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:

NUDT15 polymorphisms alter thiopurine metabolism and hematopoietic toxicity

Abstract

Widely used as anticancer and immunosuppressive agents, thiopurines have narrow therapeutic indices owing to frequent toxicities, partly explained by TPMT genetic polymorphisms. Recent studies identified germline NUDT15 variation as another critical determinant of thiopurine intolerance, but the underlying molecular mechanisms and the clinical implications of this pharmacogenetic association remain unknown. In 270 children enrolled in clinical trials for acute lymphoblastic leukemia in Guatemala, Singapore and Japan, we identified four NUDT15 coding variants (p.Arg139Cys, p.Arg139His, p.Val18Ile and p.Val18_Val19insGlyVal) that resulted in 74.4–100% loss of nucleotide diphosphatase activity. Loss-of-function NUDT15 diplotypes were consistently associated with thiopurine intolerance across the three cohorts (P = 0.021, 2.1 × 10−5 and 0.0054, respectively; meta-analysis P = 4.45 × 10−8, allelic effect size = −11.5). Mechanistically, NUDT15 inactivated thiopurine metabolites and decreased thiopurine cytotoxicity in vitro, and patients with defective NUDT15 alleles showed excessive levels of thiopurine active metabolites and toxicity. Taken together, these results indicate that a comprehensive pharmacogenetic model integrating NUDT15 variants may inform personalized thiopurine therapy.

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: NUDT15 genetic variants and their effects on nucleotide diphosphatase activity.
Figure 2: Association of NUDT15 diplotype with mercaptopurine tolerance during ALL therapy in Guatemala, Singapore and Japan.
Figure 3: Effects of NUDT15 on thiopurine metabolism and cytotoxicity.
Figure 4: NUDT15 variants and mercaptopurine metabolism in children during ALL therapy.

Similar content being viewed by others

Accession codes

Accessions

NCBI Reference Sequence

References

  1. Karran, P. & Attard, N. Thiopurines in current medical practice: molecular mechanisms and contributions to therapy-related cancer. Nat. Rev. Cancer 8, 24–36 (2008).

    Article  CAS  PubMed  Google Scholar 

  2. Pui, C.H., Carroll, W.L., Meshinchi, S. & Arceci, R.J. Biology, risk stratification, and therapy of pediatric acute leukemias: an update. J. Clin. Oncol. 29, 551–565 (2011).

    Article  PubMed  Google Scholar 

  3. Vora, A. et al. Treatment reduction for children and young adults with low-risk acute lymphoblastic leukaemia defined by minimal residual disease (UKALL 2003): a randomised controlled trial. Lancet Oncol. 14, 199–209 (2013).

    Article  CAS  PubMed  Google Scholar 

  4. Reinisch, W. et al. Azathioprine versus mesalazine for prevention of postoperative clinical recurrence in patients with Crohn's disease with endoscopic recurrence: efficacy and safety results of a randomised, double-blind, double-dummy, multicentre trial. Gut 59, 752–759 (2010).

    Article  CAS  PubMed  Google Scholar 

  5. Maltzman, J.S. & Koretzky, G.A. Azathioprine: old drug, new actions. J. Clin. Invest. 111, 1122–1124 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Goldberg, R. & Irving, P.M. Toxicity and response to thiopurines in patients with inflammatory bowel disease. Expert Rev. Gastroenterol. Hepatol. 9, 891–900 (2015).

    Article  CAS  PubMed  Google Scholar 

  7. Koren, G. et al. Systemic exposure to mercaptopurine as a prognostic factor in acute lymphocytic leukemia in children. N. Engl. J. Med. 323, 17–21 (1990).

    Article  CAS  PubMed  Google Scholar 

  8. Elion, G.B. The purine path to chemotherapy. Science 244, 41–47 (1989).

    Article  CAS  PubMed  Google Scholar 

  9. Childhood ALL Collaborative Group. Duration and intensity of maintenance chemotherapy in acute lymphoblastic leukaemia: overview of 42 trials involving 12 000 randomised children. Lancet 347, 1783–1788 (1996).

