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:

Ferroptosis as a p53-mediated activity during tumour suppression

This article has been updated

Abstract

Although p53-mediated cell-cycle arrest, senescence and apoptosis serve as critical barriers to cancer development, emerging evidence suggests that the metabolic activities of p53 are also important. Here we show that p53 inhibits cystine uptake and sensitizes cells to ferroptosis, a non-apoptotic form of cell death, by repressing expression of SLC7A11, a key component of the cystine/glutamate antiporter. Notably, p533KR, an acetylation-defective mutant that fails to induce cell-cycle arrest, senescence and apoptosis, fully retains the ability to regulate SLC7A11 expression and induce ferroptosis upon reactive oxygen species (ROS)-induced stress. Analysis of mutant mice shows that these non-canonical p53 activities contribute to embryonic development and the lethality associated with loss of Mdm2. Moreover, SLC7A11 is highly expressed in human tumours, and its overexpression inhibits ROS-induced ferroptosis and abrogates p533KR-mediated tumour growth suppression in xenograft models. Our findings uncover a new mode of tumour suppression based on p53 regulation of cystine metabolism, ROS responses and ferroptosis.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Identification of SLC7A11 as a target of p53.
Figure 2: p533KR in regulating SLC7A11 and cystine uptake activity.
Figure 3: Roles of p53 in ferroptosis.
Figure 4: Regulation of p53-mediated ferroptosis by SLC7A11.
Figure 5: p53-mediated metabolic regulation in embryonic development.
Figure 6: p53-mediated ferroptosis in ROS responses.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

Data deposits

Microarray data can be accessed through NCBI Gene Expression Omnibus (GEO) database with accession number GSE57841.

Change history

  • 01 April 2015

    Minor changes were made to Extended Data Figs 2 and 8.

References

  1. Berkers, C. R., Maddocks, O. D., Cheung, E. C., Mor, I. & Vousden, K. H. Metabolic regulation by p53 family members. Cell Metab. 18, 617–633 (2013)

    Article  CAS  Google Scholar 

  2. Jackson, J. G. & Lozano, G. The mutant p53 mouse as a pre-clinical model. Oncogene 32, 4325–4330 (2013)

    Article  CAS  Google Scholar 

  3. Aylon, Y. & Oren, M. New plays in the p53 theater. Curr. Opin. Genet. Dev. 21, 86–92 (2011)

    Article  CAS  Google Scholar 

  4. Junttila, M. R. & Evan, G. I. p53–a Jack of all trades but master of none. Nature Rev. Cancer 9, 821–829 (2009)

    Article  CAS  Google Scholar 

  5. Wang, S. J. & Gu, W. To be, or not to be: functional dilemma of p53 metabolic regulation. Curr. Opin. Oncol. 26, 78–85 (2014)

    Article  CAS  Google Scholar 

  6. Bieging, K. T. & Attardi, L. D. Deconstructing p53 transcriptional networks in tumor suppression. Trends Cell Biol. 22, 97–106 (2012)

    Article  CAS  Google Scholar 

  7. Brady, C. A. et al. Distinct p53 transcriptional programs dictate acute DNA-damage responses and tumor suppression. Cell 145, 571–583 (2011)

    Article  CAS  Google Scholar 

  8. Valente, L. J. et al. p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and proapoptotic effectors p21, Puma, and Noxa. Cell Rep. 3, 1339–1345 (2013)

    Article  CAS  Google Scholar 

  9. Kruse, J. P. & Gu, W. Modes of p53 regulation. Cell 137, 609–622 (2009)

    Article  CAS  Google Scholar 

  10. Li, T. et al. Tumor suppression in the absence of p53-mediated cell-cycle arrest, apoptosis, and senescence. Cell 149, 1269–1283 (2012)

    Article  CAS  Google Scholar 

  11. Lo, M., Wang, Y. Z. & Gout, P. W. The cystine/glutamate antiporter: a potential target for therapy of cancer and other diseases. J. Cell. Physiol. 215, 593–602 (2008)

    Article  CAS  Google Scholar 

  12. Conrad, M. & Sato, H. The oxidative stress-inducible cystine/glutamate antiporter, system : cystine supplier and beyond. Amino Acids 42, 231–246 (2012)

    Article  CAS  Google Scholar 

  13. Sato, H., Tamba, M., Kuriyama-Matsumura, K., Okuno, S. & Bannai, S. Molecular cloning and expression of human xCT, the light chain of amino acid transport system . Antioxid. Redox Signal. 2, 665–671 (2000)

