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:

DNAJA1 controls the fate of misfolded mutant p53 through the mevalonate pathway

Abstract

Stabilization of mutant p53 (mutp53) in tumours greatly contributes to malignant progression. However, little is known about the underlying mechanisms and therapeutic approaches to destabilize mutp53. Here, through high-throughput screening we identify statins, cholesterol-lowering drugs, as degradation inducers for conformational or misfolded p53 mutants with minimal effects on wild-type p53 (wtp53) and DNA contact mutants. Statins preferentially suppress mutp53-expressing cancer cell growth. Specific reduction of mevalonate-5-phosphate by statins or mevalonate kinase knockdown induces CHIP ubiquitin ligase-mediated nuclear export, ubiquitylation, and degradation of mutp53 by impairing interaction of mutp53 with DNAJA1, a Hsp40 family member. Knockdown of DNAJA1 also induces CHIP-mediated mutp53 degradation, while its overexpression antagonizes statin-induced mutp53 degradation. Our study reveals that DNAJA1 controls the fate of misfolded mutp53, provides insights into potential strategies to deplete mutp53 through the mevalonate pathway–DNAJA1 axis, and highlights the significance of p53 status in impacting statins’ efficacy on cancer 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: Statins reduce levels of conformational or misfolded mutp53.
Figure 2: Statins induce mutp53 degradation by CHIP.
Figure 3: Reduced MVP mirrors statins’ effects on mutp53.
Figure 4: Statins favourably impede mutp53-tumour growth.
Figure 5: Statins reduce growth of p53R172H MEFs.
Figure 6: DNAJA1 knockdown induces mutp53 degradation.
Figure 7: DNAJA1 nullifies statins’ effects on mutp53.

Similar content being viewed by others

References

  1. Lane, D. & Levine, A. p53 research: the past thirty years and the next thirty years. Cold Spring Harb. Perspect. Biol. 2, a000893 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Levav-Cohen, Y. et al. The p53-Mdm2 loop: a critical juncture of stress response. Subcell. Biochem. 85, 161–186 (2014).

    Article  Google Scholar 

  3. Rivlin, N., Koifman, G. & Rotter, V. p53 orchestrates between normal differentiation and cancer. Semin. Cancer Biol. 32, 10–17 (2015).

    Article  CAS  Google Scholar 

  4. Boeckler, F. M. et al. Targeted rescue of a destabilized mutant of p53 by an in silico screened drug. Proc. Natl Acad. Sci. USA 105, 10360–10365 (2008).

    Article  CAS  Google Scholar 

  5. Terzian, T. et al. The inherent instability of mutant p53 is alleviated by Mdm2 or p16INK4a loss. Genes Dev. 22, 1337–1344 (2008).

    Article  CAS  Google Scholar 

  6. Lukashchuk, N. & Vousden, K. H. Ubiquitination and degradation of mutant p53. Mol. Cell. Biol. 27, 8284–8295 (2007).

    Article  CAS  Google Scholar 

  7. Alexandrova, E. M. et al. Improving survival by exploiting tumour dependence on stabilized mutant p53 for treatment. Nature 523, 352–356 (2015).

    Article  CAS  Google Scholar 

  8. Masciarelli, S. et al. Gain-of-function mutant p53 downregulates miR-223 contributing to chemoresistance of cultured tumor cells. Oncogene 33, 1601–1608 (2014).

    Article  CAS  Google Scholar 

  9. Rivlin, N., Brosh, R., Oren, M. & Rotter, V. Mutations in the p53 tumor suppressor gene: important milestones at the various steps of tumorigenesis. Genes Cancer 2, 466–474 (2011).

    Article  CAS  Google Scholar 

  10. Zhong, S. et al. Statin use and mortality in cancer patients: Systematic review and meta-analysis of observational studies. Cancer Treat. Rev. 41, 554–567 (2015).

    Article  CAS  Google Scholar 

  11. Li, M. et al. Mono- versus polyubiquitination: differential control of p53 fate by Mdm2. Science 302, 1972–1975 (2003).

    Article  CAS  Google Scholar 

  12. Altwairgi, A. K. Statins are potential anticancerous agents (review). Oncol. Rep. 33, 1019–1039 (2015).

    Article  CAS  Google Scholar 

  13. Shimoyama, S. Statins are logical candidates for overcoming limitations of targeting therapies on malignancy: their potential application to gastrointestinal cancers. Cancer Chemother. Pharmacol. 67, 729–739 (2011).

