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:

Methionine is a metabolic dependency of tumor-initiating cells

A Publisher Correction to this article was published on 21 May 2019

This article has been updated

Abstract

Understanding cellular metabolism holds immense potential for developing new classes of therapeutics that target metabolic pathways in cancer. Metabolic pathways are altered in bulk neoplastic cells in comparison to normal tissues. However, carcinoma cells within tumors are heterogeneous, and tumor-initiating cells (TICs) are important therapeutic targets that have remained metabolically uncharacterized. To understand their metabolic alterations, we performed metabolomics and metabolite tracing analyses, which revealed that TICs have highly elevated methionine cycle activity and transmethylation rates that are driven by MAT2A. High methionine cycle activity causes methionine consumption to far outstrip its regeneration, leading to addiction to exogenous methionine. Pharmacological inhibition of the methionine cycle, even transiently, is sufficient to cripple the tumor-initiating capability of these cells. Methionine cycle flux specifically influences the epigenetic state of cancer cells and drives tumor initiation. Methionine cycle enzymes are also enriched in other tumor types, and MAT2A expression impinges upon the sensitivity of certain cancer cells to therapeutic inhibition.

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: Metabolomic characterization of lung tumor-initiating cells and differentiated cells.
Fig. 2: The metabolic requirements of lung tumor-initiating cells.
Fig. 3: Metabolic labeling and tracking of methionine cycle flux.
Fig. 4: Functional and clinical relevance of methionine cycle enzymes in NSCLC.
Fig. 5: Small-molecule inhibition of the methionine cycle disrupts the tumorigenicity of lung tumor-initiating cells.

Similar content being viewed by others

Data availability

The metabolomics datasets generated or analyzed during this study are included in this published article in Supplementary Tables 2 and 3. Additional datasets are also available from the corresponding author upon reasonable request. Source data are available online for Figs. 15 and Extended Data Figs. 1 and 35.

Change history

  • 21 May 2019

    In the version of this article originally published, there is an error in Fig. 5a. Originally, ‘MAT2A’ appeared between ‘Methionine’ and ‘Homocysteine’. ‘MAT2A’ should have been ‘MTR’. The error has been corrected in the PDF and HTML versions of this article.

References

  1. Chaneton, B. et al. Serine is a natural ligand and allosteric activator of pyruvate kinase M2. Nature 491, 458–462 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. DeBerardinis, R. J., Lum, J. J., Hatzivassiliou, G. & Thompson, C. B. The biology of cancer: metabolic reprogramming fuels cell growth and proliferation. Cell Metab. 7, 11–20 (2008).

    CAS  PubMed  Google Scholar 

  3. Jain, M. et al. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science 336, 1040–1044 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Maddocks, O. D. et al. Serine starvation induces stress and p53-dependent metabolic remodelling in cancer cells. Nature 493, 542–546 (2013).

    CAS  PubMed  Google Scholar 

  5. Mayers, J. R. & Vander Heiden, M. G. Famine versus feast: understanding the metabolism of tumors in vivo. Trends Biochem. Sci. 40, 130–140 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Vander Heiden, M. G. et al. Metabolic pathway alterations that support cell proliferation. Cold Spring Harb. Symp. Quant. Biol. 76, 325–334 (2011).

    Google Scholar 

  7. Wise, D. R. & Thompson, C. B. Glutamine addiction: a new therapeutic target in cancer. Trends Biochem. Sci. 35, 427–433 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Anastasiou, D. et al. Inhibition of pyruvate kinase M2 by reactive oxygen species contributes to cellular antioxidant responses. Science 334, 1278–1283 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Kim, D. et al. SHMT2 drives glioma cell survival in ischaemia but imposes a dependence on glycine clearance. Nature 520, 363–367 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Son, J. et al. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature 496, 101–105 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Locasale, J. W. et al. Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to oncogenesis. Nat. Genet. 43, 869–874 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Possemato, R. et al. Functional genomics reveal that the serine synthesis pathway is essential in breast cancer. Nature 476, 346–350 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Kaelin, W. G. Jr & McKnight, S. L. Influence of metabolism on epigenetics and disease. Cell 153, 56–69 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Kaelin, W. G. Jr Cancer and altered metabolism: potential importance of hypoxia-inducible factor and 2-oxoglutarate-dependent dioxygenases. Cold Spring Harb. Symp. Quant. Biol. 76, 335–345 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Lu, C. et al. IDH mutation impairs histone demethylation and results in a block to cell differentiation. Nature 483, 474–478 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Letouze, E. et al. SDH mutations establish a hypermethylator phenotype in paraganglioma. Cancer Cell 23, 739–752 (2013).