    Article  CAS  Google Scholar 

  14. Vu, B. T. & Vassilev, L. Small-molecule inhibitors of the p53–MDM2 interaction. Curr. Top. Microbiol. Immunol. 348, 151–172 (2011)

    CAS  PubMed  Google Scholar 

  15. Dixon, S. J. et al. Ferroptosis: an iron-dependent form of nonapoptotic cell death. Cell 149, 1060–1072 (2012)

    Article  CAS  Google Scholar 

  16. Huang, Y., Dai, Z., Barbacioru, C. & Sadee, W. Cystine-glutamate transporter SLC7A11 in cancer chemosensitivity and chemoresistance. Cancer Res. 65, 7446–7454 (2005)

    Article  CAS  Google Scholar 

  17. Liu, X. X. et al. MicroRNA-26b is underexpressed in human breast cancer and induces cell apoptosis by targeting SLC7A11. FEBS Lett. 585, 1363–1367 (2011)

    Article  CAS  Google Scholar 

  18. Guo, W. et al. Disruption of xCT inhibits cell growth via the ROS/autophagy pathway in hepatocellular carcinoma. Cancer Lett. 312, 55–61 (2011)

    Article  CAS  Google Scholar 

  19. Montes de Oca Luna, R., Wagner, D. S. & Lozano, G. Rescue of early embryonic lethality in mdm2-deficient mice by deletion of p53. Nature 378, 203–206 (1995)

    Article  ADS  CAS  Google Scholar 

  20. Jones, S. N., Roe, A. E., Donehower, L. A. & Bradley, A. Rescue of embryonic lethality in Mdm2-deficient mice by absence of p53. Nature 378, 206–208 (1995)

    Article  ADS  CAS  Google Scholar 

  21. Gannon, H. S. & Jones, S. N. Using mouse models to explore MDM–p53 signaling in development, cell growth, and tumorigenesis. Genes Cancer 3, 209–218 (2012)

    Article  CAS  Google Scholar 

  22. Mendrysa, S. M. et al. Mdm2 is critical for inhibition of p53 during lymphopoiesis and the response to ionizing irradiation. Mol. Cell. Biol. 23, 462–472 (2003)

    Article  CAS  Google Scholar 

  23. Marine, J. C. & Lozano, G. Mdm2-mediated ubiquitylation: p53 and beyond. Cell Death Differ. 17, 93–102 (2010)

    Article  CAS  Google Scholar 

  24. Chavez-Reyes, A. et al. Switching mechanisms of cell death in mdm2- and mdm4-null mice by deletion of p53 downstream targets. Cancer Res. 63, 8664–8669 (2003)

    CAS  PubMed  Google Scholar 

  25. Yang, W. S. et al. Regulation of ferroptotic cancer cell death by GPX4. Cell 156, 317–331 (2014)

    Article  CAS  Google Scholar 

  26. Hughes, R. H., Silva, V. A., Ahmed, I., Shreiber, D. I. & Morrison, B., III Neuroprotection by genipin against reactive oxygen and reactive nitrogen species-mediated injury in organotypic hippocampal slice cultures. Brain Res. 1543, 308–314 (2014)

    Article  CAS  Google Scholar 

  27. Wang, Z., Jiang, H., Chen, S., Du, F. & Wang, X. The mitochondrial phosphatase PGAM5 functions at the convergence point of multiple necrotic death pathways. Cell 148, 228–243 (2012)

    Article  CAS  Google Scholar 

  28. Lu, M. et al. Restoring p53 function in human melanoma cells by inhibiting MDM2 and cyclin B1/CDK1-phosphorylated nuclear iASPP. Cancer Cell 23, 618–633 (2013)

    Article  ADS  CAS  Google Scholar 

  29. Wade, M. & Wahl, G. M. Targeting Mdm2 and Mdmx in cancer therapy: better living through medicinal chemistry? Mol. Cancer Res. 7, 1–11 (2009)

    Article  CAS  Google Scholar 

  30. Wang, P. Y. et al. Increased oxidative metabolism in the Li-Fraumeni syndrome. N. Engl. J. Med. 368, 1027–1032 (2013)

    Article  CAS  Google Scholar 

  31. Liang, Y., Liu, J. & Feng, Z. The regulation of cellular metabolism by tumor suppressor p53. Cell Biosci. 3, 9 (2013)

    Article  Google Scholar 

  32. Bensaad, K. et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell 126, 107–120 (2006)

    Article  CAS  Google Scholar 

  33. Cairns, R. A., Harris, I. S. & Mak, T. W. Regulation of cancer cell metabolism. Nature Rev. Cancer 11, 85–95 (2011)

    Article  CAS  Google Scholar 

  34. Cheung, E. C., Ludwig, R. L. & Vousden, K. H. Mitochondrial localization of TIGAR under hypoxia stimulates HK2 and lowers ROS and cell death. Proc. Natl Acad. Sci. USA 109, 20491–20496 (2012)

    Article  ADS  CAS  Google Scholar 

  35. Cheung, E. C. et al. TIGAR is required for efficient intestinal regeneration and tumorigenesis. Dev. Cell 25, 463–477 (2013)

    Article  CAS  Google Scholar 

  36. Hu, W. et al. Glutaminase 2, a novel p53 target gene regulating energy metabolism and antioxidant function. Proc. Natl Acad. Sci. USA 107, 7455–7460 (2010)

    Article  ADS  CAS  Google Scholar 

  37. Suzuki, S. et al. Phosphate-activated glutaminase (GLS2), a p53-inducible regulator of glutamine metabolism and reactive oxygen species. Proc. Natl Acad. Sci. USA 107, 7461–7466 (2010)

    Article  ADS  CAS  Google Scholar 

  38. Schmidt, D. et al. ChIP-seq: using high-throughput sequencing to discover protein-DNA interactions. Methods 48, 240–248 (2009)

    Article  CAS  Google Scholar 

  39. Zheng, H. et al. A posttranslational modification cascade involving p38, Tip60, and PRAK mediates oncogene-induced senescence. Mol. Cell 50, 699–710 (2013)

    Article  CAS  Google Scholar 

  40. Kon, N. et al. Inactivation of HAUSP in vivo modulates p53 function. Oncogene 29, 1270–1279 (2010)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Cancer Institute of the National Institutes of Health under awards 5R01CA172023, 5RO1CA166294, 5RO1CA169246, 5RO1CA085533 and 2P01CA080058 to W.G. It was also supported by the National Cancer Institute under award 2P01CA097403 to R.B. and W.G. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health. L.J. and S.-J.W. were supported by NIH cancer biology training grant T32-CA09503. We thank S. Mendrysa for Mdm2 mutant mice.

Author information

Authors and Affiliations

Authors

Contributions

The experiments were conceived and designed by L.J., N. K. and W.G. Experiments were performed mainly by L.J. and N.K. Some of the experiments were performed with help from T.L, S.-J.W., T.S., H.H. and R.B. The paper was written by J.L., N.K., R.B. and W.G.

Corresponding author

Correspondence to Wei Gu.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 SLC7A11 expression is downregulated by p53 and identification of p53 binding sites for mouse Slc7a11 gene.

a, Messenger RNA levels of SLC7A11 in tet-on wild-type p53 stable line and parental H1299 cells treated with doxycycline (0.1 µg ml−1). b, U2OS cells were treated with doxorubicin (0.2 µg ml−1) and mRNA was quantified. c, Osteosarcoma cell lines, U2OS (p53 wild type) and SAOS-2 (p53 null) cells, were treated with doxorubicin (0.2 µg ml−1) and mRNA levels were determined. d, Lung cancer cell lines, H1299 (p53 null) and H460 (p53 wild type) cells, were treated with doxorubicin (0.2 µg ml−1) and RT–PCR was used to determine mRNA expression. e, The breast cancer cell line MCF7 was treated with doxorubicin (0.2 µg ml−1) for indicated duration and RT–qPCR was used to measure mRNA expression. f, RT–qPCR were used to determine the mRNA level of Slc7a11 in MEFs with indicated genotype. g, Schematic diagram representing potential p53 binding locations and sequences on the mouse Slc7a11 gene. TSS, transcription start site; light blue box, 5′-UTR. h, ChIP–qPCR was performed on MEFs that were treated with nutlin (10 µM) for 6 h. All qPCR was performed in two technical replicates and mean ± s.d. are shown. All experiments were repeated independently three times.

Extended Data Figure 2 Characterization of erastin-induced cell death in MEFs.

a, Quantification of cell death as shown in Fig. 3a. Error bars are s.d. from two technical replicates. b, Kinetics of cell death induced by erastin (1 µM) over a 24-h period in MEFs with indicated genotypes. Technical replicates were performed and mean ± s.d. are shown (n = 2). c, Transmission electron microscopy image of wild-type MEFs that were treated with TNFα (20 ng ml−1) and CHX (5 µg ml−1) for 16 h with arrows pointing to fragmented nuclei. d, Wild-type MEFs were treated with mouse TNFα (20 ng ml−1) and CHX (5 µg ml−1) or erastin (1 µM) for 8 h followed by western blots. e, TUNEL assay was carried out using wild-type MEFs treated as in d. f, Quantification of TUNEL signals for e. Mean ± s.d. from ten random microscope views are shown (magnification, ×20). g, MEFs with indicated p53 status were treated with erastin (4 µM) and specific cell death inhibitors for 8 h before images were taken (magnification, ×10). 3-MA, 3-methylademine. All experiments were repeated at least three times and representative data are shown.

Extended Data Figure 3 Effectiveness of cell death inhibitors.

a, Wild-type MEF cells were starved in DMEM medium deprived of glucose, sodium pyruvate or l-glutamine for 2 h with or without 3-methylademine (2 mM) followed by western blots. b, Wild-type MEFs were treated for 48 h with TNFα (20 ng ml−1), SMAC mimetic (100 nM) and Z-VAD-FMK (10 µg ml−1) to induced necroptosis with or without the presence of necrostatin-1 (10 µg ml−1) (magnification, ×10). c, Quantification of cell death as shown in b. PI, propidium iodide. Mean ± s.d. from two technical replicates are shown. d, Wild-type MEFs were treated for 48 h with TNFα (20 ng ml−1), SMAC mimetic (100 nM) and necrostatin-1 (10 µg ml−1) to induce apoptosis with or without the presence of Z-VAD-FMK (10 µg ml−1) (magnification, ×10). e, Quantification of cell death as shown in d. Mean ± s.d. from two technical replicates are shown. f, p533KR/3KR MEFs were treated with erastin (4 µM) and various chemicals that block ferroptosis for 24 h before the percentage of cell death was determined; error bars, s.d. from two technical replicates. DMSO, dimethyl sulfoxide; DFO, deferoxamine; U0126, 1,4-diamino-2,3-dicyano-1,4-bis[2-aminophenylthio] butadiene; β-ME, β-mercaptoethanol; NAC, N-acetyl-l-cysteine. All experiments were independently repeated three times.

Extended Data Figure 4 SLC7A11 is overexpressed in tumours of human cancer patients.

ac, Quantitative RT–PCR was used to determine the expression levels of SLC7A11 in paired normal and cancer tissues from colon (a), kidney (b) and liver (c); average expression levels from normal tissues were normalized to 1 in each type of cancer. Mean ± s.d. from two technical replicates are shown. d, Representative heamotoxylin and eosin (H&E) and immunofluorescence staining of SLC7A11 on frozen sections of paired patient cancer and adjacent normal tissues. Magnifcation, ×20. N, normal tissue; C, cancer tissue. Blue, DAPI; green, anti-ATP1A1; red, anti-SLC7A11. e, DNA sequencing was performed on colon cancer samples and specific mutations were identified. Independent experiments were repeated three times and representative data are shown.

Extended Data Figure 5 Cell death induced by p533KR and erastin is ferroptosis and effect of SLC7A11 overexpression on colony formation.

a, Representative phase-contrast images of cell cultures as treated in Fig. 4b (magnification, ×10). b, p533KR tet-on stable line cells were treated as indicated and the percentage of cell death was quantified (DFO, deferoxamine; U0126, 1,4-diamino-2,3-dicyano-1,4-bis[2-aminophenylthio] butadiene; β-ME, beta-mercaptoethanol; NAC, N-acetyl-l-cysteine). Mean ± s.d. from two technical replicates are shown. c, H1299 cells were transfected with indicated plasmids followed by western blot 24 h later. HA, haemagglutinin d, Representative images of colony formation assay in 10-cm plates as transfected in c. e, Quantification of colony formation assay as shown in d. Numbers of colonies formed in control plates were normalized to 100 and mean ± s.d. from two technical replicates are shown. All experiments were repeated three times with representative data shown.

Extended Data Figure 6 p533KR/3KRMdm2−/− mice are embryonic lethal.

a, Representative gel images for genotyping of Mdm2 status in p533KR/3KR background mice. b, Summary of numbers of live embryos and pups recovered from p533KR/3KRMdm2+/− intercross breeding. c, Haematoxylin and eosin (H&E) and immunohistochemistry staining of p53 and TUNEL assay on E7.5 embryos of indicated genotype (magnification, ×20). d, Percentage of cells with positive TUNEL or BrdU were determined by counting 100 cells in each section from three different embryos. Error bars, s.d.; N.S., not significant. e, Whole-embryo extracts from E9.5 embryos were used for western blot. As positive controls, thymus protein lysate from irradiated (IR) wild-type mouse (8 Gy) was used. f, Messenger RNA expression levels of Puma were determined by RT–qPCR using E9.5 embryos with indicated genotype (n = 3 for p533KR/3KRMdm2+/+ and n = 5 for p533KR/3KRMdm2−/−; error bars, s.d.; N.S, not significant). g, Representative images of whole-mount senescence-associated β-galactosidase staining using E9.5 mouse embryos with indicated genotype (magnification, ×2). h, Same protocol as in g was used to stain control wild-type embryos and embryos of HAUSP heterozygous knockouts, which express β-galactosidase40 (magnification, ×2). i, Late passage senescent wild-type MEFs were stained for senescence-associated β-galactosidase activity using the same protocol as in g (magnification, ×10).

Extended Data Figure 7 Ferrostatin-1 partially rescues p533KR/3KRMdm2−/− mice.

a, Representative morphologies of E11.5 mouse embryos of indicated genotype (magnification, ×3). b, Representative embryos recovered from p533KR/3KRMdm2+/− intercross with or without ferr-1 injection. The dashed line marked by an asterisk highlights the body of a dead p533KR/3KRMdm2−/− embryo, which disintegrated upon further dissection (magnification, ×1.5). c, Head-to-tail lengths of p533KR/3KRMdm2−/− embryos were measured and compared to p533KR/3KRMdm2+/+ controls (n = 4 for each group of p533KR/3KRMdm2−/− embryos with or without ferr-1 treatment; error bars, s.d.). d, Representative haematoxylin and eosin (H&E) staining of eye structures of p533KR/3KRMdm2+/+ and p533KR/3KRMdm2−/− mouse embryos (magnification, ×20 and ×40 as indicated).

Extended Data Figure 8 Synergized ferroptosis by nutlin/ROS.

a, Percentage of cell death as shown in Fig. 6a was quantified. Mean ± s.d. from two technical replicates are shown. b, U2OS cells with stable knockdown of p53 were treated by nultin (10 µM) for 24 h followed by addition of ROS (tert-butyl hydroperoxide, 350 µM) for 4 h. Western blots were performed. c, Quantification of cell death as shown in Fig. 6c. Mean ± s.d. from two technical replicates are shown. d, U2OS cells were treated with nutlin (10 µM) for 24 h first, followed by ROS (tert-butyl hydroperoxide, 350 µM) along with indicated cell death inhibitors; cell death were quantified 24 h later. Error bars, s.d. from two technical replicates. e, U2OS cells were treated with DNA-damaging agents (etoposide, 20 µM; doxorubicin, 0.2 µg ml−1) for 48 h with or without the presence of ferr-1 (2 µM) (magnification, ×10); cell death was quantified in f with mean ± s.d. shown (n = 2 technical replicates). All data were repeated three times independently.

Extended Data Figure 9 Generation of BAC transgenic mice for Slc7a11 overexpression.

a, Schematic diagram showing the procedure for generation of Slc7a11-BAC transgenic mice. b, Snap shot of BACs surrounding mouse Slc7a11 genes. BAC (RP24-242E11) that contains only the Slc7a11 gene was selected for injection. c, PCR at both ends of the BAC construct identified founders (no. 21 and no. 22) as positive BAC transgenic mice. d, Germline transmission was confirmed from both founders identified in c. NC, no template control. e, Thymus and brain tissues from 3-week-old litter mates of control and Slc7a11-BAC transgenic mice were lysed and examined by western blots. f, MEF cells with indicated genotypes were treated as in Fig. 6e for 2 h and mRNA levels were determined by RT–qPCR. Mean ± s.d. from two technical replicates are shown. g, Representative images of cells treated as in Fig. 6e (magnification, ×10).

Extended Data Table 1 p53-regulated genes identified in the wild-type p53 inducible stable line through microarray analysis

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jiang, L., Kon, N., Li, T. et al. Ferroptosis as a p53-mediated activity during tumour suppression. Nature 520, 57–62 (2015). https://doi.org/10.1038/nature14344

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature14344

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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