    Article  CAS  Google Scholar 

  14. Freed-Pastor, W. A. et al. Mutant p53 disrupts mammary tissue architecture via the mevalonate pathway. Cell 148, 244–258 (2012).

    Article  CAS  Google Scholar 

  15. Varley, J. M. et al. A detailed study of loss of heterozygosity on chromosome 17 in tumours from Li-Fraumeni patients carrying a mutation to the TP53 gene. Oncogene 14, 865–871 (1997).

    Article  CAS  Google Scholar 

  16. Venkatachalam, S. et al. Retention of wild-type p53 in tumors from p53 heterozygous mice: reduction of p53 dosage can promote cancer formation. EMBO J. 17, 4657–4667 (1998).

    Article  CAS  Google Scholar 

  17. McDonough, H. & Patterson, C. CHIP: a link between the chaperone and proteasome systems. Cell Stress Chaperones 8, 303–308 (2003).

    Article  CAS  Google Scholar 

  18. Edkins, A. L. CHIP: a co-chaperone for degradation by the proteasome. Subcell. Biochem. 78, 219–242 (2015).

    Article  CAS  Google Scholar 

  19. Peng, Y., Chen, L., Li, C., Lu, W. & Chen, J. Inhibition of MDM2 by hsp90 contributes to mutant p53 stabilization. J. Biol. Chem. 276, 40583–40590 (2001).

    Article  CAS  Google Scholar 

  20. Li, D. et al. Functional inactivation of endogenous MDM2 and CHIP by HSP90 causes aberrant stabilization of mutant p53 in human cancer cells. Mol. Cancer Res. 9, 577–588 (2011).

    Article  CAS  Google Scholar 

  21. Muller, P., Hrstka, R., Coomber, D., Lane, D. P. & Vojtesek, B. Chaperone-dependent stabilization and degradation of p53 mutants. Oncogene 27, 3371–3383 (2008).

    Article  CAS  Google Scholar 

  22. Zhou, P. et al. ErbB2 degradation mediated by the co-chaperone protein CHIP. J. Biol. Chem. 278, 13829–13837 (2003).

    Article  CAS  Google Scholar 

  23. Murata, S., Minami, Y., Minami, M., Chiba, T. & Tanaka, K. CHIP is a chaperone-dependent E3 ligase that ubiquitylates unfolded protein. EMBO Rep. 2, 1133–1138 (2001).

    Article  CAS  Google Scholar 

  24. Hiraki, M. et al. Small-molecule reactivation of mutant p53 to wild-type-like p53 through the p53-Hsp40 regulatory axis. Chem. Biol. 22, 1206–1216 (2015).

    Article  CAS  Google Scholar 

  25. King, F. W., Wawrzynow, A., Hohfeld, J. & Zylicz, M. Co-chaperones Bag-1, Hop and Hsp40 regulate Hsc70 and Hsp90 interactions with wild-type or mutant p53. EMBO J. 20, 6297–6305 (2001).

    Article  CAS  Google Scholar 

  26. Yue, X. et al. BAG2 promotes tumorigenesis through enhancing mutant p53 protein levels and function. Elife 4, e08401 (2015).

    Article  Google Scholar 

  27. Fan, C. Y., Lee, S. & Cyr, D. M. Mechanisms for regulation of Hsp70 function by Hsp40. Cell Stress Chaperones 8, 309–316 (2003).

    Article  CAS  Google Scholar 

  28. Gandelman, K., Malhotra, B., LaBadie, B. B., Crownover, P. & Bergstrom, T. Analytes of interest and choice of dose: two important considerations in the design of bioequivalence studies with atorvastatin. Bioequiv. Bioavailab. 3, 62–68 (2011).

    CAS  Google Scholar 

  29. Bisgaier, C. L. et al. Attenuation of plasma low density lipoprotein cholesterol by select 3-hydroxy-3-methylglutaryl coenzyme A reductase inhibitors in mice devoid of low density lipoprotein receptors. J. Lipid Res. 38, 2502–2515 (1997).

    CAS  PubMed  Google Scholar 

  30. Moon, H., Hill, M. M., Roberts, M. J., Gardiner, R. A. & Brown, A. J. Statins: protectors or pretenders in prostate cancer? Trends Endocrinol. Metab. 25, 188–196 (2014).

