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.

  • Review Article
  • Published:

Deoxyribonucleotide metabolism, mutagenesis and cancer

Key Points

  • Eukaryotic cells contain two distinct but interrelated pools of DNA precursors for the replication of nuclear and mitochondrial DNA. Deoxyribonucleoside triphosphate (dNTP) biosynthesis begins either with the reduction of ribonucleotides or with the salvage of preformed nucleosides or nucleobases.

  • dNTP pool sizes are regulated mostly at the level of biosynthesis, particularly involving ribonucleotide reductase. However, controlled dNTP degradation brought about by the sterile α-motif and histidine-aspartate domain-containing protein (SAMHD1) protein has recently emerged as a major regulatory mechanism.

  • An increase in the spontaneous mutation rate is almost certainly an essential feature of carcinogenesis. Abnormal regulation of dNTP pool sizes contributes to determination of the mutation rate.

  • Several oncogenes and tumour suppressors control dNTP pool sizes and have as-yet-unexplained effects on oncogene-induced senescence.

  • Much evidence indicates that 'sanitation', the enzymatic hydrolysis of abnormal or damaged nucleotides, is important to the maintenance of genomic stability. However, recent evidence indicates that one sanitizing enzyme, MTH1, is an important target for inhibition in cancer treatment.

  • Many anticancer drugs achieve their effectiveness because they are analogues to DNA precursors. Effective use of these analogues requires a thorough understanding of nucleotide biosynthesis, transport processes, cell cycle regulation and interactions with target enzymes.

Abstract

Cancer was recognized as a genetic disease at least four decades ago, with the realization that the spontaneous mutation rate must increase early in tumorigenesis to account for the many mutations in tumour cells compared with their progenitor pre-malignant cells. Abnormalities in the deoxyribonucleotide pool have long been recognized as determinants of DNA replication fidelity, and hence may contribute to mutagenic processes that are involved in carcinogenesis. In addition, many anticancer agents antagonize deoxyribonucleotide metabolism. Here, we consider the extent to which aspects of deoxyribonucleotide metabolism contribute to our understanding of both carcinogenesis and to the effective use of anticancer agents.

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: Timeline of important events that have developed our understanding of deoxyribonucleotide metabolism, mutagenesis and cancer, and their relationships.
Figure 2: Pathways de novo dNTP biosynthesis in mammalian cells.
Figure 3: Principal reactions in salvage routes to deoxyribonucleotides in human cells.
Figure 4: The thymidylate synthesis cycle.
Figure 5: Replication errors leading to mutations induced by dNTP pool imbalances.

Similar content being viewed by others

References

  1. Cohen, S. S. & Barner, H. D. Studies on unbalanced growth in Escherichia coli. Proc. Natl Acad. Sci. USA 40, 885–893 (1954).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Cohen, S. S., Flaks, J. G., Barner, H. D., Loeb, M. R. & Lichtenstein, J. The mode of action of 5-fluorouracil and its derivatives. Proc. Natl Acad. Sci. USA 44, 1004–1012 (1958).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Hitchings, G. H. Chemotherapy and comparative biochemistry: G. H. A. Clowes memorial lecture. Cancer Res. 29, 1895–1903 (1969). This is a personal account of the early era of target seeking for disease treatment in nucleotide metabolism.

    CAS  PubMed  Google Scholar 

  4. Kunz, B. A. et al. Deoxyribonucleoside triphosphate levels: a critical factor in the maintenance of genetic stability. Mutat. Res. 318, 1–64 (1994).

    CAS  PubMed  Google Scholar 

  5. Meuth, M. The molecular basis of mutations induced by deoxyribonucleoside pool imbalances in mammalian cells. Exp. Cell Res. 181, 305–316 (1989).

    CAS  PubMed  Google Scholar 

  6. Mathews, C. K. DNA precursor metabolism and genomic stability. FASEB J. 20, 1300–1314 (2006).

    CAS  PubMed  Google Scholar 

  7. Bestwick, R. K., Moffett, G. L. & Mathews, C. K. Selective expansion of mitochondrial deoxyribonucleoside triphosphate pools in antimetaboite-treated HeLa cells. J. Biol. Chem. 257, 9300–9304 (1982).

    CAS  PubMed  Google Scholar 

  8. Rampazzo, C. et al. Mitochondrial deoxyribonucleotides, pool sizes, synthesis, and regulation. J. Biol. Chem. 279, 17019–17026 (2004).

    CAS  PubMed  Google Scholar 

  9. Ferraro, P. et al. Mitochondrial deoxynucleotide pools in quiescent fibroblasts. J. Biol. Chem. 280, 24472–24480 (2005).