    CAS  PubMed  Google Scholar 

  17. Xiao, M. et al. Inhibition of α-KG-dependent histone and DNA demethylases by fumarate and succinate that are accumulated in mutations of FH and SDH tumor suppressors. Genes Dev. 26, 1326–1338 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Kreso, A. & Dick, J. E. Evolution of the cancer stem cell model. Cell Stem Cell 14, 275–291 (2014).

    CAS  PubMed  Google Scholar 

  19. Meacham, C. E. & Morrison, S. J. Tumour heterogeneity and cancer cell plasticity. Nature 501, 328–337 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. O’Brien, C. A., Kreso, A. & Dick, J. E. Cancer stem cells in solid tumors: an overview. Semin. Radiat. Oncol. 19, 71–77 (2009).

    PubMed  Google Scholar 

  21. Oskarsson, T., Batlle, E. & Massague, J. Metastatic stem cells: sources, niches, and vital pathways. Cell Stem Cell 14, 306–321 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Tan, W. L. et al. Novel therapeutic targets on the horizon for lung cancer. Lancet Oncol. 17, e347–e362 (2016).

    CAS  PubMed  Google Scholar 

  23. Lee, Y. A. et al. Identification of tumor initiating cells with a small-molecule fluorescent probe by using vimentin as a biomarker. Angew. Chem. Int. Ed. Engl. 57, 2851–2854 (2018).

    CAS  PubMed  Google Scholar 

  24. Zhang, W. C. et al. Glycine decarboxylase activity drives non-small cell lung cancer tumor-initiating cells and tumorigenesis. Cell 148, 259–272 (2012).

    CAS  PubMed  Google Scholar 

  25. Shiraki, N. et al. Methionine metabolism regulates maintenance and differentiation of human pluripotent stem cells. Cell Metab. 19, 780–794 (2014).

    CAS  PubMed  Google Scholar 

  26. Tsuyama, T., Shiraki, N. & Kume, S. Definitive endoderm differentiation of human embryonic stem cells combined with selective elimination of undifferentiated cells by methionine deprivation. Methods Mol. Biol. 1341, 173–180 (2016).

    CAS  PubMed  Google Scholar 

  27. Shyh-Chang, N. et al. Influence of threonine metabolism on S-adenosylmethionine and histone methylation. Science 339, 222–226 (2013).

    PubMed  Google Scholar 

  28. Maddocks, O. D., Labuschagne, C. F., Adams, P. D. & Vousden, K. H. Serine metabolism supports the methionine cycle and DNA/RNA methylation through de novo ATP synthesis in cancer cells. Mol. Cell 61, 210–221 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Labuschagne, C. F., van den Broek, N. J., Mackay, G. M., Vousden, K. H. & Maddocks, O. D. Serine, but not glycine, supports one-carbon metabolism and proliferation of cancer cells. Cell Rep. 7, 1248–1258 (2014).

    CAS  PubMed  Google Scholar 

  30. Martinez-Chantar, M. L. et al. l-Methionine availability regulates expression of the methionine adenosyltransferase 2A gene in human hepatocarcinoma cells: role of S-adenosylmethionine. J. Biol. Chem. 278, 19885–19890 (2003).

    CAS  PubMed  Google Scholar 

  31. Pendleton, K. E. et al. The U6 snRNA m6A methyltransferase METTL16 regulates SAM synthetase intron retention. Cell 169, 824–835 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Quinlan, C. L. et al. Targeting S-adenosylmethionine biosynthesis with a novel allosteric inhibitor of Mat2A. Nat. Chem. Biol. 13, 785–792 (2017).

    CAS  PubMed  Google Scholar 

  33. Eagle, H. The specific amino acid requirements of a human carcinoma cell (Stain HeLa) in tissue culture. J. Exp. Med. 102, 37–48 (1955).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Eagle, H., Oyama, V. I. & Levy, M. Amino acid requirements of normal and malignant human cells in tissue culture. Arch. Biochem. Biophys. 67, 432–446 (1957).