    Article  CAS  Google Scholar 

  31. Baandrup, L., Dehlendorff, C., Friis, S., Olsen, J. H. & Kjaer, S. K. Statin use and risk for ovarian cancer: a Danish nationwide case-control study. Br. J. Cancer 112, 157–161 (2015).

    Article  CAS  Google Scholar 

  32. Zhang, X. L. et al. Statin use and risk of kidney cancer: a meta-analysis of observational studies and randomized trials. B. J. Clin. Pharmacol. 77, 458–465 (2014).

    Article  CAS  Google Scholar 

  33. Qiu, X. B., Shao, Y. M., Miao, S. & Wang, L. The diversity of the DnaJ/Hsp40 family, the crucial partners for Hsp70 chaperones. Cell Mol. Life Sci. 63, 2560–2570 (2006).

    Article  CAS  Google Scholar 

  34. Kota, P., Summers, D. W., Ren, H. Y., Cyr, D. M. & Dokholyan, N. V. Identification of a consensus motif in substrates bound by a Type I Hsp40. Proc. Natl Acad. Sci. USA 106, 11073–11078 (2009).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank B. Vogelstein (Johns Hopkins Medicine, USA), D. R. Welch (University of Kansas Medical Center, USA), S. Anant (University of Kansas Medical Center, USA), D. A. Dixon (University of Kansas Medical Center, USA), J. Lewis-Wambi (University of Kansas Medical Center, USA), S. Thomas (University of Kansas Medical Center, USA), T. Komiya (University of Kansas Medical Center, USA) and G. Lozano (MD Anderson Cancer Center, USA) for providing cell lines. We also thank A. K. Godwin, S. Hyter, R. Perez, N. K. Sharma, R. Pradhan, M. Danley, T. Izumi and R. Stein for technical assistance and helpful discussion. Research reported in this publication was supported by NIH R01-CA174735-01A1 (T.I.), P30-GM103495 (T.I.) and P30-CA168524-02 (T.I.) grants, and utilized the Lead Development and Optimization Shared Resource.

Author information

Authors and Affiliations

Authors

Contributions

T.I. supervised the project. A.P., A.Ranjan, S.V.I. and A.Roy performed the experiments. A.Roy performed high-throughput analyses. T.I. and A.P. wrote the manuscript. A.Ranjan, S.V.I., S.P., S.J.W. and A.Roy commented on experiments and edited the manuscript.

Corresponding author

Correspondence to Tomoo Iwakuma.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Screening strategy to identify statins as compounds that deplete p53R175H.

(a) Saos2 (p53null) cells were infected with a lentiviral vector encoding a chimeric fusion protein of p53R175H and luciferase reporter (p53R175H-Luc). Compounds that induce p53R175H degradation are also expected to degrade the p53R175H-Luc fusion protein, thereby reducing luciferase activities. The p53R175H-Luc cells were exposed to chemical libraries containing 9,000 compounds (2.5 μM, bioactives and FDA approved drugs) and DMSO (control) for 24 h (1st screening). Forty four (44) compounds showing reduced luciferase activity compared with DMSO were examined for the dose dependency in the luciferase activity using p53R175H-Luc cells, as well as for a counter screening using Saos2 cells expressing only luciferase reporter (Luc, 2nd screening). Top 10 compounds that preferentially reduced luciferase activity in p53R175H-Luc cells with dose dependency were selected. (b) Representative results of WB for p53 and vinculin using MG63 (p53null) exogenously expressing p53R175H or SK-Br-3 (p53R175H) and U2OS (p53wt) treated with 7 compounds other than statins for 24 h. (c) WB for p53 and vinculin using SK-Br-3 cells treated with DMSO (D) or lovastatin (L; 4 μM) for indicated time points (h: h). NT, not treated. (d) WB for p53 and vinculin using SJSA-1 (p53wt) cells infected with retroviral vector encoding control empty (control) or p53R175H cDNA (+ p53R175H) and treated with D or L (4 μM) for 24 h. (e) WB for p53 and vinculin using H-2087 (p53V 157F) cells treated with D or L for 24 h.

Supplementary Figure 2 Statins induce ubiquitination and degradation of mutp53 via an E3 ubiquitin ligase CHIP, but not MDM2.

(a) Quantitative RT-PCR for human p53 and GAPDH using total RNAs from SK-Br-3 and U2OS cells treated with DMSO (D) or lovastatin (L) at 4 μM for 24 h. Data are shown as relative p53 mRNA expression to that of GAPDH. Error bars, means ± s.d. (n = 3 independent experiments), NS: not significant; Student’s t test (two-tailed). (b) WB for p53 and vinculin using KHOS/NP (p53R156P) cells treated with cyclohexamide (CHX, 50 nM) for indicated time period following pre-incubation with D or 4 μM of L for 12 h (top). Graph showing relative p53R175H levels with time compared to those without CHX treatment (bottom). (c) Ubiquitination assays for p53 using KHOS/NP (left) or CAL33 (p53R175H, right) cells treated with D or 4 μM of L for 22 h. High molecular bands indicate ubiquitinated mutp53. (d) WB for indicated proteins using KHOS/NP cells infected with non-silencing control or MDM2 (M1, 2, 3) shRNA-encoding lentiviral vectors and treated with L at 4 μM for 24 h. (e) WB for indicated proteins using KHOS/NP cells infected with non-silencing control or CHIP (CH1, 2) shRNA-encoding lentiviral vectors and treated with L at 4 μM for 24 h. (f) WB for p53 and vinculin using SK-Br-3 (p53R175H) cells infected with non-silencing control or CHIP shRNA-encoding lentiviral vectors and treated with cyclohexamide (CHX, 50 nM) for indicated time period following pre-incubation with D or 4 μM of L for 12 h (left). Graph showing relative p53R175H levels with time, compared to those without CHX treatment (right). Error bars, means ± s.d. (n = 3 independent experiments). Statistics source data for a and f are provided in Supplementary Table 1.

Supplementary Figure 3 Both MVP and MVA-5PP rescue cholesterol production that is inhibited by statins, whereas only MVP, but not MVA-5PP, nullifies statin-induced mutp53 degradation.

(a) WB for p53 and vinculin using KHOS/NP cells treated with DMSO (D) or lovastatin (L; 4 μM) with or without supplementation with MVA-5PP (100 and 200 μM). (b) WB for p53 and vinculin using SK-Br-3 (p53R175H) cells treated with cyclohexamide (CHX, 50 nM) for indicated time period, following pre-incubation with D, L (4 μM), L + MVP (200 μM), or L + MVA-5PP (200 μM) for 12 h (left). In Supplementary Fig. 2f and 3b, the same samples for DMSO and lovastatin treatments were used till 8 h-time points, and hence, the same data are used in both figures. Graph showing relative p53R175H levels with time compared to those without CHX treatment (right). Error bars, means ± s.d. (n = 3 independent experiments). (c) Measurement of intracellular cholesterol. SK-Br-3 (left) and KHOS/NP (right) cells were treated with L, along with MVP (200 μM) or MVA-5PP (200 μM) for 24 h, and the levels of cholesterol were determined using the Total Cholesterol Assay Kit (Colorimetric; Cell Biolabs, San Diego, CA). Graphs showing cholesterol values relative to that in D-treated cells. Error bars, means ± s.d. (n = 3 independent experiments), P < 0.05, P < 0.01; Student’s t test (two-tailed). (d) Measurement of intracellular cholesterol levels in CAL33 cells infected with non-silencing control or MVK shRNA-encoding lentiviral vectors and treated with vehicle (water), MVA (200 μM), or MVP (200 μM). Error bars, means ± s.d. (n = 3 independent experiments), P < 0.05, P < 0.01, NS: not significant; Student’s t test (two-tailed). Statistics source data for b, c, and d are provided in Supplementary Table 1.

Supplementary Figure 4 Effects of statins on oncogenic properties of human cancer cells.