    CAS  PubMed  Google Scholar 

  10. Di Noia, M. A. et al. The human SLC25A33 and SLC25A36 genes of solute carrier family 25 encode two mitochondrial pyrimidine nucleotide transporters. J. Biol. Chem. 289, 33137–33148 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Weinberg, S. E. & Chandel, N. S. Targeting mitochondrial metabolism for cancer therapy. Nat. Chem. Biol. 11, 9–15 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Leeds, J. M., Slabaugh, M. B. & Mathews, C. K. DNA precursor pools and ribonucleotide reductase activity: distribution between the nucleus and cytoplasm of mammalian cells. Mol. Cell. Biol. 5, 3443–3450 (1985).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Poli, J. et al. dNTP pools determine fork progression and origin usage under replication stress. EMBO J. 31, 883–894 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Nordlund, P. & Reichard, P. Ribonucleotide reductases. Ann. Rev. Biochem. 75, 681–706 (2006).

    CAS  PubMed  Google Scholar 

  15. Stubbe, J. Ribonucleotide reductases. Adv. Enzymol. 63, 349–419 (1990).

    CAS  PubMed  Google Scholar 

  16. Tanaka, H. et al. A ribonucleotide reductase gene involved in a p53-dependent cell-cycle checkpoint for DNA damage. Nature 404, 42–49 (2000).

    CAS  PubMed  Google Scholar 

  17. Guittet, O. et al. Mammalian p53R2 protein forms an active ribonucleotide reductase in vitro with the R1 protein, which is expressed both in resting cells in response to DNA damage and in proliferating cells. J. Biol. Chem. 276, 40647–40651 (2001).

    CAS  PubMed  Google Scholar 

  18. Shao, J. et al. In vitro characterization of enzymatic properties and inhibition of the p53R2 subunit of human ribonucleotide reductase. Cancer Res. 64, 1–6 (2004).

    CAS  PubMed  Google Scholar 

  19. Bourdon, A. et al. Mutation of RRM2B, encoding p53-controlled ribonucleotide reductase (p53R2), causes severe mitochondrial depletion. Nat. Genet. 39, 776–780 (2007).

    CAS  PubMed  Google Scholar 

  20. Pontarin, G., Ferraro, P., Bee, L., Reichard, P. & Bianchi, V. Mammalian ribonucleotide reductase subunit p53R2 is required for mitochondrial DNA replication and DNA repair in quiescent cells. Proc. Natl Acad. Sci. USA 109, 13302–13307 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Chimploy, K. & Mathews, C. K. Mouse ribonucleotide reductase control. Influence of substrate binding upon interactions with allosteric inhibitors. J. Biol. Chem. 276, 7093–7100 (2001).

    CAS  PubMed  Google Scholar 

  22. Fairman, J. W. et al. Structural basis for allosteric regulation of human ribonucleotide reductase by nucleotide-induced oligomerization. Nat. Struct. Mol. Biol. 18, 316–322 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Martomo, S. A. & Mathews, C. K. Effects of biological DNA precursor pool asymmetry upon accuracy of DNA replication in vitro. Mutat. Res. 499, 197–211 (2002).

    CAS  PubMed  Google Scholar 

  24. Wheeler, L. J. & Mathews, C. K. Nucleoside triphosphate pool asymmetry in mammalian mitochondria. J. Biol. Chem. 286, 16992–16996 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Postel, E. H., Berberich, S. J., Flint, S. J. & Ferrone, C. A. Human c-myc transcription factor PuF identified as nm23-H2 nucleoside diphosphate kinase, a candidate suppressor of tumor metastasis. Science 261, 478–481 (1993).

    CAS  PubMed  Google Scholar 

  26. Postel, E. H. Cleavage of DNA by human NM23-H2/nucleoside diphosphate kinase involves formation of a covalent protein–DNA complex. J. Biol. Chem. 274, 22821–22829 (1999).

    CAS  PubMed  Google Scholar 

  27. Engström, Y. & Rozell, B. Immunocytochemical evidence for the cytoplasmic localization and differential expression during the cell cycle of the M1 and M2 subunits of mammalian ribonucleotide reductase. EMBO J. 7, 1615–1620 (1988).

    PubMed  PubMed Central  Google Scholar 

  28. Pontarin, G. et al. Ribonucleotide reduction is a cytosolic process in mammalian cells independently of DNA damage. Proc. Natl Acad. Sci. USA 105, 17801–17806 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Niida, H. et al. Essential role of Tip60-dependent recruitment of ribonucleotide reductase at DNA damage sites in DNA repair during G1 phase. Genes Dev. 24, 333–338 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Hu, C.-M. et al. Tumor cells require thymidylate kinase to prevent dUTP incorporation during DNA repair. Cancer Cell 22, 36–50 (2012). This paper identifies thymidylate kinase as an under-investigated but important therapeutic target.

    CAS  PubMed  Google Scholar 

  31. MacFarlane, A. et al. Nuclear localization of de novo thymidylate biosynthesis pathway is required to prevent uracil accumulation in DNA. J. Biol. Chem. 286, 44015–44022 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Anderson, D. D., Eom, J. Y. & Stover, P. J. Competition between sumoylation and ubiquitination of serine hydroxymethyltransferase 1 determines its nuclear localization and its accumulation in the nucleus. J. Biol. Chem. 287, 4790–4799 (2012).

    CAS  PubMed  Google Scholar 

  33. Anderson, D. D., Woeller, C. F., Chiang, E.-P., Shane, B. & Stover, P. J. Serine hydroxymethyltransferase anchors de novo thymidylate synthesis pathway to nuclear lamina for DNA synthesis. J. Biol. Chem. 287, 7051–7062 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Field, M. S. et al. Nuclear enrichment of folate cofactors and methylenetetrahydrofolate dehydrogenase 1 (MTHFD1) protect de novo thymidylate biosynthesis during folate deficiency. J. Biol. Chem. 289, 29642–29650 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Field, M. S., Kamynina, E., Watkins, D., Rosenblatt, D. S. & Stover, P. J. Human mutations in methylenetetrahydrofolate dehydrogenase 1 impair nuclear de novo thymidylate biosynthesis. Proc. Natl Acad. Sci. USA 112, 400–405 (2015).

    CAS  PubMed  Google Scholar 

  36. Chabes, A. L., Björklund, S. & Thelander, L. S. Phase-specific transcription of the mouse ribonucleotide reductase R2 gene requires both a proximal repressive EF2-binding site and an upstream promoter activating region. J. Biol. Chem. 279, 10796–10807 (2004).

    CAS  PubMed  Google Scholar 

  37. Zhang, Y.-W. et al. Implication of checkpoint kinase-dependent up-regulation of ribonucleotide reductase R2 in DNA damage response. J. Biol. Chem. 284, 18085–18095 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Salguero, I. et al. Ribonucleotide reductase activity is coupled to DNA synthesis via proliferating cell nuclear antigen. Curr. Biol. 22, 720–726 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Chabes, A. & Thelander, L. Controlled protein degradation regulates ribonucleotide reductase activity in proliferating mammalian cells during the normal cell cycle and in response to DNA damage and replication blocks. J. Biol. Chem. 275, 17747–17753 (2000).

    CAS  PubMed  Google Scholar 

  40. Bianchi, V. et al. Cell cycle-dependent metabolism of pyrimidine deoxynucleoside triphosphates in CEM cells. J. Biol. Chem. 272, 16118–16124 (1997).

    CAS  PubMed  Google Scholar 

  41. Rampazzo, C. et al. Regulation by degradation, a cellular defense against deoxyribonucleotide pool imbalances. Mut. Res. 703, 2–10 (2010).

    CAS  Google Scholar 

  42. Tzoneva, G. et al. Activating mutations in the NT5C2 nucleotidase gene drive chemotherapy resistance in relapsed ALL. Nature Med. 19, 368–371 (2013).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  45. Laguette, N. et al. SAMHD1 is the dendritic and myeloid-cell-specific HIV-1 restriction factor counteracted by Vpx. Nature 474, 654–657 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Clifford, R. et al. SAMHD1 is mutated recurrently in chronic lymphocytic leukemia and is involved in response to DNA damage. Blood 123, 1021–1031 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Rossi, D. SAMHD1: a new gene for CLL. Blood 123, 951–952 (2014).

    CAS  PubMed  Google Scholar 

  48. Amie, S. M., Bambara, R. A. & Kim, B. GTP is the primary activator of the anti-HIV restriction factor SAMHD1. J. Biol. Chem. 288, 25001–25006 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Koharudin, L. M. I. et al. Structural basis of allosteric activation of SAMHD1 by nucleoside triphosphates. J. Biol. Chem. 289, 32617–32627 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Ji, X., Tang, C., Zhao, Q., Wang, W. & Xiong, Y. Structural basis of cellular dNTP regulation by SAMHD1. Proc. Natl Acad. Sci. USA 111, E4305–E4314 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Franzolin, E. et al. The deoxynucleotide triphosphohydrolase SAMDH1 is a major regulator of DNA precursor pools in mammalian cells. Proc. Natl Acad. Sci. USA 110, 14272–14277 (2013). This paper presents the first evidence to support a role for SAMHD1 in the normal regulation of dNTP pool sizes.