    CAS  PubMed  Google Scholar 

  35. Carey, B. W., Finley, L. W. S., Cross, J. R., Allis, C. D. & Thompson, C. B. Intracellular α-ketoglutarate maintains the pluripotency of embryonic stem cells. Nature 518, 413–416 (2015).

    CAS  PubMed  Google Scholar 

  36. Guo, H. Y., Herrera, H., Groce, A. & Hoffman, R. M. Expression of the biochemical defect of methionine dependence in fresh patient tumors in primary histoculture. Cancer Res. 53, 2479–2483 (1993).

    CAS  PubMed  Google Scholar 

  37. Yano, S. et al. Selective methioninase-induced trap of cancer cells in S/G2 phase visualized by FUCCI imaging confers chemosensitivity. Oncotarget 5, 8729–8736 (2014).

    PubMed  PubMed Central  Google Scholar 

  38. De La Haba, G. & Cantoni, G. L. The enzymatic synthesis of S-adenosyl-l-homocysteine from adenosine and homocysteine. J. Biol. Chem. 234, 603–608 (1959).

    Google Scholar 

  39. Zhang, W. et al. Fluorinated N,N-dialkylaminostilbenes repress colon cancer by targeting methionine S-adenosyltransferase 2A. ACS Chem. Biol. 8, 796–803 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Jiang, X. et al. Functional characterization of D9, a novel deazaneplanocin A (DZNep) analog, in targeting acute myeloid leukemia (AML). PloS One 10, e0122983 (2015).

    PubMed  PubMed Central  Google Scholar 

  41. Miranda, T. B. et al. DZNep is a global histone methylation inhibitor that reactivates developmental genes not silenced by DNA methylation. Mol. Cancer Ther. 8, 1579–1588 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Tan, J. et al. Pharmacologic disruption of Polycomb-repressive complex 2-mediated gene repression selectively induces apoptosis in cancer cells. Genes Dev. 21, 1050–1063 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Cantor, J. R. & Sabatini, D. M. Cancer cell metabolism: one hallmark, many faces. Cancer Discov. 2, 881–898 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).

    CAS  PubMed  Google Scholar 

  45. Ward, P. S. & Thompson, C. B. Metabolic reprogramming: a cancer hallmark even Warburg did not anticipate. Cancer Cell 21, 297–308 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Halpern, B. C., Clark, B. R., Hardy, D. N., Halpern, R. M. & Smith, R. A. The effect of replacement of methionine by homocystine on survival of malignant and normal adult mammalian cells in culture. Proc. Natl Acad. Sci. USA 71, 1133–1136 (1974).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Hoffman, R. M. & Erbe, R. W. High in vivo rates of methionine biosynthesis in transformed human and malignant rat cells auxotrophic for methionine. Proc. Natl Acad. Sci. USA 73, 1523–1527 (1976).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Stern, P. H., Wallace, C. D. & Hoffman, R. M. Altered methionine metabolism occurs in all members of a set of diverse human tumor cell lines. J. Cell. Physiol. 119, 29–34 (1984).

    CAS  PubMed  Google Scholar 

  49. Batlle, E. & Clevers, H. Cancer stem cells revisited. Nat. Med. 23, 1124–1134 (2017).

    CAS  PubMed  Google Scholar 

  50. Lyssiotis, C. A. & Kimmelman, A. C. Metabolic interactions in the tumor microenvironment. Trends Cell Biol. 27, 863–875 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Mayers, J. R. et al. Tissue of origin dictates branched-chain amino acid metabolism in mutant Kras-driven cancers. Science 353, 1161–1165 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Davidson, S. M. et al. Environment impacts the metabolic dependencies of Ras-driven non-small cell lung cancer. Cell Metab. 23, 517–528 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Guo, H. et al. Therapeutic tumor-specific cell cycle block induced by methionine starvation in vivo. Cancer Res. 53, 5676–5679 (1993).

    CAS  PubMed  Google Scholar 

  54. Ikeda, S., Kawahara-Miki, R., Iwata, H., Sugimoto, M. & Kume, S. Role of methionine adenosyltransferase 2A in bovine preimplantation development and its associated genomic regions. Sci. Rep. 7, 3800 (2017).