  10. Lennard, L. & Lilleyman, J.S. Variable mercaptopurine metabolism and treatment outcome in childhood lymphoblastic leukemia. J. Clin. Oncol. 7, 1816–1823 (1989).

    Article  CAS  PubMed  Google Scholar 

  11. Lilleyman, J.S. & Lennard, L. Mercaptopurine metabolism and risk of relapse in childhood lymphoblastic leukaemia. Lancet 343, 1188–1190 (1994).

    Article  CAS  PubMed  Google Scholar 

  12. Relling, M.V., Hancock, M.L., Boyett, J.M., Pui, C.H. & Evans, W.E. Prognostic importance of 6-mercaptopurine dose intensity in acute lymphoblastic leukemia. Blood 93, 2817–2823 (1999).

    Article  CAS  PubMed  Google Scholar 

  13. Schmiegelow, K., Nielsen, S.N., Frandsen, T.L. & Nersting, J. Mercaptopurine/Methotrexate maintenance therapy of childhood acute lymphoblastic leukemia: clinical facts and fiction. J. Pediatr. Hematol. Oncol. 36, 503–517 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Teml, A., Schaeffeler, E., Herrlinger, K.R., Klotz, U. & Schwab, M. Thiopurine treatment in inflammatory bowel disease: clinical pharmacology and implication of pharmacogenetically guided dosing. Clin. Pharmacokinet. 46, 187–208 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Candy, S. et al. A controlled double blind study of azathioprine in the management of Crohn's disease. Gut 37, 674–678 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Fraser, A.G., Orchard, T.R. & Jewell, D.P. The efficacy of azathioprine for the treatment of inflammatory bowel disease: a 30 year review. Gut 50, 485–489 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Hanauer, S.B. et al. Postoperative maintenance of Crohn's disease remission with 6-mercaptopurine, mesalamine, or placebo: a 2-year trial. Gastroenterology 127, 723–729 (2004).

    Article  CAS  PubMed  Google Scholar 

  18. Connell, W.R., Kamm, M.A., Ritchie, J.K. & Lennard-Jones, J.E. Bone marrow toxicity caused by azathioprine in inflammatory bowel disease: 27 years of experience. Gut 34, 1081–1085 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. de Boer, N.K., van Bodegraven, A.A., Jharap, B., de Graaf, P. & Mulder, C.J. Drug insight: pharmacology and toxicity of thiopurine therapy in patients with IBD. Nat. Clin. Pract. Gastroenterol. Hepatol. 4, 686–694 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. de Jong, D.J., Goullet, M. & Naber, T.H. Side effects of azathioprine in patients with Crohn's disease. Eur. J. Gastroenterol. Hepatol. 16, 207–212 (2004).

    Article  CAS  PubMed  Google Scholar 

  21. Hindorf, U., Lindqvist, M., Hildebrand, H., Fagerberg, U. & Almer, S. Adverse events leading to modification of therapy in a large cohort of patients with inflammatory bowel disease. Aliment. Pharmacol. Ther. 24, 331–342 (2006).

    Article  CAS  PubMed  Google Scholar 

  22. Posthuma, E.F. et al. Fatal infectious mononucleosis: a severe complication in the treatment of Crohn's disease with azathioprine. Gut 36, 311–313 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Schwab, M. et al. Azathioprine therapy and adverse drug reactions in patients with inflammatory bowel disease: impact of thiopurine S-methyltransferase polymorphism. Pharmacogenetics 12, 429–436 (2002).

    Article  CAS  PubMed  Google Scholar 

  24. Fotoohi, A.K., Coulthard, S.A. & Albertioni, F. Thiopurines: factors influencing toxicity and response. Biochem. Pharmacol. 79, 1211–1220 (2010).

    Article  CAS  PubMed  Google Scholar 

  25. Hedeland, R.L. et al. DNA incorporation of 6-thioguanine nucleotides during maintenance therapy of childhood acute lymphoblastic leukaemia and non-Hodgkin lymphoma. Cancer Chemother. Pharmacol. 66, 485–491 (2010).