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Miazzi, C. et al. Allosteric regulation of the human and mouse deoxyribonucleotide triphospholydrolase SAMHD1. J. Biol. Chem. 289, 18339–18346 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Hansen, E. C., Seamon, K. J., Cravens, C. N. & Stivers, J. T. GTP activator and dNTP substrates of HIV-1 restriction factor SAMHD1 generate a long-lived activated state. Proc. Natl Acad. Sci. USA 111, E1843–E1851 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Ryoo, J. et al. The ribonuclease activity of SAMHD1 is required for HIV-1 restriction. Nat. Med. 20, 936–941 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Loeb, L. A., Springgate, C. F. & Battula, N. Errors in DNA replication as a basis of malignant change. Cancer Res. 34, 2311–2321 (1974). This is perhaps the earliest suggestion that the spontaneous mutation rate must increase to explain the many factors that distinguish a tumour from its normal tissue of origin.

    CAS  PubMed  Google Scholar 

  56. Sjöblom, T. et al. The consensus coding sequences of human breast and colorectal cancers. Science 314, 268–274 (2006).

    PubMed  Google Scholar 

  57. Bielas, J. H., Loeb, K. R., Rubin, B. P., True, L. D. & Loeb, L. A. Human cancers express a mutator phenotype. Proc. Natl Acad. Sci. USA 103, 18238–18242 (2006). This paper describes the use of an ultra-sensitive technique for detecting mutations and its use to analyse mutation frequencies in cancer cells.

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Tomasetti, C., Marchionni, L., Nowak, M. A., Parmigiani, V. & Vogelstein, B. Only three driver mutations are required for the development of lung and colorectal cancers. Proc. Natl Acad. Sci. USA 112, 118–123 (2015). This paper expands on a series of studies that define genetic differences between tumour cells and their tissue of origin.

    CAS  PubMed  Google Scholar 

  59. Kennedy, S. R. et al. Volatility of mutator phenotypes at single cell resolution. PLOS Genet. 11,e1005151 (2015).

    PubMed  PubMed Central  Google Scholar 

  60. Tomasetti, C. & Vogelstein, B. Variations in cancer risk among tissues can be explained by the number of stem cell divisions. Science 347, 78–81 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Kumar, D. et al. Mechanisms of mutagenesis in vivo due to imbalanced dNTP pools. Nucleic Acids Res. 39, 1360–1371 (2010).

    PubMed  PubMed Central  Google Scholar 

  62. Ahluwalia, D., Bienstock, R. & Schaaper, R. Novel mutator mutants of E. coli ribonucleotide reductase: insights into allosteric regulation and control of mutation rates. DNA Repair 11, 480–487 (2012). This paper describes a powerful approach both for identifying mutations caused by dNTP imbalance and for understanding the molecular basis for allosteric regulation of RNR.

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Buckland, R. J. et al. Increased and imbalanced dNTP pools symmetrically promote both leading and lagging strand infidelity. PLOS Genet. 10, e10004846 (2014).

    Google Scholar 

  64. Kunkel, T. A. & Burgers, P. M. Dividing the workload at a eukaryotic replication fork. Trends Cell Biol. 18, 521–527 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Johansson, E. & MacNeill, S. A. The eukaryotic replicative DNA polymerases take shape. Trends Biochem. Sci. 35, 339–347 (2010).

    CAS  PubMed  Google Scholar 

  66. St. Charles, J. A., Liberti, S. E., Williams, J. S., Lujan, S. A. & Kunkel, T. A. Quantifying the contributions of base selectivity, proofreading and mismatch repair to nuclear DNA replication in Saccharomyces cerevisiae. DNA Repair 31, 41–51 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Supek, F. & Lehner, B. Differential mismatch repair underlies mutation rate variation across the human genome. Nature 521, 81–84 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Chabes, A. et al. Survival of DNA damage in yeast directly depends on increased dNTP levels allowed by relaxed feedback inhibition of ribonucleotide reductase. Cell 112, 391–401 (2003).

    CAS  PubMed  Google Scholar 

  69. Wheeler, L. J., Rajagopal, I. & Mathews, C. K. Stimulation of mutagenesis by proportional deoxyribonucleoside triphosphate accumulation in Escherichia coli. DNA Repair 4, 1450–1456 (2005).

    CAS  PubMed  Google Scholar 

  70. Davidson, M. B. et al. Endogenous replication stress results in expansion of dNTP pools and a mutator phenotype. EMBO J. 31, 895–907 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Gon, S., Napolitano, R., Rocha, W., Coulton, S. & Fuchs, R. P. Increase in dNTP pool size during the DNA damage response plays a key role in spontaneous and induced mutagenesis in Escherichia coli. Proc. Natl Acad. Sci. USA 108, 19311–19316 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Bester, A. C. et al. Nucleotide deficiency promotes genomic instability in early stage of cancer development. Cell 145, 435–446 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Fearon, E. R. Molecular genetics of colorectal cancer. Annu. Rev. Pathol. 6, 479–507 (2011).