    PubMed  PubMed Central  Google Scholar 

  55. Katoh, Y. et al. Methionine adenosyltransferase II serves as a transcriptional corepressor of Maf oncoprotein. Mol. Cell 41, 554–566 (2011).

    CAS  PubMed  Google Scholar 

  56. Kera, Y. et al. Methionine adenosyltransferase II-dependent histone H3K9 methylation at the COX-2 gene locus. J. Biol. Chem. 288, 13592–13601 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Marjon, K. et al. MTAP deletions in cancer create vulnerability to targeting of the MAT2A/PRMT5/RIOK1 axis. Cell Rep. 15, 574–587 (2016).

    CAS  PubMed  Google Scholar 

  58. McDonald, O. G., Wu, H., Timp, W., Doi, A. & Feinberg, A. P. Genome-scale epigenetic reprogramming during epithelial-to-mesenchymal transition. Nat. Struct. Mol. Biol. 18, 867–874 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Feinberg, A. P., Ohlsson, R. & Henikoff, S. The epigenetic progenitor origin of human cancer. Nat. Rev. Genet. 7, 21–33 (2006).

    CAS  PubMed  Google Scholar 

  60. Hansen, K. D. et al. Increased methylation variation in epigenetic domains across cancer types. Nat. Genet. 43, 768–775 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Kahlert, C. et al. Increased expression of ALCAM/CD166 in pancreatic cancer is an independent prognostic marker for poor survival and early tumour relapse. Br. J. Cancer 101, 457–464 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Papal, S., Monti, C. E., Tennison, M. E. & Swaroop, A. Molecular dissection of cone photoreceptor-enriched genes encoding transmembrane and secretory proteins. J. Neurosci. Res. 97, 16–28 (2019).

    CAS  PubMed  Google Scholar 

  63. Wu, M. et al. Multiparameter metabolic analysis reveals a close link between attenuated mitochondrial bioenergetic function and enhanced glycolysis dependency in human tumor cells. Am. J. Physiol Cell Physiol. 292, C125–C136 (2007).

    CAS  PubMed  Google Scholar 

  64. Bligh, E. G. & Dyer, W. J. A rapid method of total lipid extraction and purification. Can. J. Biochem. Physiol. 37, 911–917 (1959).

    CAS  PubMed  Google Scholar 

  65. Smith, C. A., Want, E. J., O’Maille, G., Abagyan, R. & Siuzdak, G. XCMS: processing mass spectrometry data for metabolite profiling using nonlinear peak alignment, matching and identification. Anal. Chem. 78, 779–787 (2006).

    CAS  PubMed  Google Scholar 

  66. Alves, T. C. et al. Integrated, step-wise, mass-isotopomeric flux analysis of the TCA cycle. Cell Metab. 22, 936–947 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Davis, M. A. et al. Calpain drives pyroptotic vimentin cleavage, intermediate filament loss, and cell rupture that mediates immunostimulation. Proc. Natl Acad. Sci. USA 116, 5061–5070 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Zizza, P. et al. TRF2 positively regulates SULF2 expression increasing VEGF-A release and activity in tumor microenvironment. Nucleic Acids Res. https://doi.org/10.1093/nar/gkz041(2019).

  69. Garcia-Lopez, S. et al. Deregulation of the imprinted DLK1-DIO3 locus ncRNAs is associated with replicative senescence of human adipose-derived stem cells. PloS One 13, e0206534 (2018).

    PubMed  PubMed Central  Google Scholar 

  70. Wang, L. et al. H3K36 trimethylation mediated by SETD2 regulates the fate of bone marrow mesenchymal stem cells. PLoS Biol. 16, e2006522 (2018).

    PubMed  PubMed Central  Google Scholar 

  71. Fofou-Caillierez, M. B. et al. Interaction between methionine synthase isoforms and MMACHC: characterization in cblG-variant, cblG and cblC inherited causes of megaloblastic anaemia. Hum. Mol. Genet. 22, 4591–4601 (2013).

