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:

Cellular energy stress induces AMPK-mediated regulation of YAP and the Hippo pathway

This article has been updated

Abstract

YAP (Yes-associated protein) is a transcription co-activator in the Hippo tumour suppressor pathway and controls cell growth, tissue homeostasis and organ size. YAP is inhibited by the kinase Lats, which phosphorylates YAP to induce its cytoplasmic localization and proteasomal degradation. YAP induces gene expression by binding to the TEAD family transcription factors. Dysregulation of the Hippo–YAP pathway is frequently observed in human cancers. Here we show that cellular energy stress induces YAP phosphorylation, in part due to AMPK-dependent Lats activation, thereby inhibiting YAP activity. Moreover, AMPK directly phosphorylates YAP Ser 94, a residue essential for the interaction with TEAD, thus disrupting the YAP–TEAD interaction. AMPK-induced YAP inhibition can suppress oncogenic transformation of Lats-null cells with high YAP activity. Our study establishes a molecular mechanism and functional significance of AMPK in linking cellular energy status to the Hippo–YAP pathway.

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: Cellular energy starvation activates Lats and inhibits YAP.
Figure 2: AMPK is required for Hippo–YAP regulation by energy stress.
Figure 3: Energy stress induces YAP phosphorylation via both Lats-dependent and Lats-independent mechanisms.
Figure 4: AMPK phosphorylates Ser 94 in YAP.
Figure 5: AMPK inhibits YAP activity through phosphorylation of Ser 94.
Figure 6: Energy stress attenuates tumorigenicity of Lats1/2 DKO MEFs.
Figure 7: AMPK is required for energy stress to inhibit YAP activity.

Similar content being viewed by others

Change history

  • 14 April 2015

    In the Methods section 'Antibodies and reagents' there was a typographical error in the company name GeneTex. The sentence should have read 'Anti-phosphorylated Ser 94 antibody was generated by immunizing rabbits with phosphopeptides (Abbiotec and GeneTex, 1:500)'. This has now been corrected online.

References

  1. Oh, H. & Irvine, K. D. Yorkie: the final destination of Hippo signaling. Trends Cell Biol. 20, 410–417 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Pan, D. The Hippo signaling pathway in development and cancer. Dev. Cell 19, 491–505 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Zhao, B., Tumaneng, K. & Guan, K. L. The Hippo pathway in organ size control, tissue regeneration and stem cell self-renewal. Nat. Cell Biol. 13, 877–883 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Zhao, B. et al. Inactivation of YAP oncoprotein by the Hippo pathway is involved in cell contact inhibition and tissue growth control. Genes Dev. 21, 2747–2761 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Dong, J. et al. Elucidation of a universal size-control mechanism in Drosophila and mammals. Cell 130, 1120–1133 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Hao, Y., Chun, A., Cheung, K., Rashidi, B. & Yang, X. Tumor suppressor LATS1 is a negative regulator of oncogene YAP. J. Biol. Chem. 283, 5496–5509 (2008).

    CAS  PubMed  Google Scholar 

  7. Zhao, B., Li, L., Tumaneng, K., Wang, C. Y. & Guan, K. L. A coordinated phosphorylation by Lats and CK1 regulates YAP stability through SCFβ−TRCP. Genes Dev. 24, 72–85 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Liu, C. Y. et al. The Hippo tumor pathway promotes TAZ degradation by phosphorylating a phosphodegron and recruiting the SCF{β}−TrCP E3 ligase. J. Biol. Chem. 285, 37159–37169 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Lu, L. et al. Hippo signaling is a potent in vivo growth and tumor suppressor pathway in the mammalian liver. Proc. Natl Acad. Sci. USA 107, 1437–1442 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Cai, J. et al. The Hippo signaling pathway restricts the oncogenic potential of an intestinal regeneration program. Genes Dev. 24, 2383–2388 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Lee, K. P. et al. The Hippo-Salvador pathway restrains hepatic oval cell proliferation, liver size, and liver tumorigenesis. Proc. Natl Acad. Sci. USA 107, 8248–8253 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Ota, M. & Sasaki, H. Mammalian Tead proteins regulate cell proliferation and contact inhibition as transcriptional mediators of Hippo signaling. Development 135, 4059–4069 (2008).

    CAS  PubMed  Google Scholar 

  13. Zhao, B. et al. TEAD mediates YAP-dependent gene induction and growth control. Genes Dev. 22, 1962–1971 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Halder, G. & Johnson, R. L. Hippo signaling: growth control and beyond. Development 138, 9–22 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Harvey, K. F., Zhang, X. & Thomas, D. M. The Hippo pathway and human cancer. Nat. Rev. Cancer 13, 246–257 (2013).