    CAS  PubMed  Google Scholar 

  74. Church, D. N. et al. DNA polymerase ɛ and δ exonuclease domains in endometrial cancer. Hum. Mol. Genet. 22, 2820–2828 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Palles, C. et al. Germline mutations affecting the proofreading domains of POLE and POLD1 predispose toward colorectal adenomas and carcinomas. Nat. Genet. 45, 136–144 (2012).

    PubMed  PubMed Central  Google Scholar 

  76. Mertz, T. M., Sharma, S., Chabes, A. & Shcherbakova, P. V. A colon cancer-associated mutator DNA polymerase δ variant causes expansion of dNTP pools increasing its own infidelity. Proc. Natl Acad. Sci. USA 112, E2467–E2476 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Williams, L. N. et al. dNTP pool levels modulate mutator phenotypes of error-prone DNA polymerase ɛ variants. Proc. Natl Acad. Sci. USA 112, E2457–E2466 (2015). References 76 and 77 use the yeast system to model mutations in human replicative DNA polymerases found in some colorectal and endometrial cancers.

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Chabes, A. & Stillman, B. Constitutively high dNTP concentration inhibits cell cycle progression and the DNA damage checkpoint in yeast Saccharomyces cerevisiae. Proc. Natl Acad. Sci. USA 104, 1183–1188 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Sanvisens, N. et al. Yeast Dun1 kinase regulates ribonucleotide reductase inhibitor Sml1 in response to iron deficiency. Mol. Cell. Biol. 34, 3259–3271 (2014).

    PubMed  PubMed Central  Google Scholar 

  80. Aye, Y., Li, M., Long, M. U. C. & Weiss, R. S. Ribonucleotide reductase and cancer: biological mechanisms and targeted therapies. Oncogene 34, 2011–2021 (2014). This is an extensively referenced review of RNR and its alterations in many cancers.

    PubMed  Google Scholar 

  81. Xu, X. et al. Broad overexpression of ribonucleotide reductase genes in mice specifically induces lung neoplasms. Cancer Res. 68, 2652–2660 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Taricani, L., Shanahan, F., Malinao, M.-C. & Parry, D. A functional approach reveals a genetic and physical interaction between ribonucleotide reductase and CHK1 in mammalian cells. PLoS ONE 9, e111714 (2014).

    PubMed  PubMed Central  Google Scholar 

  83. Åkerblom, L. et al. Overproduction of the free radical of ribonucleotide reductase in hydroxyurea-resistant mouse fibroblast 3T6 cells. Proc. Natl Acad. Sci. USA 78, 2159–2163 (1981).

    PubMed  PubMed Central  Google Scholar 

  84. Wang, J., Lohman, G. J. S. & Stubbe, J. Mechanism of inactivation of human ribonucleotide reductase with p53R2 by gemcitabine 5′-diphosphate. Biochemistry 48, 11612–11621 (2009).

    CAS  PubMed  Google Scholar 

  85. Artin, E. et al. Insight into the mechanism of inactivation of ribonucleotide reductase by gemcitabine 5′-diphosphate in the presence and absence of reductant. Biochemistry 48, 11622–11629 (2009).

    CAS  PubMed  Google Scholar 

  86. Aye, Y. & Stubbe, J. Clofarabine 5′-di- and -triphosphates inhibit human ribonucleotide reductase by altering the quaternary structure of its large subunit. Proc. Natl Acad. Sci. USA 108, 9815–9820 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Chen, M.-C. et al. The novel ribonucleotide reductase inhibitor COH29 inhibits DNA repair in vitro. Mol. Pharmacol. 87, 996–1005 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Zhao, X., Chabes, A., Domkin, V., Thelander, L. & Rothstein, R. The ribonucleotide reductase inhibitor Sml1 is a new target of the Mec1/Rad53 kinase cascade during growth and in response to DNA damage. EMBO J. 20, 3544–3553 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Zhao, X. & Rothstein, R. The Dun1 checkpoint kinase phosphorylates and regulates the ribonucleotide reductase inhibitor Sml1. Proc. Natl Acad. Sci. USA 99, 3746–3751 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  90. Tsaponina, O., Barsoum, E., Åström, S. U. & Chabes, A. Ixr1 is required for the expression of the ribonucleotide reductase Rnr1 and maintenance of dNTP pools. PLOS Genet. 7, e1002061 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  91. Kumar, D., Viberg, J., Nilsson, A. K. & Chabes, A. Highly mutagenic and severely unbalanced dNTP pool can escape detection by the S-phase checkpoint. Nucleic Acids Res. 38, 3975–3983 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Tsaponina, O. & Chabes, A. Pre-activation of the genome integrity checkpoint increases DNA damage tolerance. Nucleic Acids Res. 41, 10371–10378 (2013). Understanding checkpoint control of dNTP metabolism and genomic stability in human cells is one of the most important issues in cancer research.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Angus, S. P. et al. Retinoblastoma tumor suppressor targets dNTP metabolism to regulate DNA replication. J. Biol. Chem. 277, 44376–44384 (2002).