    CAS  PubMed  Google Scholar 

  72. Zhu, B. et al. The protective role of DOT1L in UV-induced melanomagenesis. Nat. Commun. 9, 259 (2018).

    PubMed  PubMed Central  Google Scholar 

  73. Limm, K., Dettmer, K., Reinders, J., Oefner, P. J. & Bosserhoff, A. K. Characterization of the methylthioadenosine phosphorylase polymorphism rs7023954—incidence and effects on enzymatic function in malignant melanoma. PloS One 11, e0160348 (2016).

    PubMed  PubMed Central  Google Scholar 

  74. Cooper, S. et al. Jarid2 binds mono-ubiquitylated H2A lysine 119 to mediate crosstalk between Polycomb complexes PRC1 and PRC2. Nat. Commun. 7, 13661 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Chen, X., Zhi, X., Wang, J. & Su, J. RANKL signaling in bone marrow mesenchymal stem cells negatively regulates osteoblastic bone formation. Bone Res. 6, 34 (2018).

    PubMed  PubMed Central  Google Scholar 

  76. Li, Y. et al. Genome-wide analyses reveal a role of Polycomb in promoting hypomethylation of DNA methylation valleys. Genome Biol. 19, 18 (2018).

    PubMed  PubMed Central  Google Scholar 

  77. Fang, D., Gan, H., Wang, H., Zhou, H. & Zhang, Z. Probe the function of histone lysine 36 methylation using histone H3 lysine 36 to methionine mutant transgene in mammalian cells. Cell Cycle 16, 1781–1789 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Zhang, X. et al. H3 ubiquitination by NEDD4 regulates H3 acetylation and tumorigenesis. Nat. Commun. 8, 14799 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Freeman, T. J. et al. Inhibition of pannexin 1 reduces the tumorigenic properties of human melanoma cells. Cancers 11, E102 (2019).

    PubMed  Google Scholar 

  80. Gerhart, S. V. et al. Activation of the p53–MDM4 regulatory axis defines the anti-tumour response to PRMT5 inhibition through its role in regulating cellular splicing. Sci. Rep. 8, 9711 (2018).

    PubMed  PubMed Central  Google Scholar 

  81. Sengupta, S. et al. Genome-wide expression profiling reveals EBV-associated inhibition of MHC class I expression in nasopharyngeal carcinoma. Cancer Res. 66, 7999–8006 (2006).

    CAS  PubMed  Google Scholar 

  82. Haslinger, C. et al. Microarray gene expression profiling of B-cell chronic lymphocytic leukemia subgroups defined by genomic aberrations and VH mutation status. J. Clin. Oncol. 22, 3937–3949 (2004).

    CAS  PubMed  Google Scholar 

  83. Zhang, J. et al. The genetic basis of early T-cell precursor acute lymphoblastic leukaemia. Nature 481, 157–163 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Piccaluga, P. P. et al. Gene expression analysis of peripheral T cell lymphoma, unspecified, reveals distinct profiles and new potential therapeutic targets. J. Clin. Invest. 117, 823–834 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Bonome, T. et al. A gene signature predicting for survival in suboptimally debulked patients with ovarian cancer. Cancer Res. 68, 5478–5486 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Talantov, D. et al. Novel genes associated with malignant melanoma but not benign melanocytic lesions. Clin. Cancer Res. 11, 7234–7242 (2005).

    CAS  PubMed  Google Scholar 

  87. Wallace, T. A. et al. Tumor immunobiological differences in prostate cancer between African-American and European-American men. Cancer Res. 68, 927–936 (2008).

    CAS  PubMed  Google Scholar 

  88. Zhao, H. et al. Different gene expression patterns in invasive lobular and ductal carcinomas of the breast. Mol. Biol. Cell 15, 2523–2536 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank SingHealth Advanced Molecular Pathology Laboratory for their assistance in generating tumor microarray data, M.Y. Lee, K.H.E. Lim, X.H. Yeo, G.S. Tan, M. Nichane, L. Chen, W.A. Zaw, S.L. Khaw, M.S. Noghabi and R. Ettikan for technical assistance, and S.-C. Ng for critical comments. This research is supported by the National Research Foundation, Singapore (NRF-NRFF2015-04), the National Medical Research Council, Singapore (LCG17MAY004; NMRC/OFIRG/0064/2017; NMRC/TCR/007-NCC/2013; OFYIRG16nov013), the Agency for Science, Research and Technology, Singapore (1331AEG071; 334I00053; SPF 2012/001), and the Singapore Ministry of Education under its Research Centers of Excellence initiative. Z. Wang dedicates this manuscript to the memory of Joseph P. Calarco, a wonderful friend and scientist.

Author information

Authors and Affiliations

Authors

Contributions

L.Y.Y., P.K.W.C., C.C.T., K.L.E.P., N.B. and Y.S.H. performed metabolomic studies. Z. Wang, J.H.J.L., H.Y.-K.A., L.S.K.C., H.Y.C., X.J. and Z. Wu performed molecular, cell-based and mouse xenograft experiments. A.T., A.M.H., Q.Y., E.H.T., W.T.L., T.K.H.L., J.Y., S.M. and D.S.W.T. provided key reagents and interpreted the data. Z. Wang, B.L. and W.L.T. designed the research and wrote the manuscript.

Corresponding authors

Correspondence to Bing Lim or Wai Leong Tam.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Extended data

Extended Data Fig. 1 Metabolomic characterization of lung tumor-initiating cells and differentiated cells.

a, Mean tumor mass of implanted cells. Error bars, s.d.; n = 6 tumors. b, Left, mean percentage of CD166+ cells. Error bars, s.d.; n = 4 biologically independent experiments. Right, representative flow cytometry plots, independently repeated four times. Unstained cells are in blue. c, Proliferation curves generated from mean cell viability. Error bars, s.d.; n = 6 biologically independent experiments. d, Mean ECAR. Error bars, s.d.; n = 6 biologically independent experiments. e,f, Top, immunoblots of the indicated cells, independently repeated three times. Bottom, mean ATP levels. Formate was supplemented at 0.5 mM. Error bars, s.d.; n = 6 (e) and n = 5 (f) biologically independent experiments. g, Top, mean tumor mass of implanted cells. Error bars, s.d.; n = 5 tumors. Bottom, mean tumor volume. Error bars, s.e.m.; n = 4 tumors. Mean tumor mass and volume for Control-knockdown (a and Fig. 1c) are included. h, Left, mean percentage of CD166+ cells. Error bars, s.d.; n = 4 biologically independent experiments. Right, representative flow cytometry plots, independently repeated four times. Unstained cells are in blue. i, Mean tumor mass and volume of implanted cells. Error bars, s.d. (left) and s.e.m. (right); n = 4 tumors. Mean tumor mass and volume for TS and Adh cells (a and Fig. 1c) are included. j, Immunoblots of cells grown in the indicated conditions. Independent blots were repeated at least three times with similar results. Histone H3 is used as a loading control. k, Mean abundance of metabolites in cells sorted by CD166 expression from three different tumors. Error bars, s.d. *P < 0.05, one-sided multiple t test corrected for multiple comparisons by the Holm–Sidak method. Exact P values (vs. CD166 cells) are as follows: methionine, 0.0102522; SAM, 0.01934. l, Immunoblots of sorted cells. Independent blots were repeated at least three times with similar results. GAPDH and total histone H3 were used as loading controls.

Source Data

Extended Data Fig. 2 The metabolic requirements of lung tumor-initiating cells.