    Article  CAS  PubMed  Google Scholar 

  26. Ebbesen, M.S. et al. Incorporation of 6-thioguanine nucleotides into DNA during maintenance therapy of childhood acute lymphoblastic leukemia—the influence of thiopurine methyltransferase genotypes. J. Clin. Pharmacol. 53, 670–674 (2013).

    Article  CAS  PubMed  Google Scholar 

  27. Diouf, B. et al. Somatic deletions of genes regulating MSH2 protein stability cause DNA mismatch repair deficiency and drug resistance in human leukemia cells. Nat. Med. 17, 1298–1303 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Krynetskaia, N.F. et al. Msh2 deficiency attenuates but does not abolish thiopurine hematopoietic toxicity in Msh2−/− mice. Mol. Pharmacol. 64, 456–465 (2003).

    Article  CAS  PubMed  Google Scholar 

  29. Swann, P.F. et al. Role of postreplicative DNA mismatch repair in the cytotoxic action of thioguanine. Science 273, 1109–1111 (1996).

    Article  CAS  PubMed  Google Scholar 

  30. Tidd, D.M. & Paterson, A.R. Distinction between inhibition of purine nucleotide synthesis and the delayed cytotoxic reaction of 6-mercaptopurine. Cancer Res. 34, 733–737 (1974).

    CAS  PubMed  Google Scholar 

  31. Krynetski, E.Y., Krynetskaia, N.F., Bianchi, M.E. & Evans, W.E. A nuclear protein complex containing high mobility group proteins B1 and B2, heat shock cognate protein 70, ERp60, and glyceraldehyde-3-phosphate dehydrogenase is involved in the cytotoxic response to DNA modified by incorporation of anticancer nucleoside analogues. Cancer Res. 63, 100–106 (2003).

    CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Meyer, J.A. et al. Relapse-specific mutations in NT5C2 in childhood acute lymphoblastic leukemia. Nat. Genet. 45, 290–294 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Relling, M.V. et al. Mercaptopurine therapy intolerance and heterozygosity at the thiopurine S-methyltransferase gene locus. J. Natl. Cancer Inst. 91, 2001–2008 (1999).

    Article  CAS  PubMed  Google Scholar 

  35. Lennard, L., Lilleyman, J.S., Van Loon, J. & Weinshilboum, R.M. Genetic variation in response to 6-mercaptopurine for childhood acute lymphoblastic leukaemia. Lancet 336, 225–229 (1990).

    Article  CAS  PubMed  Google Scholar 

  36. Evans, W.E. et al. Preponderance of thiopurine S-methyltransferase deficiency and heterozygosity among patients intolerant to mercaptopurine or azathioprine. J. Clin. Oncol. 19, 2293–2301 (2001).

    Article  CAS  PubMed  Google Scholar 

  37. Gardiner, S.J., Gearry, R.B., Begg, E.J., Zhang, M. & Barclay, M.L. Thiopurine dose in intermediate and normal metabolizers of thiopurine methyltransferase may differ three-fold. Clin. Gastroenterol. Hepatol. 6, 654–660 quiz 604 (2008).

    Article  CAS  PubMed  Google Scholar 

  38. Regueiro, M. & Mardini, H. Determination of thiopurine methyltransferase genotype or phenotype optimizes initial dosing of azathioprine for the treatment of Crohn's disease. J. Clin. Gastroenterol. 35, 240–244 (2002).

    Article  CAS  PubMed  Google Scholar 

  39. Yates, C.R. et al. Molecular diagnosis of thiopurine S-methyltransferase deficiency: genetic basis for azathioprine and mercaptopurine intolerance. Ann. Intern. Med. 126, 608–614 (1997).