    CAS  PubMed  Google Scholar 

  94. Mannava, S. et al. Direct role of nucleotide metabolism in c-Myc-dependent proliferation of melanoma cells. Cell Cycle 7, 2392–2400 (2008).

    CAS  PubMed  Google Scholar 

  95. Cunningham, J. T., Moreno, M. V., Lodi, A., Ronen, S. M. & Ruggiero, D. Protein and nucleotide biosynthesis are coupled by a single rate-limiting enzyme, PRPS2, to drive cancer. Cell 157, 1008–1103 (2014).

    Google Scholar 

  96. Mannava, S. et al. Depletion of deoxyribonucleotide pools is an endogenous source of DNA damage in cells undergoing oncogene-induced senescence. Am. J. Pathol. 182, 142–150 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Mannava, S. et al. Ribonucleotide reductase and thymidylate synthase or exogenous deoxyribonucleosides reduce DNA damage and senescence caused by C-MYC depletion. Aging 4, 917–922 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  98. Aird, K. M. et al. Suppression of nucleotide metabolism underlies the establishment and maintenance of oncogene-induced senescence. Cell Rep. 3, 1252–1265 (2013).

    CAS  PubMed  Google Scholar 

  99. Aird, K. M. & Zhang, R. Nucleotide metabolism, oncogene-induced senescence, and cancer. Cancer Lett. 356, 204–210 (2015). This recent review describes relationships between nucleotide metabolism and OIS, which may lead to new cancer treatment strategies.

    CAS  PubMed  Google Scholar 

  100. Marusyk, A., Wheeler, L. J., Mathews, C. K. & DeGregori, J. D. p53 mediates senescence-like arrest induced by chronic replication stress. Mol. Cell. Biol. 27, 5336–5351 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Acosta, J. C. & Gil, J. Senescence: a new weapon for cancer therapy. Trends Cell Biol. 22, 211–219 (2012).

    CAS  PubMed  Google Scholar 

  102. Saldivar, J. C. et al. Initiation of genome instability and preneoplastic processes through loss of Fhit expression. PLOS Genet. 8, e1003077 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Xie, M. et al. Bcl2 induces replication stress by inhibiting ribonucleotide reductase. Cancer Res. 74, 212–223 (2014).

    CAS  PubMed  Google Scholar 

  104. Jain, D. & Cooper, J. P. Telomeric strategies: means to an end. Ann. Rev. Genet. 44, 243–269 (2010).

    CAS  PubMed  Google Scholar 

  105. 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).

    CAS  PubMed  Google Scholar 

  106. Hornsby, P. J. Telomerase and the aging process. Exp. Gerontol. 42, 575–581 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Blackburn, E. H. Telomerase and cancer. Mol. Cancer Res. 3, 477–482 (2005).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  109. Gupta, A. et al. Telomere length homeostasis responds to changes in intracellular dNTP pools. Genetics 193, 1095–1105 (2013). A possible link between dNTP and cancer via the regulation of telomere length, as described in this study using yeast, is extremely intriguing.

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Friedberg, E. et al. in DNA Repair and Mutagenesis 2nd edn 17–25 (ASM Press, 2006).

    Google Scholar 

  111. Topal, M. D. & Baker, M. S. DNA precursor pool: a significant target for N-methyl-N-nitrosourea in CH3/T101/2 clone 8 cells. Proc. Natl Acad. Sci. USA 79, 2211–2215 (1982).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. Maki, H. & Sekiguchi, M. MutT protein specifically hydrolyzes a potent mutagenic substrate for DNA synthesis. Nature 355, 273–275 (1992).

    CAS  PubMed  Google Scholar 

  113. Kamiya, H. Mutations caused by oxidized DNA precursors and their prevention by nucleotide pool sanitization enzymes. Genes Environ. 29, 133–140 (2007).

    CAS  Google Scholar 

  114. Mao, J.-Y., Maki, H. & Sekiguchi, M. Hydrolytic elimination of a mutagenic nucleotide, 8-oxodGTP, by human 18-kilodalton protein: sanitization of nucleotide pool. Proc. Natl Acad. Sci. USA 89, 11021–11025 (1992).

    Google Scholar 

  115. Sakai, Y. et al. A molecular basis for the selective recognition of 2-hydroxy-dATP and 8-oxo-dGTP by human MTH-1. J. Biol. Chem. 277, 8579–8587 (2002).