a, Left, mean number of colonies. Error bars, s.d.; n = 4 biologically independent experiments. Right, mean tumor mass. Error bars, s.d.; n = 9 (no methionine, no serine/glycine), n = 6 (no glutamine) tumors. Mean TS tumor masses (Extended Data Fig. 1a) were included. b, Imaging experiments were independently repeated three times. Top, lesion from TS-implanted lung (l) and normal bronchiole (r). Black bars, 50 μm. Middle, GFP-positive lesion (l) and normal bronchiole (r). Scale bars, 40 μm. Bottom, mean number of GFP-positive lesions. Error bars, s.d.; n = 5 lungs. ****P < 0.0001 by unpaired two-sided Student’s t test. c, Mean tumor mass. Error bars, s.d.; n = 5 tumors. d, Top, mean percentage of CD166+ cells. n = 4 biologically independent experiments. Error bars, s.d. Bottom, representative flow cytometry plots independently repeated four times. Unstained cells are in blue. e, Proliferation curves generated from mean cell viability. Error bars, s.d.; n = 6 biologically independent experiments. f, Left, mean number of colonies. Error bars, s.d.; n = 5 (no methionine + homocysteine), n = 3 (no methionine + SAM), n = 6 (48/48) biologically independent replicates. Right, mean tumor mass. Error bars, s.d.; n = 9 (no methionine + SAM, 48/48), n = 5 (no methionine + homocysteine) tumors. Tumor masses for no methionine (a) and TS cells (Extended Data Fig. 1a) are included. ****P < 0.0001, two-sided Student’s t test with Welch’s correction. P values (vs. no methionine): no methionine + homocysteine, 0.0505; no methionine + SAM, <0.0001; 48/48, <0.0001. g, Mean α-ketoglutarate/succinate ratios. Error bars, s.d.; n = 4 biologically independent experiments. h, Proliferation curves generated from mean cell viability. Error bars, s.d.; n = 6 biologically independent experiments. i, Top, representative flow cytometry plots independently repeated four times. Left, Complete condition. Right, Thymidine treated positive control. Bottom, mean percentage of cells in G2/M. Error bars, s.d.; n = 4 biologically independent experiments. j, Mean percentage of Annexin V+ cells. Error bars, s.d.; n = 4 biologically independent experiments. k, Proliferation curves generated from mean cell viability. Error bars, s.d.; n = 6 biologically independent experiments.

Extended Data Fig. 3 Metabolic labeling and tracking of methionine cycle flux.

a, Changes to [13C]methionine through the methionine cycle. Red circles, 13C; blue triangle, ATP; +, positive charge; black circle, 12C; black letters, enzymes. b, Top, cells were starved of methionine (16 h; l) or not starved (r) before [13C]methionine pulse–chase. Bottom, metabolite species detected are indicated on the right, and proportional abundance (% APE) is indicated on the left. Data represent the mean ± s.e.m.; n = 3 technical replicates. Technical replicates are shown to demonstrate technical consistency. n = 2 (bottom left) and n = 3 (bottom right) biologically independent experiments. c,d, Immunoblots of the indicated cells. β-actin (c) and total histone H3 (d) were used as loading controls. Independent blots were repeated at least three times with similar results. e, Mean tumor mass of implanted cells. Tumor masses from control- and GLDC-knockdown cells (Extended Data Fig. 1a) are included. Error bars, s.d.; n = 6 tumors. f, Immunoblots of the indicated lines. β-actin was used as a loading control. Independent blots were repeated at least three times with similar results. g, Percent APE of metabolite species derived from deuterated homocysteine. Data represent the mean ± s.e.m.; n = 3 technical replicates. Technical replicates are shown to demonstrate technical consistency. n = 3 biologically independent experiments. h, Mean tumor mass. Tumor masses from control- and GLDC-knockdown cells (Extended Data Fig. 1a) are included. Error bars, s.d.; n = 7 tumors. i, Mean ATP levels in the indicated cells supplemented with formate (0.5 mM), methyl-THF (20 μM) or adenosine (200 μM). Error bars, s.d.; n = 6 biologically independent experiments; **P < 0.005, Student’s two-sided t test with Welch’s correction. Exact P values (vs. GLDC shRNA) are as follows: GLDC shRNA + methyl-THF, 0.7947; GLDC shRNA + adenosine, 0.0011; GLDC shRNA + formate, 0.0010.

Source Data

Extended Data Fig. 4 Functional and clinical relevance of methionine cycle enzymes in NSCLC.

a, Immunoblots of the indicated enzymes. β-actin was used as a loading control. Independent blots were repeated at least three times with similar results. b, MTHFR immunohistochemistry (performed once) of a tumor microarray (n = 47) containing paired tumor and normal sections. Top, representative staining intensity. White bar, 20 µm. Bottom, box-and-whisker plots comparing the intensity of tumor and normal sections. Intensity was defined as the product of the maximum immunostaining intensity and the percentage of tumor cells stained per section. Box, twenty-fifth to seventy-fifth percentile; the median value coincides with the seventy-fifth percentile; whiskers indicate the minima and maxima. **P = 0.0005, paired Student’s two-sided t test. t = 3.776, d.f. = 46. c, MTHFR immunohistochemistry (performed once) of an NSCLC tumor microarray (n = 153). Top, representative staining intensity. White bar, 200 µm. Bottom, contingency table correlating staining intensity with NSCLC grade. Chi-squared test P value (P = 0.2297) is indicated at the bottom right. χ2 = 8.116, d.f. = 6. d,e, Immunoblots of MAT2A in the indicated cells or tumors. GAPDH was used as a loading control. Independent blots were repeated at least three times with similar results. f, Proliferation curves generated from mean cell viability of the indicated lines. Error bars, s.d.; n = 10 biologically independent experiments.