    Article  CAS  PubMed  Google Scholar 

  40. Krynetski, E.Y. et al. A single point mutation leading to loss of catalytic activity in human thiopurine S-methyltransferase. Proc. Natl. Acad. Sci. USA 92, 949–953 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Tai, H.L. et al. Thiopurine S-methyltransferase deficiency: two nucleotide transitions define the most prevalent mutant allele associated with loss of catalytic activity in Caucasians. Am. J. Hum. Genet. 58, 694–702 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Relling, M.V., Pui, C.H., Cheng, C. & Evans, W.E. Thiopurine methyltransferase in acute lymphoblastic leukemia. Blood 107, 843–844 (2006).

    Article  CAS  PubMed  Google Scholar 

  43. Relling, M.V., Altman, R.B., Goetz, M.P. & Evans, W.E. Clinical implementation of pharmacogenomics: overcoming genetic exceptionalism. Lancet Oncol. 11, 507–509 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  44. Aricó, M. et al. The seventh international childhood acute lymphoblastic leukemia workshop report: Palermo, Italy, January 29–30, 2005. Leukemia 19, 1145–1152 (2005).

    Article  PubMed  Google Scholar 

  45. Schmiegelow, K. et al. Thiopurine methyltransferase activity is related to the risk of relapse of childhood acute lymphoblastic leukemia: results from the NOPHO ALL-92 study. Leukemia 23, 557–564 (2009).

    Article  CAS  PubMed  Google Scholar 

  46. Yang, S.K. et al. A common missense variant in NUDT15 confers susceptibility to thiopurine-induced leukopenia. Nat. Genet. 46, 1017–1020 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Yang, J.J. et al. Inherited NUDT15 variant is a genetic determinant of mercaptopurine intolerance in children with acute lymphoblastic leukemia. J. Clin. Oncol. 33, 1235–1242 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kham, S.K. et al. Thiopurine S-methyltransferase activity in three major Asian populations: a population-based study in Singapore. Eur. J. Clin. Pharmacol. 64, 373–379 (2008).

    Article  CAS  PubMed  Google Scholar 

  49. Kakuta, Y. et al. NUDT15 R139C causes thiopurine-induced early severe hair loss and leukopenia in Japanese patients with IBD. Pharmacogenomics J. 10.1038/tpj.2015.43 (2015).

  50. Tanaka, Y. et al. Susceptibility to 6-MP toxicity conferred by a NUDT15 variant in Japanese children with acute lymphoblastic leukaemia. Br. J. Haematol. 171, 109–115 (2015).

    Article  CAS  PubMed  Google Scholar 

  51. Relling, M.V. et al. Clinical Pharmacogenetics Implementation Consortium guidelines for thiopurine methyltransferase genotype and thiopurine dosing. Clin. Pharmacol. Ther. 89, 387–391 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Carter, M. et al. Crystal structure, biochemical and cellular activities demonstrate separate functions of MTH1 and MTH2. Nat. Commun. 6, 7871 (2015).

    Article  CAS  PubMed  Google Scholar 

  53. Tai, H.L., Krynetski, E.Y., Schuetz, E.G., Yanishevski, Y. & Evans, W.E. Enhanced proteolysis of thiopurine S-methyltransferase (TPMT) encoded by mutant alleles in humans (TPMT*3A, TPMT*2): mechanisms for the genetic polymorphism of TPMT activity. Proc. Natl. Acad. Sci. USA 94, 6444–6449 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Takagi, Y. et al. Human MTH3 (NUDT18) protein hydrolyzes oxidized forms of guanosine and deoxyguanosine diphosphates: comparison with MTH1 and MTH2. J. Biol. Chem. 287, 21541–21549 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. McLennan, A.G., Cartwright, J.L. & Gasmi, L. The human NUDT family of nucleotide hydrolases. Enzymes of diverse substrate specificity. Adv. Exp. Med. Biol. 486, 115–118 (2000).

    Article  CAS  PubMed  Google Scholar 

  56. 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 

  57. Antillón-Klussmann, F.V.P., Garrido, C., Castellanos, M., De Alarcon, P. & Ribeiro, R. Treatment for acute lymphoblastic leukemia in limited income country: the experience of the Unidad Nacional de Oncología Pediátrica (UNOP) of Guatemala. Pediatr. Blood Cancer 55, 861 (2010).