    CAS  PubMed  Google Scholar 

  116. Yoshimura, D. et al. An oxidized purine nucleoside triphosphatase, MTH1, suppresses cell death caused by oxidative stress. J. Biol. Chem. 278, 37965–37973 (2003).

    CAS  PubMed  Google Scholar 

  117. Takagi, Y. et al. Human MTH3 (NUDT18) protein hydrolyzes oxidized forms of guanosine and deoxyguanosine diphosphates. J. Biol. Chem. 287, 21541–21549 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Tsuzuki, T. et al. Spontaneous tumorigenesis in mice defective in the MTH1 gene encoding 8-oxo-dGTPase. Proc. Natl Acad. Sci. USA 98, 11456–11461 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Sakumi, K. et al. Ogg1 knockout-associated lung tumorigenesis and its suppression by Mth1 gene disruption. Cancer Res. 63, 902–905 (2003).

    CAS  PubMed  Google Scholar 

  120. Tassotto, M.-L. & Mathews, C. K. Assessing the metabolic function of the MutT 8-oxodeoxyguanosine triphosphatase in Escherichia coli by nucleotide pool analysis. J. Biol. Chem. 277, 15807–15812 (2002).

    CAS  PubMed  Google Scholar 

  121. Pursell, Z. F., McDonald, J. T., Mathews, C. K. & Kunkel, T. A. Trace amounts of 8-oxo-dGTP in mitochondrial extracts reduce DNA polymerase γ replication fidelity. Nucleic Acids Res. 36, 4990–4995 (2008).

    Google Scholar 

  122. Nakabeppu, Y., Kajitani, K., Sakamoto, K., Yamaguchi, H. & Tsuchimoto, D. MTH1, an oxidized purine nucleoside triphosphatase, prevents the cytotoxicity and neurotoxicity of oxidized purine nucleotides. DNA Repair 5, 761–772 (2006).

    CAS  PubMed  Google Scholar 

  123. Nagy, G. N., Leveles, I. & Vértessy, B. G. Preventive DNA repair by sanitizing the cellular (deoxy)nucleoside triphosphate pool. FEBS J. 281, 4207–4223 (2014).

    CAS  PubMed  Google Scholar 

  124. Gad, H. et al. MTH1 inhibition eradicates cancer by preventing sanitation of the dNTP pool. Nature 508, 215–221 (2014).

    CAS  PubMed  Google Scholar 

  125. Huber, K. V. M. et al. Stereospecific targeting of MTH1 by S-crizotinib as an anticancer strategy. Nature 508, 222–226 (2014). References 124 and 125 present strong evidence that inhibiting dNTP pool sanitation is an effective approach to cancer treatment.

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Heidelberger, C. Chemical carcinogenesis, chemotherapy: cancer's continuing core challenges — G. H. A. Clowes memorial lecture. Cancer Res. 30, 1549–1569 (1970).

    CAS  PubMed  Google Scholar 

  127. Carreras, C. & Santi, D. V. The catalytic mechanism and structure of thymidylate synthase. Annu. Rev. Biochem. 64, 721–762 (1995).

    CAS  PubMed  Google Scholar 

  128. Curtin, N. J., Harris, A. L. & Aherne, G. W. Mechanism of cell death following thymidylate synthase inhibition: 2′-deoxyuridine-5′-triphosphate accumulation, DNA damage, and growth inhibition following exposure to CB3717 and dipyridamole. Cancer Res. 51, 2346–2352 (1991).

    CAS  PubMed  Google Scholar 

  129. Parker, J. B. & Stivers, J. T. Dynamics of uracil and 5-fluorouracil in DNA. Biochemistry 50, 612–617 (2011).

    CAS  PubMed  Google Scholar 

  130. Miyahara, S. et al. Discovery of a novel class of potent human deoxyuridine triphosphatase inhibitors remarkably enhancing the antitumor activity of thymidylate synthase inhibitors. J. Med. Chem. 55, 2970–2980 (2012).

    CAS  PubMed  Google Scholar 

  131. Longley, D. B., Harkin, D. P. & Johnston, P. G. 5-fluorouracil: mechanisms of action and clinical strategies. Nat. Rev. Cancer 3, 330–338 (2003).

    CAS  PubMed  Google Scholar 

  132. Doong, S.-L. & Dolnick, B. J. 5-fluorouracil substitution alters pre-mRNA splicing in vitro. J. Biol. Chem. 263, 4467–4473 (1988).

    CAS  PubMed  Google Scholar 

  133. Yoo, B. K. et al. Identification of genes conferring resistance to 5-fluorouracil. Proc. Natl Acad. Sci. USA 106, 12938–12943 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  134. Chabner, B. A. et al. in Goodman & Gilman's Pharmacological Basis of Therapeutics 12th edn (eds Brunton, L. L. et al.) 1677–1730 (McGraw-Hill, 2011). This chapter in a recent edition of a classic pharmacology textbook describes practical aspects of treating cancer with antagonists of nucleotide metabolism.

    Google Scholar 

  135. Itsko, M. & Schaaper, R. M. dGTP starvation in Escherichia coli provides new insights into the thymineless-death phenomenon. PLOS Genet. 10, e1004310 (2014).

    PubMed  PubMed Central  Google Scholar 

  136. Cannazza, G. et al. Internalization and stability of a thymidylate synthase peptide inhibitor in ovarian cancer cells. J. Med. Chem. 57, 10551–10556 (2014).

    CAS  PubMed  Google Scholar 

  137. Tochowitz, A. et al. Alanine mutants of the interface residues of human thymidylate synthase decode key features of the binding mode of allosteric anticancer peptides. J. Med. Chem. 58, 1012–1018 (2014).

    Google Scholar 

  138. Salo-Ahen, O. M. H. et al. Hotspots in an obligate homodimeric anticancer target. Structural and functional effects of interfacial mutations in human thymidylate synthase. J. Med. Chem. 58, 3572–3581 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Farber, S. et al. Temporary remissions in acute leukemia of children produced by folic acid antagonist, 4-aminopteroyl-glutamic acid (aminopterin). N. Engl. J. Med. 239, 779–787 (1948).

    Google Scholar 

  140. Osborn, M. J., Freeman, M. & Huennekens, F. M. Inhibition of dihydrofolic reductase by aminopterin and amethopterin. Proc. Soc. Exp. Biol. Med. 97, 429–431 (1958).

    CAS  PubMed  Google Scholar 

  141. Reichard, P., Baldesten, A. & Rutberg, L. Formation of deoxycytidine phosphates from cytidine phosphates in extracts from Escherichia coli. J. Biol. Chem. 236, 1150–1155 (1961).

    CAS  PubMed  Google Scholar 

  142. Fersht, A. R. Fidelity of replication of phage ΦX174 by DNA polymerase III holoenzyme: spontaneous mutation by misincorporation. Proc. Natl Acad. Sci. USA 76, 4946–4950 (1979).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The author thanks former students and colleagues who contributed to work cited from his laboratory, particularly L. Wheeler, K. Chimploy, I. Rajagopal, M. L. Tassotto and S. Martomo. He would also like to thank two anonymous referees for their help in improving this article. The author apologizes to those whose work he was unable to cite because of space limitations.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christopher K. Mathews.

Ethics declarations

Competing interests

The author declares no competing financial interests.

PowerPoint slides

Glossary

Thymineless death

A loss of cell viability that occurs when an otherwise well-nourished cell is deprived of thymine nucleotides.

Metastable

Defining a state that does not represent thermodynamic equilibrium but that is nearly stable over the time frame of interest because progress towards equilibrium is slow.

Mutator phenotype

A phenotype characterized by an abnormally high spontaneous mutation rate.

Nucleobases

A term used to define the purine and pyrimidine bases that are found in nucleic acids.

Base excision repair

A DNA repair process that begins with cleavage of the glycosidic bond between a damaged base and the deoxyribose to which it is linked.

Exergonic

Referring to a process or chemical reaction in a non-isolated system that proceeds with a net negative free energy change.

Aicardi–Goutières syndrome

A rare inflammatory disorder, usually of early childhood onset, most typically affecting the brain and the skin; it arises as a result of mutations in one of at least seven human genes, including SAMHD1.

Next-nucleotide effect

An increase in replication errors brought about by correct pairing between a template DNA base and an incoming deoxyribonucleotide that occurs before an upstream insertion error has been corrected by 3′-exonucleolytic proofreading.

Replication stress

DNA damage endogenously generated by errors during DNA replication.

Checkpoint

A point during the cell cycle at which the cell monitors its readiness to proceed into the next phase.

Sliding clamp

A protein associated with replicative DNA polymerases that physically surrounds a DNA strand that is being replicated and prevents the dissociation of the polymerase until replication of the template has been completed.

Free radical scavenger

A compound that can donate one electron to a compound containing a free radical and thereby convert an unpaired electron to an electron pair; a one-electron reductant.

Transversion mutagenesis

A substitution mutation resulting from the change of a purine–pyrimidine base pair to a pyrimidine–purine base pair (for example, AT → CG).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mathews, C. Deoxyribonucleotide metabolism, mutagenesis and cancer. Nat Rev Cancer 15, 528–539 (2015). https://doi.org/10.1038/nrc3981

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrc3981

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