Source Data

Extended Data Fig. 5 Small-molecule inhibition of the methionine cycle disrupts the tumorigenicity of lung tumor-initiating cells.

a, Mean tumor mass of implanted cells. Error bars, s.d.; n = 6 (D9, DMSO), n = 9 (FIDAS, FIDAS + SAM) tumors. b, Left, mean percentage of CD166+ cells. Error bars, s.d.; n = 4 biologically independent experiments. Right, representative flow cytometry plots independently repeated four times. Unstained cells are in blue. c,d, Immunoblots of the indicated cells. β-catenin was used as a loading control. Independent blots were repeated at least three times with similar results. e, Proliferation curves generated from mean cell viability. Error bars, s.d.; n = 10 biologically independent experiments. f, Mean percentage of Annexin V+ cells. Error bars, s.d.; n = 4 (DMSO, D9), n = 3 (FIDAS) biologically independent experiments. g,h, Proliferation curves generated from mean cell viability. Error bars, s.d.; n = 10 biologically independent experiments. i, Mean tumor mass of implanted cells. Error bars, s.d.; n = 7 (control), n = 6 (FIDAS) and n = 9 (cisplatin) tumors. j, Mean number of GFP+ lesions. Error bars, s.d.; n = 5 lungs. ****P < 0.0001, two-sided unpaired Student’s t test. k, Individual weight plots of nine mice undergoing the indicated treatment. l, Mean MAT2A mRNA levels in normal versus tumor tissue, including glioblastoma (TCGA), colorectal cancer (TCGA), nasopharyngeal carcinoma81, leukemia82,83, lymphoma84, ovarian carcinoma85, melanoma86, prostate adenocarcinoma87 and breast cancer88. ****P < 0.0001, ***P < 0.001, **P ≤ 0.01, *P ≤ 0.05, Student’s unpaired two-sided t test with Welch’s correction. P values and numbers of normal and tumor samples are as follows: brain: P < 0.0001; n = 10 and n = 547; nasopharynx, P = 0.0005; n = 10 and n = 31; skin: P = 0.0015; n = 7 and n = 63; lymphatic system, P < 0.0001; n = 20 and n = 40; bone marrow (childhood acute lymphatic leukemia): P = 0.0492; n = 8 and n = 566; bone marrow (chronic lymphatic leukemia): P = 0.0003; n = 11 and n = 100; ovary: P < 0.0001; n = 10 and n = 185; breast, P = 0.0215; n = 5 and n = 59; prostate: P = 0.0010; n = 20 and n = 69; colon: P < 0.0001; n = 22 and n = 215. m, Immunoblots of the indicated cells. Independent blots were repeated at least three times with similar results.

Source Data

Supplementary information

Supplementary Information

Supplementary Tables 1–7 and Figures 1 and 2

Reporting Summary

Source data

Source Data Fig. 1

Unprocessed Western Blots

Source Data Fig. 2

Unprocessed Western Blots

Source Data Fig. 3

Unprocessed Western Blots

Source Data Fig. 4

Unprocessed Western Blots

Source Data Fig. 5

Unprocessed Western Blots

Source Data Extended Data Fig. 1

Unprocessed Western Blots

Source Data Extended Data Fig. 3

Unprocessed Western Blots

Source Data Extended Data Fig. 4

Unprocessed Western Blots

Source Data Extended Data Fig. 5

Unprocessed Western Blots

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, Z., Yip, L.Y., Lee, J.H.J. et al. Methionine is a metabolic dependency of tumor-initiating cells. Nat Med 25, 825–837 (2019). https://doi.org/10.1038/s41591-019-0423-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41591-019-0423-5

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer