISH.

a-c, Histogram of GFP fluorescence from K562 cells (a), K562 cells transduced with the pXPR-011 vector which expresses eGFP and an sgRNA targeting eGFP (b), and K562 cells expressing Cas9 and transduced with the pXPR-011 vector, 13 days post transduction (c). Presence of GFP-negative cells in c indicates functional Cas9 nuclease activity. d-g, Representative gating strategy for flow-FISH telomere length screening. Data from nucleotide metabolism library infected K562 cells, replicate 1. Cells are gated to enrich for single cells (e, f), and gated on low DAPI fluorescence to enrich for cells with 2 N genome copy number and aid in identifying sgRNAs which promote telomere elongation independent from changes in total DNA content (g) followed by gating on high and low TelC-Alexa 647 probe fluorescence populations. Gates adjusted to maintain approximately 5% of cells throughout the duration of the sort. h, i, sgRNA enrichment in high (h) and low (i) telomere fluorescence populations compared to unsorted populations from K562 cells expressing Cas9 that were infected with the Brunello sgRNA library and then cultured for 49 days followed by flow-FISH sorting of the 5% of cells with the highest and lowest telomere fluoresence in two replicates performed on consecutive days. Enrichment score calculated using the MAGeCK RRA software. Known telomere length regulating genes indicated with orange dots; other genes indicated are involved in nucleotide metabolism. j, KEGG pathway enrichment analysis performed on the genes with sgRNAs enriched in the sorted short telomere population (i), analysis performed using the MAGeCKFlute software package (see Methods). Plot includes top enriched KEGG terms, plotting -log10 adjusted P value, which includes q-value estimation for false discovery rate control; dot size indicates number of genes identified in that pathway out of the short telomere enriched genes.

Extended Data Fig. 2 Characterization of TERC-null 293T cells.

a, Schematic of TERC genotypes in TERC-null 293T cells generated by genome-editing, including a deletion of the essential box H domain on one allele, and an 821-bp TERC locus deletion that encompasses 74 bp from the 3’ end of TERC on the other allele. b, Ethidium bromide stained agarose gel of PCR of 293T or 293T TERC-null genomic DNA using primers flanking the deletions indicated in a. c, Sanger sequencing of gel-purified PCR products from the (1) higher molecular weight bands in b, indicating that the non-deleted allele lacks the box H domain, and (2) the ∆821 bp deleted band from b, with trace file showing the deletion junction in a genomic context. d, RT-qPCR of TERC expression relative to GAPDH in wild-type 293T and TERC-null 293T cells, performed in technical triplicate. P value calculated by unpaired t test. Data are shown are means and error bars indicate standard deviation. e, Telomerase activity measured via the TRAP assay, performed on 5-fold serial dilutions of lysates. HI indicates heat-inactivated lysate. IC indicates the internal control product. f, TRF of wild-type and TERC-null 293T cells. Days of culture were recorded beginning approximately two months after gene editing. Telomere length gradually declines with passage until cells universally senesce. g, Quantification of f, line fit using simple linear regression. Data presented in this figure are the results of single experiments unless otherwise indicated.

Extended Data Fig. 3 dT nucleotide metabolism perturbations and their effects on telomere length and polar metabolite homeostasis.

a, TRF of 293T cells treated with the indicated dose of dT for 10 days. The 0 µM dT lane is the same image as the rightmost lane in Fig. 2f. Manufacturer 2, Santa Cruz Biotechnology; Manufacturer 3, MP Biomedicals. b,c, Genomic DNA from 293T or K562 cells manipulated with the indicated sgRNA(s) followed by dT treatment was PCR amplified using primers specific to the TK1 (b) or TK2 (c) genomic loci. Amplicons were separated by agarose gel electrophoresis, demonstrating the three pooled sgRNAs targeting either TK1 or TK2 generated on-target genomic deletions. First lane is a molecular weight marker. PCR products were Sanger sequenced and editing efficiency was quantified using the Synthego ICE algorithm (shown as ‘Knockout Score’). The model fit of the ICE quantification is also displayed. Genomic DNA used in this figure was the same as the DNA used to generate Fig. 3b–e. d, Quantification of Fig. 3b. n = 3 biological replicates for each cell line, P value calculated using paired two-sided t test. e, Polar metabolite profiling by liquid chromatography mass spectrometry of 293 T cells treated with or without 100 µM dT for 24 hours, performed in biological triplicate. P value calculated by unpaired two-sided Student’s t test of average signal intensity in treatment vs. control samples; nucleotide and nucleoside species detected in all samples displayed. Note: dGTP not detected. f, TRF of 293T cells treated with the indicated compound for 10 days. dU, deoxyuridine. g, Detection of CRISPR/Cas9 editing of TYMS locus using TYMS-specific primers, performed as in b and c. Genomic DNA used was the same as the DNA used to generate data in Fig. 3g–j. Data in d and e are shown as means, and error bars indicate standard deviation.

Extended Data Fig. 4 Manipulation of SAMHD1 levels by CRISPR/Cas9, shRNA, and lentiviral expression.

a, Immunoblot of 293T and K562 cells electroporated with Cas9 and the indicated sgRNA(s) using primary antibodies against SAMHD1 or β-Actin, corresponding to cell lines evaluated in Fig. 4b. b, Immunoblot of K562 cells transduced with vectors expressing the indicated shRNA, corresponding to cell lines evaluated in Fig. 4c. c, qRT-PCR of SAMHD1 expression compared to β-Actin, performed in technical triplicate. Means of the replicates are shown. d, Immunoblot of 293T cells transduced with vectors expressing the indicated shRNA. e, TRF of indicated cell lines transduced with the indicated shRNA and cultured for 15 days, and quantification of the TRF using the WALTER webtool. The boxplot displays the 75th, 50th and 25th percentile molecular weight of the telomere signal distribution in the TRF blot. f, Immunoblot of 293T cells transduced with vectors to overexpress either eGFP or the indicated SAMHD1 variant and treated with the indicated dose of dT, corresponding to cell lines evaluated in Fig. 4i, j. Data presented in this figure are the results of single experiments unless otherwise stated. Full-length western blots are presented as source data.

Source data

Extended Data Fig. 5 Cell cycle assessment of cells treated with dT.

ad, DAPI staining of 293T (a, b) and K562 (c, d) cells treated with the indicated dose of dT for 7 or 8 days, respectively, measured by flow cytometry, plotted as histograms of DAPI intensity, displaying representative samples from each treatment arm (a, c), and the percentage of cells in different stages of the cell cycle (b, d) gated based on the lines drawn on the histogram, gates determined based on untreated samples. n = 2 biological replicates for 293T cells treated with 200 µM dT and n = 3 biological replicates for all other conditions. Data presented are means; error bars indicate standard deviation.

Extended Data Fig. 6 Evaluation of telomere length and cell cycle progression changes from treatment with dT, aphidicolin, 5FU, or hydroxyurea.

a, TRF Southern blot of 293T cells treated with the indicated doses of aphidicolin and hydroxyurea for 10 days. b, TRF Southern blot of 293T TERC-null cells transfected with TERT in addition to the indicated vector, cultured for 18 hours, then treated with the indicated dose of dT for 30 hours. c, TRF Southern blot of 293T TERC-null cells transfected with TERT in addition to the indicated vector, cultured for 18 hours, then treated with the indicated dose of dT for two days. d, TRF Southern blot of 293T TERC-null cells transfected with the indicated expression vectors, cultured for 18 hours, then treated with the indicated dose of dT for five days. eh, Cell cycle analysis by DAPI staining and flow cytometry of 293T TERC-null cells transfected with TERC and TERT expression vectors, cultured for 18 hours, then treated with the indicated of dose of dT (e), aphidicolin (f), 5FU (g), or hydroxyurea (h), displayed as histograms of DAPI intensity of representative samples from each treatment arm, corresponding to cells in Fig. 6b–l. Gating based on untreated cells. TRFs presented in this figure show the results of single experiments.

Extended Data Fig. 7 T-free telomerase is sensitive to dT nucleotide manipulations.

a, Representative modified TRAP assay performed on super-telomerase extracts using the indicated dose of dTTP and physiologic levels of dATP, dCTP and dGTP (see Methods). b, Quantification of a. n = 2 biological replicates. c, GGAAAG TRAP assay performed on lysates from 293T TERC-null cells overexpressing T-free super-telomerase demonstrates linearity between cell input amount and telomerase signal. Five-fold serial dilutions performed. HI, heat inactivated. d, Quantification of lanes 1–3 from c. e, Representative modified GGAAAG TRAP assay performed on super-telomerase extracts generated using the indicated TERC vector. Assay performed with the indicated dose of dTTP and physiologic levels of dATP, dCTP and dGTP (see Methods). HI, heat inactivated. f, Quantification of e using two-sided unpaired Student’s t test; n = 3 biological replicates. g, Representative modified GGAAAG TRAP assay performed on T-free super-telomerase extracts supplemented with the indicated dose of dTTP and physiologic levels of dATP and dGTP. h, Quantification of g as in f, n = 3 biological replicates. i, Diagram of GGAAAG TRAP product sequencing and analysis strategy. Note * indicates T’s encoded by the partially complementary reverse primer, preventing analysis of base composition in that portion of the read. j, Quantification of base pair composition of representative GGAAAG TRAP products from g with 0 μM or 25 μM dTTP by nanopore sequencing (see Methods). Bits of information calculated using Shannon entropy and plotted using ggseqlogo. k, Quantification of base pair composition of GGAAAG TRAP products from g using nanopore sequencing (see Methods). P value calculated using two-sided Student’s t test; n = 3 biological replicates. l, Quantification of Fig. 7d, plotting the signal in the indicated telomerase product repeat relative to the signal of the corresponding repeat in the lane without dTTP added, normalized for loading (see Methods). m, TRF Southern blot of 293T TERC-null cells transfected with TERT in addition to the indicated vector, cultured for 18 hours, then treated with the indicated dose of dT for 30 hours, and probed with a GGTTAG complementary probe. Lanes 1–4 are the same blot shown in Extended Data Fig. 6b. n, Blot from m was stripped and re-probed with a probe complementary to the GGAAAG repeat. o, Slot blot of DNA from 293T TERC-null cells overexpressing eGFP and TERT showing linear relationship between DNA input and signal; rows are technical triplicates. p, Quantification of o. q, Slot blot of DNA from 293T TERC-null cells overexpressing T-free super-telomerase; rows are technical triplicates. r, Quantification of q. s, Slot blot of DNA from 293T-TERC null cells transfected with TERT and TERC, cultured for 18 hours, then treated with dT as indicated for 30 hours. Denatured DNA for each sample was split and loaded onto parallel blots, which were probed for the indicated target. Performed in technical triplicate. t, Quantification of s. P values calculated with one way ANOVA using Dunnett’s multiple comparisons test for each probe. For b, d, f, h, k, p, r and t, the mean of the data is presented and error bars indicate s.d.; ns, P > 0.05.

Extended Data Fig. 8 Effects of dT on iPSC cell cycle progression and replication stress signaling.

a, Cell cycle analysis of wild-type iPSCs cultured in the indicated of dose of dT for 24 hours, measured by DAPI staining and flow cytometry, displayed as histograms of DAPI intensity. n = 2 biological replicates; the mean of the replicates is presented. b, Representative histograms of DAPI signal for cells in a. Gates defined based on untreated cells. c, Immunoblot of cells treated as in a; all images of the same membrane blotted with the indicated primary antibodies. UV- treated cells used as a positive control. Blot shows the results from a single experiment. Full-length blots are provided as source data.

Source data

Extended Data Fig. 9 Model of relationship between dT nucleotide metabolism and telomere synthesis.

a-e, Schematics illustrate conditions of homeostasis (a), dT supplementation (b), dU supplementation (c), loss of SAMHD1 (d), and treatment with hydroxyurea or 5-fluorouracil (e).

Supplementary information

Reporting Summary

Supplementary Tables

Supplementary Tables 1–8.

Source data

Source Data Fig. 6

Unprocessed western blots.

Source Data Fig. 7

Unprocessed gels.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 8

Unprocessed western blots.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mannherz, W., Agarwal, S. Thymidine nucleotide metabolism controls human telomere length. Nat Genet 55, 568–580 (2023). https://doi.org/10.1038/s41588-023-01339-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-023-01339-5

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