    CAS  PubMed  Google Scholar 

  16. Johnson, R. & Halder, G. The two faces of Hippo: targeting the Hippo pathway for regenerative medicine and cancer treatment. Nat. Rev. Drug Discov. 13, 63–79 (2014).

    CAS  PubMed  Google Scholar 

  17. Strano, S. et al. The transcriptional coactivator Yes-associated protein drives p73 gene-target specificity in response to DNA damage. Mol. Cell 18, 447–459 (2005).

    CAS  PubMed  Google Scholar 

  18. Lapi, E. et al. PML, YAP, and p73 are components of a proapoptotic autoregulatory feedback loop. Mol. Cell 32, 803–814 (2008).

    CAS  PubMed  Google Scholar 

  19. Cottini, F. et al. Rescue of Hippo coactivator YAP1 triggers DNA damage-induced apoptosis in hematological cancers. Nat. Med. 20, 599–606 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Schlegelmilch, K. et al. Yap1 acts downstream of α-catenin to control epidermal proliferation. Cell 144, 782–795 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Zhao, B. et al. Cell detachment activates the Hippo pathway via cytoskeleton reorganization to induce anoikis. Genes Dev. 26, 54–68 (2012).

    PubMed  PubMed Central  Google Scholar 

  22. Yu, F. X. et al. Regulation of the Hippo-YAP pathway by G-protein-coupled receptor signaling. Cell 150, 780–791 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Mo, J. S., Yu, F. X., Gong, R., Brown, J. H. & Guan, K. L. Regulation of the Hippo-YAP pathway by protease-activated receptors (PARs). Genes Dev. 26, 2138–2143 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Miller, E. et al. Identification of serum-derived sphingosine-1-phosphate as a small molecule regulator of YAP. Chem. Biol. 19, 955–962 (2012).

    CAS  PubMed  Google Scholar 

  25. Yu, F. X. et al. Protein kinase A activates the Hippo pathway to modulate cell proliferation and differentiation. Genes Dev. 27, 1223–1232 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Dupont, S. et al. Role of YAP/TAZ in mechanotransduction. Nature 474, 179–183 (2011).

    CAS  PubMed  Google Scholar 

  27. Wada, K., Itoga, K., Okano, T., Yonemura, S. & Sasaki, H. Hippo pathway regulation by cell morphology and stress fibers. Development 138, 3907–3914 (2011).

    CAS  PubMed  Google Scholar 

  28. Sansores-Garcia, L. et al. Modulating F-actin organization induces organ growth by affecting the Hippo pathway. EMBO J. 30, 2325–2335 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Hardie, D. G. AMP-activated protein kinase: an energy sensor that regulates all aspects of cell function. Genes Dev. 25, 1895–1908 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Vassilev, A., Kaneko, K. J., Shu, H., Zhao, Y. & DePamphilis, M. L. TEAD/TEF transcription factors utilize the activation domain of YAP65, a Src/Yes-associated protein localized in the cytoplasm. Genes Dev. 15, 1229–1241 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Mahoney, W. M. Jr, Hong, J. H., Yaffe, M. B. & Farrance, I. K. The transcriptional co-activator TAZ interacts differentially with transcriptional enhancer factor-1 (TEF-1) family members. Biochem. J. 388, 217–225 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Tian, W., Yu, J., Tomchick, D. R., Pan, D. & Luo, X. Structural and functional analysis of the YAP-binding domain of human TEAD2. Proc. Natl Acad. Sci. USA 107, 7293–7298 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Chan, E. H. et al. The Ste20-like kinase Mst2 activates the human large tumor suppressor kinase Lats1. Oncogene 24, 2076–2086 (2005).

    CAS  PubMed  Google Scholar 

  34. Praskova, M., Xia, F. & Avruch, J. MOBKL1A/MOBKL1B phosphorylation by MST1 and MST2 inhibits cell proliferation. Curr. Biol. 18, 311–321 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Knowler, W. C. et al. Reduction in the incidence of type 2 diabetes with lifestyle intervention or metformin. N. Engl. J. Med. 346, 393–403 (2002).

    CAS  PubMed  Google Scholar 

  36. Zhou, G. et al. Role of AMP-activated protein kinase in mechanism of metformin action. J. Clin. Invest. 108, 1167–1174 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Evans, J. M., Donnelly, L. A., Emslie-Smith, A. M., Alessi, D. R. & Morris, A. D. Metformin and reduced risk of cancer in diabetic patients. BMJ 330, 1304–1305 (2005).