(a) MTT assays using SK-Br-3, HT29 cells, as well as HCT116 sub-cell lines (p53wt, p53null, p53null+R175H, p53null+R273H) treated with different concentrations of DMSO or lovastatin. Error bars, means ± s.d. (n = 3 independent experiments), P < 0.05, P < 0.01, NS: not significant; Student’s t test (two-tailed). Results of WB for p53 in HCT116 p53wt, p53null+R175H, and p53null+R273H cells treated with DMSO or lovastatin are also shown on the right. (b) Colony formation assays using SK-Br-3, HT29, and HCT116 sub-cell lines seeded onto 6-well plates (500 cells) and treated with DMSO or lovastatin (4 μM) every other day for 10 days. Formed colonies were stained with crystal violet and counted. Representative images (left) and a summarized graph (right). Error bars, means ± s.d. (n = 3 independent experiments), P < 0.01, NS: not significant; Student’s t test (two-tailed). (c) Representative images of PI-staining and flow cytometry using U2OS and KHOS/NP cells treated with MDSO (D) or lovastatin (L) for 72 h (left). Summarized graph showing percentages (%) of sub-G0/G1 fraction and G1/S ratio (right). Error bars, means ± s.d. (n = 3 independent experiments), P < 0.05; Student’s t test (two-tailed). (d) Tumor formation assays in mice subcutaneously injected with CAL33 cells (1 × 106). When subcutaneous tumors reached 3 mm in diameter, mice were intraperitoneally injected with DMSO (D) or atorvastatin (A, 30 mg kg−1) daily, and tumors were measured every 2–3 day until day 21. Representative images of tumors are shown in insets. Error bars, means ± s.d. (n = 5 animals for each group), P < 0.05; Student’s t test (two-tailed). Scale bar, 1 cm. (e) Body weights of mice bearing KHOS/NP-tumors treated with DMSO or atorvastatin (30 mg kg−1, daily). Error bars, means ± s.d. (n = 5 animals for each group). (f) Representative images of IHC for p53, Ki67, and cleaved caspase 3 using CAL33 tumors treated with DMSO or atorvastatin. Scale bar, 50 μm. Statistics source data for ac are provided in Supplementary Table 1.

Supplementary Figure 5 Response of MEFs with different p53 genotypes to statins.

(a) WB for p53 and vinculin using E1A/H-RasG12V-transformed p53+/+, p53+/R172H, and p53R172H/R172H MEFs treated with DMSO, lovastatin (L, 4 μM), or atorvastatin (A, 4 μM) for 24 h. (b) Quantitative RT-PCR for p21 and PUMA using mRNA from E1A/H-RasG12V-transformed MEFs treated with (+) or without (−) 0.5 μM of doxorubicin for 24 h. Data are shown as relative mRNA expression to that of GAPDH. Error bars, means ± s.d. (n = 3 independent experiments), P < 0.01, P < 0.05, NS: not significant; Student’s t test (two-tailed). (c) Colony formation assays using transformed MEFs (500) treated with D or L (4 μM) for 10 days. Representative images (left) and a summarized graph (right). Error bars, means ± s.d. (n = 3 independent experiments), P < 0.01; Student’s t test (two-tailed). (d) Representative images of IHC for p53, Ki67, and cleaved caspase 3 using transformed MEFs-derived tumors treated with D or A (magnification, ×40). Scale bar, 50 μm. Statistics source data for b and c are provided in Supplementary Table 1.

Supplementary Figure 6 Hsp40 (DNAJA1), but not Hsp90, is involved in statin-induced mutp53 degradation, and reduced MVP by statins changes the binding partner of mutp53 from DNAJA1 to CHIP.

(a) WB for Hsp90, Hsp90 client proteins (p53R175H, EGFR, Raf-1, and ErbB2), and β-actin or GAPDH using SK-Br-3 cells treated with DMSO (D) or lovastatin (4 μM). (b) Immunofluorescence for p53R175H using SK-Br-3 cells infected with lentiviral vectors encoding control or DNAJA1 shRNAs and treated with vehicle (ethanol) or leptomycin B (LMB, 50 nM). Scale bar, 50 μm.

Supplementary Figure 7 Reduced MVP inhibits mutp53’s binding to DNAJA1 but increases its interaction with CHIP.

(a) Co-immunoprecipitation and WB for p53R175H and DNAJA1 using CAL33 cells treated with D or L for 11 h, before the degradation of mutp53. (b) Co-immunoprecipitation and WB for p53Y 220C and DNAJA1 using BxPC-3 cells treated with D, L, or L + MVP for 11 h. (c) Co-immunoprecipitation and WB for p53Y 220C and CHIP using BxPC-3 cells treated with D or L for 11 h. (d) A schematic view of a proposed mechanism for statin-induced mutp53 degradation. In the absence of DNAJA1, conformational/misfolded mutp53 is degraded by CHIP, suggesting that DNAJA1 determines the fate of mutp53.

Supplementary information

Supplementary Information

Supplementary Information (PDF 14009 kb)

Supplementary Table 1

Supplementary Information (XLSX 23 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Parrales, A., Ranjan, A., Iyer, S. et al. DNAJA1 controls the fate of misfolded mutant p53 through the mevalonate pathway. Nat Cell Biol 18, 1233–1243 (2016). https://doi.org/10.1038/ncb3427

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3427

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