    Google Scholar 

  58. Yeoh, A.E. et al. Minimal residual disease-guided treatment deintensification for children with acute lymphoblastic leukemia: results from the Malaysia-Singapore acute lymphoblastic leukemia 2003 study. J. Clin. Oncol. 30, 2384–2392 (2012).

    Article  CAS  PubMed  Google Scholar 

  59. Stephens, M., Smith, N.J. & Donnelly, P. A new statistical method for haplotype reconstruction from population data. Am. J. Hum. Genet. 68, 978–989 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Yang, J.J. & Bhojwani, D. Thiopurine S-methyltransferase pharmacogenetics in childhood acute lymphoblastic leukemia. Methods Mol. Biol. 999, 273–284 (2013).

    Article  CAS  PubMed  Google Scholar 

  61. Xu, H. et al. Inherited coding variants at the CDKN2A locus influence susceptibility to acute lymphoblastic leukaemia in children. Nat. Commun. 6, 7553 (2015).

    Article  CAS  PubMed  Google Scholar 

  62. Hofmann, U. et al. Simultaneous quantification of eleven thiopurine nucleotides by liquid chromatography–tandem mass spectrometry. Anal. Chem. 84, 1294–1301 (2012).

    Article  CAS  PubMed  Google Scholar 

  63. Jacobsen, J.H., Schmiegelow, K. & Nersting, J. Liquid chromatography–tandem mass spectrometry quantification of 6-thioguanine in DNA using endogenous guanine as internal standard. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 881–882, 115–118 (2012).

    Article  PubMed  CAS  Google Scholar 

  64. Holleman, A. et al. Gene-expression patterns in drug-resistant acute lymphoblastic leukemia cells and response to treatment. N. Engl. J. Med. 351, 533–542 (2004).

    Article  CAS  PubMed  Google Scholar 

  65. Higgins, J.P., Thompson, S.G. & Spiegelhalter, D.J. A re-evaluation of random-effects meta-analysis. J. R. Stat. Soc. Ser. A Stat. Soc. 172, 137–159 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the patients and parents who participated in the clinical trials included in this study, H. Toyoda at Mie University for his assistance in processing the Japanese Pediatric Leukemia/Lymphoma Study Group samples and C. Smith at St. Jude Children's Research Hospital for querying the 1000 Genomes Project data. This work was supported by the US National Institutes of Health (CA021765 and GM115279), the American Lebanese Syrian Associated Charities of St. Jude Children's Research Hospital, the Order of St. Francis Foundation, the V Foundation for Cancer Research and the Danish Childhood Cancer Foundation. The Japanese Pediatric Leukemia/Lymphoma Study Group ALL-B12 study is supported by the Japanese Ministry of Health, and the MaSpore ALL studies are supported by the National Medical Research Council (Singapore), Children's Cancer Foundation and the Viva Foundation for Children with Cancer. J.J.Y. is an American Society of Hematology Scholar. T.M. is supported by the Mie Prefecture Study-Abroad Scholarship (Mie, Japan). U.H. and M.S. are supported by the Robert Bosch Foundation (Stuttgart, Germany). K.H. is supported by the Pediatric Oncology Education Program grant (CA23944). T.I. is supported by Alex's Lemonade Stand Foundation's pediatric oncology student training (POST) program.

Author information

Authors and Affiliations

Authors

Contributions

Supervised research: J.J.Y. Conceived and designed the experiments: T.M., R.N., H.H., K.S., A.E.J.Y., W.E.E. and J.J.Y. Performed the experiments: T.M., R.N., V.P.-A., X.Z., T.-N.L., K.H., J.N., K. Kihira, U.H., R.M., L.L., C.R.N., T.I., Z.C., E.K.-H.C., C.J., Y.L. and M.S. Performed statistical analysis: V.P.-A. and W.Y. Analyzed the data: T.M., R.N., V.P.-A., W.Y., J.N., U.H., R.M., L.L., T.I., C.J., Y.L., M.S., H.H., K.S. and J.J.Y. Contributed reagents, materials and analysis tools: F.A.K., K. Kihira, Y.K., M.K., K. Koh, C.R.N., S.K.-Y.K., Z.C., E.K.-H.C., D.B., H.I., C.-H.P., M.V.R., A.M., H.H., K.S. and A.E.J.Y. Wrote the manuscript: T.M., W.E.E. and J.J.Y.

Corresponding author

Correspondence to Jun J Yang.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 NUDT15 diplotypes in worldwide populations.

(a) NUDT15 diplotype frequency was based on phased data from the 1000 Genomes Project. (b) Diplotype nomenclatures. Populations include European (CEU, TSI, FIN, GBR and IBS), African (YRI, LWK, GWD, MSL, ESN and ASW), American (MXL, PUR, CLM and PEL), South Asian (GIH, PJL, BEB, STU and ITU) and East Asian (CHB, JPT, CHS, CDX and KHV), as described at http://www.1000genomes.org/category/frequently-asked-questions/population.

Supplementary Figure 2 Enzymatic activity of wild-type and variant NUDT15 with TdGTP as the substrate.

(a) Each variant NUDT15 was expressed in E. coli, and purified protein was subjected to diphosphatase activity measurement with TdGTP as the substrate. (b) Variant or wild-type proteins were combined to determine the level of NUDT15 activity in patients with different diplotypes. Each experiment was performed in triplicate and was repeated at least three times. Center values (dots) represent the means of triplicate experiments; error bars, s.d.

Supplementary Figure 3 Thermostability assay for wild-type and variant NUDT15.

(af) Purified NUDT15 was incubated with SyproOrange, and the mixture was heated from 20 ºC to 95 ºC in increments of 0.2 ºC in the Quant Studio 12K Flex Real-Time PCR system. The temperature midpoint for protein unfolding transition, Tm (indicated by the red arrows), was calculated on the basis of the Boltzmann model. Experiments were performed in triplicate to estimate standard deviation.

Supplementary Figure 4 Association of NUDT15 diplotype with mercaptopurine tolerance during ALL therapy.

(a) Guatemalan cohort. (b) Singaporean cohort. (c) Japanese cohort. Mercaptopurine (MP) dose was adjusted during maintenance therapy to avoid excessive host toxicities (myelosuppression and infections), and the tolerated MP dosage was defined as the stable dose for at least 14 d. There were no significant differences in tolerated MP dosage between NUDT15 diplotypes within the intermediate-activity group (*1/*2, *1/*3, *1/*4 and *1/*5; P = 0.44) or within the low-activity group (*2/*3, *3/*3 and *3/*5; P = 0.41) in the combined cohorts using the Kruskal-Wallis test. Cases with TPMT variants (rs1800462, rs1800460 and rs1142345) were excluded from the analysis. Each box includes data between the 25th and 75th percentiles, with the horizontal line indicating the median.

Supplementary Figure 5 Effects of NUDT15 and TPMT genotypes on mercaptopurine tolerance during ALL therapy.

(a) Guatemalan cohort. (b) Singaporean cohort. Mercaptopurine (MP) dose was adjusted during maintenance therapy to avoid host toxicities (myelosuppression and infections), and the tolerated MP dosage was defined as the stable dose for at least 14 d. TPMT diplotypes were based on rs1800462, rs1800460 and rs1145345. No TPMT variants were observed in the Japanese cohort. Each box includes data between the 25th and 75th percentiles, with the horizontal line indicating the median. P value was estimated for the association of NUDT15 diplotype (as normal-, intermediate- and low-activity groups) with the tolerated MP dosage by using linear regression model with (*) or without (**) adjusting for TPMT genotype (as wild type or heterozygous).

Supplementary Figure 6 NUDT15 knockdown in a lymphoid cell line.

Nalm6 cells were transfected with lentiviral particles of shRNA specific to human NUDT15 (NUDT15 KD) or scrambled sequence (control), and stable clones were established by puromycin selection. (a,b) NUDT15 expression was determined in knockdown and control cells by RT-PCR (a) and immunoblotting (b). Experiments were performed in triplicate and were repeated at least three times.

Supplementary Figure 7 Effects of NUDT15 expression on mercaptopurine and thioguanine cytotoxicity.

(a,b) NUDT15-knockdown (NUDT15 KD; red) cells were established by lentiviral transduction of NUDT15-specific shRNA, and control cells (black) were transduced with non-targeted vectors. Cytotoxicity was determined by MTT assay following incubation for 72 h with increasing concentrations of MP (a) and TG (b). Mean values are plotted in each panel; error bars, s.d. from triplicate experiments.

Supplementary Figure 8 Effects of NUDT15 expression on cytotoxicity and metabolism of azathioprine.

(a,b) NUDT15-knockdown (NUDT15 KD; red) Nalm6 cells were established by lentiviral transduction of NUDT15-specific shRNA, and control cells (black) were transduced with non-targeted vectors. Azathioprine cytotoxicity was determined by MTT assay following drug exposure for 72 h (a). DNA-TG level was quantified after exposure for 48 h to azathioprine (b). Mean values are plotted in each panel; error bars, s.d. from triplicate experiments for cytotoxicity and duplicate experiments for DNA-TG level.

Supplementary Figure 9 NUDT15 variants and mercaptopurine metabolism in children during ALL therapy.

(a,b) DNA-TG levels were analyzed in the Singaporean (a) and Japanese (b) cohorts. Sixty-three and 44 samples were successfully measured from 32 cases with wild-type TPMT in the Singaporean and Japanese cohorts, respectively. An average DNA-TG level was estimated for each patient and then normalized on the basis of the tolerated MP dosage. The ratio of DNA-TG/MP dosage was plotted against the NUDT15 diplotypes. There were no significant differences in normalized DNA-TG between diplotypes within the intermediate-activity group (*1/*2, *1/*3 and *1/*5; P = 0.29) or within the low-activity group (*3/*5, *3/*3 and *2/*3; P = 0.37) in the combined cohorts using the Kruskal-Wallis test.

Supplementary Figure 10 NUDT15 genotype was associated with the thioguanine sensitivity of primary ALL blasts in vitro.

In 285 children with newly diagnosed ALL treated at St. Jude Children’s Research Hospital, primary leukemia cells at diagnosis were evaluated for sensitivity to TG using MTT assay as previously described (N. Engl. J. Med. 351, 533, 2004). NUDT15 genotype was summarized as diplotype (Online Methods), and its association with TG sensitivity (with LC50 as a continuous variable) was determined by a linear regression model (Online Methods).

Supplementary Figure 11 NUDT15 genotype and its protein stability in vitro.

Variant or wild-type NUDT15 cDNA encoding an N-terminal FLAG tag was transiently expressed in HEK293T cells using polyethylenimine reagent. Cycloheximide (CHX; 50 mg/ml) was added 48 h after transfection, and NUDT15 protein level was monitored after 0, 24 and 48 h by immunoblotting with β-actin as the loading control. Immunoblot images are shown (top), and protein level was quantified using ImageJ software (bottom).

Supplementary Figure 12 Expression and purification of wild-type and variant human NUDT15 proteins.

N-terminally His-tagged human NUDT15 was expressed in E. coli BL21 with IPTG induction and purified by affinity chromatography. One microgram of protein was loaded onto a 4–15% SDS-PAGE gel for electrophoresis, and the gel was stained with Coomassie Brilliant Blue.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–12 and Supplementary Tables 1–4. (PDF 1963 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Moriyama, T., Nishii, R., Perez-Andreu, V. et al. NUDT15 polymorphisms alter thiopurine metabolism and hematopoietic toxicity. Nat Genet 48, 367–373 (2016). https://doi.org/10.1038/ng.3508

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3508

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research