    PubMed  PubMed Central  Google Scholar 

  38. Stambolic, V., Woodgett, J. R., Fantus, I. G., Pritchard, K. I. & Goodwin, P. J. Utility of metformin in breast cancer treatment, is neoangiogenesis a risk factor? Breast Cancer Res. Treat. 114, 387–389 (2009).

    PubMed  Google Scholar 

  39. Moreno, D., Knecht, E., Viollet, B. & Sanz, P. A769662, a novel activator of AMP-activated protein kinase, inhibits non-proteolytic components of the 26S proteasome by an AMPK-independent mechanism. FEBS Lett. 582, 2650–2654 (2008).

    CAS  PubMed  Google Scholar 

  40. Zhou, D. et al. Mst1 and Mst2 maintain hepatocyte quiescence and suppress hepatocellular carcinoma development through inactivation of the Yap1 oncogene. Cancer Cell 16, 425–438 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Sorrentino, G. et al. Metabolic control of YAP and TAZ by the mevalonate pathway. Nat. Cell Biol. 16, 357–366 (2014).

    CAS  PubMed  Google Scholar 

  42. Kim, M. et al. cAMP/PKA signalling reinforces the LATS-YAP pathway to fully suppress YAP in response to actin cytoskeletal changes. EMBO J. 32, 1543–1555 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Inoki, K., Zhu, T. & Guan, K. L. TSC2 mediates cellular energy response to control cell growth and survival. Cell 115, 577–590 (2003).

    CAS  PubMed  Google Scholar 

  44. Kim, J., Kundu, M., Viollet, B. & Guan, K. L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 13, 132–141 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Budanov, A. V. & Karin, M. p53 target genes sestrin1 and sestrin2 connect genotoxic stress and mTOR signaling. Cell 134, 451–460 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Li, Z. et al. Structural insights into the YAP and TEAD complex. Genes Dev. 24, 235–240 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Chen, L. et al. Structural basis of YAP recognition by TEAD4 in the Hippo pathway. Genes Dev. 24, 290–300 (2010).

    PubMed  PubMed Central  Google Scholar 

  48. Yu, F. X. & Guan, K. L. The Hippo pathway: regulators and regulations. Genes Dev. 27, 355–371 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. DeRan, M. et al. Energy stress regulates Hippo-YAP signaling involving AMPK-mediated regulation of angiomotin-like 1 protein. Cell Rep. 9, 495–503 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Zhang, H. et al. TEAD transcription factors mediate the function of TAZ in cell growth and epithelial-mesenchymal transition. J. Biological Chem. 284, 13355–13362 (2009).

    CAS  Google Scholar 

  51. Wang, W. et al. AMPK modulates Hippo pathway activity to regulate energy homeostasis. Nat. Cell Biol. 17, http://dx.doi.org/10.1038/ncb3113 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank A. Hong and F. Flores for critical reading of the manuscript. This work was supported by grants from National Institutes of Health (CA132809, EY022611, DEO15964 and CA23100, K-L.G.). C.G.H. is supported by a Postdoctoral Fellowship from the Danish Council for Independent Research, Natural Sciences.

Author information

Authors and Affiliations

Authors

Contributions

J-S.M. and K-L.G. designed the experiments, analysed data and wrote the paper. J-S.M. performed the experiments with assistance from Z.M., Y.C.K., H.W.P., C.G.H. and S.K.; D-S.L. established the Lats knockout MEFs. All authors discussed the results and commented on the manuscript.

Corresponding author

Correspondence to Kun-Liang Guan.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Energy stress induces YAP phosphorylation.

(a) Osmotic stress does not affect YAP and TAZ mobility shift. HEK293A cells were treated with different concentrations of 2-DG or sorbitol for 1 h. Cells were then lysed and subjected to immunoblotting using the indicated antibodies. (b) 2-DG induces YAP phosphorylation. HEK293A cells were treated with different concentrations of 2-DG for 30 min. Cells were then lysed and cell lysates were subjected to immunoblotting using the indicated antibodies. (c) Experiments are the same as in panel an exception cells were treated with 25 mM 2-DG for various times. (df) Effect of 2-DG on YAP phosphorylation in different cell lines. C2C12 cells were treated with 10 mM and 25 mM 2-DG (d); MCF10A cells were treated with 10 and 25 mM 2-DG (e); HeLa cells were treated with 10 and 25 mM (f). (g) Glucose starvation inhibits YAP and TEAD1 interaction. HEK293A cells were starved with glucose for 2 h as indicated. Endogenous YAP/TAZ and the co-immunoprecipitated TEAD were detected by western blot.

Supplementary Figure 2 2-DG does not affect MST phosphorylation.

(a) Effect of MST overexpression on 2-DG induced YAP phosphorylation. HEK293A cells were transiently co-transfected with indicated plasmids. After transfection, cells were treated with 25 mM 2-DG for 1 h and YAP phosphorylation status was determined by phos-tag gel. (b) 2-DG does not affect MST1 phosphorylation. Endogenous MST1 was immunoprecipitated from control or 2-DG (25 mM, 2 h) treated HEK293A cells. MST1 phosphorylation was detected by a phospho-MST (T183) specific antibody.

Supplementary Figure 3 Activation of AMPK increases YAP phosphorylation in various cell lines.

(a) Genotyping of AMPKa1 and AMPKa2 wild-type (WT) and mutant (KO) alleles. DNA was isolated from AMPK+/+ or AMPK−/− cells as indicated. Genotyping was performed by PCR using primers specific to wild-type AMPKa (left side of the dash line) or AMPKa KO (right side of the dash line). The PCR products were run on the same agarose gel. Samples in the left panel are PCR products using primer specific to AMPKa1 where samples in the right panel are PCR products using primers specific to AMPKa2. (b) Cells were treated with 25 mM 2-DG for various times as indicated and lysed. Lysates were incubated with lambda phosphatase (λ PPase) as indicated for 1 h. Endogenous YAP mobility was examined by western blot. (c) AICAR increases YAP phosphorylation in hepatocytes. Primary mouse hepatocytes were treated with different doses of AICAR for 8 h. Cell lysates were used for immunoblotting with indicated antibodies. (d) HepG2 cells were treated with different doses of AICAR for 2 h. (e) HepG2 cells were treated with different doses of A769662 for 2 h.

Supplementary Figure 4 Identification of AMPK phosphorylation sites in YAP.

(a) AMPK induces mobility shift of YAP-5SA. GST-YAP 5SA was used as substrates for in vitro AMPK phosphorylation. GST- YAP mobility was examined by western blot (b) Mass spectrometry analyses of AMPK-phosphorylated YAP. GST-YAP-5SA was purified from E.coli and was used as a substrate for an in vitro AMPK assay. After electrophoresis, the phosphorylated GST-YAP 5SA band was cut and digested with trypsin/thermolysine followed by mass spectrometry analyses. The green coloured bars (top) and amino acid residues (lower part) are sequences that were detected by mass spectrometry. Phosphorylation sites identified by mass spectrometry are indicated on top of the amino acid sequence. (c) S94 is the major AMPK phosphorylation site in the YAP (51-121) fragment. Experiments were similar to panel a. (d) All three indicated residues of YAP were mutated to alanine. A luciferase reporter controlled by multiple TEAD binding sequences was transfected into HEK293T cells together with 5 × UAS-luciferase reporter for Gal4-TEAD4 and Renilla constructs and indicated plasmids. After 48 h, the firefly luciferase activity was measured and normalized to the co-transfected Renilla luciferase internal control. (d) Two different substrates of AMPK, TSC2 and ULK1, and YAP truncations were expressed, purified, and used as substrates for in vitro AMPK phosphorylation in the presence of 32P-ATP Phosphorylation was detected by autoradiography. GST-ULK1 (279-425), TSC2 (1300-1367) and YAP fragment proteins were detected by Coomassie staining. (e) Experiments are same as in panel c. except that a GST-YAP truncated form and UKL1 were used as substrates for in vitro AMPK phosphorylation in the presence of 32P-ATP for the indicated times. (f) Experiments are the same as in panel d. exception GST-YAP WT and UKL1 used as substrate for the indicated times. (g) AMPK has no effect on the YAP and RUNX2 interaction. HEK293A cells were transfected with the indicated plasmids. Flag-RUNX2 was immunoprecipitated and the co-precipitated HA-YAP was detected by Western blot.

Supplementary information

Supplementary Information

Supplementary Information (PDF 1808 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mo, JS., Meng, Z., Kim, Y. et al. Cellular energy stress induces AMPK-mediated regulation of YAP and the Hippo pathway. Nat Cell Biol 17, 500–510 (2015). https://doi.org/10.1038/ncb3111

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ncb3111

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing