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:

The Hippo pathway kinases LATS1 and LATS2 attenuate cellular responses to heavy metals through phosphorylating MTF1

Abstract

Heavy metals are both integral parts of cells and environmental toxicants, and their deregulation is associated with severe cellular dysfunction and various diseases. Here we show that the Hippo pathway plays a critical role in regulating heavy metal homeostasis. Hippo signalling deficiency promotes the transcription of heavy metal response genes and protects cells from heavy metal-induced toxicity, a process independent of its classic downstream effectors YAP and TAZ. Mechanistically, the Hippo pathway kinase LATS phosphorylates and inhibits MTF1, an essential transcription factor in the heavy metal response, resulting in the loss of heavy metal response gene transcription and cellular protection. Moreover, LATS activity is inhibited following heavy metal treatment, where accumulated zinc directly binds and inhibits LATS. Together, our study reveals an interplay between the Hippo pathway and heavy metals, providing insights into this growth-related pathway in tissue homeostasis and stress response.

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: Hippo signalling deficiency protects cells against heavy metal-induced toxicity.
Fig. 2: Hippo signalling deficiency induces expression of heavy metal response genes.
Fig. 3: The Hippo pathway controls the heavy metal response by inhibiting the transcriptional activity of MTF1.
Fig. 4: LATS1 binds and phosphorylates MTF1 at S152.
Fig. 5: LATS-mediated MTF1 S152 phosphorylation inhibits the heavy metal response.
Fig. 6: Inhibition of the Hippo pathway promotes heavy metal detoxification.
Fig. 7: The Hippo pathway is negatively regulated by heavy metals.
Fig. 8: Zn binds and inhibits LATS.

Similar content being viewed by others

Data availability

The RNAseq dataset is available in the Gene Expression Omnibus under the accession number GSE163156. The lung squamous cell carcinoma RNAseq datasets were downloaded from The Cancer Genome Atlas data portal (https://portal.gdc.cancer.gov/). The protein structure information (PDB: 1ZW8) was downloaded from the Protein Data Bank (https://www.rcsb.org/structure/1zw8). All other data supporting the findings of this study are available from the corresponding author on reasonable request. Source data are provided with this paper.

References

  1. Gunther, V., Lindert, U. & Schaffner, W. The taste of heavy metals: gene regulation by MTF-1. Biochim. Biophys. Acta 1823, 1416–1425 (2012).

    Article  CAS  PubMed  Google Scholar 

  2. Baltaci, A. K., Yuce, K. & Mogulkoc, R. Zinc metabolism and metallothioneins. Biol. Trace Elem. Res. 183, 22–31 (2017).

    Article  PubMed  Google Scholar 

  3. Coyle, P., Philcox, J. C., Carey, L. C. & Rofe, A. M. Metallothionein: the multipurpose protein. Cell. Mol. Life Sci. 59, 627–647 (2002).

    Article  CAS  PubMed  Google Scholar 

  4. Nath, R., Kambadur, R., Gulati, S., Paliwal, V. K. & Sharma, M. Molecular aspects, physiological function, and clinical significance of metallothioneins. Crit. Rev. Food Sci. Nutr. 27, 41–85 (1988).

    Article  CAS  PubMed  Google Scholar 

  5. Hamer, D. H. Metallothionein. Annu. Rev. Biochem. 55, 913–951 (1986).

    Article  CAS  PubMed  Google Scholar 

  6. Robbins, A. H. et al. Refined crystal structure of Cd, Zn metallothionein at 2.0 Å resolution. J. Mol. Biol. 221, 1269–1293 (1991).

    CAS  PubMed  Google Scholar 

  7. Dong, G., Chen, H., Qi, M., Dou, Y. & Wang, Q. Balance between metallothionein and metal response element binding transcription factor 1 is mediated by zinc ions (review). Mol. Med. Rep. 11, 1582–1586 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Davis, S. R. & Cousins, R. J. Metallothionein expression in animals: a physiological perspective on function. J. Nutr. 130, 1085–1088 (2000).

    Article  CAS  PubMed  Google Scholar 

  9. Stuart, G. W., Searle, P. F. & Palmiter, R. D. Identification of multiple metal regulatory elements in mouse metallothionein-I promoter by assaying synthetic sequences. Nature 317, 828–831 (1985).

    Article  CAS  PubMed  Google Scholar 

  10. Westin, G. & Schaffner, W. A zinc-responsive factor interacts with a metal-regulated enhancer element (MRE) of the mouse metallothionein-I gene. EMBO J. 7, 3763–3770 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Radtke, F. et al. Cloned transcription factor MTF-1 activates the mouse metallothionein I promoter. EMBO J. 12, 1355–1362 (1993).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Brugnera, E. et al. Cloning, chromosomal mapping and characterization of the human metal-regulatory transcription factor MTF-1. Nucleic Acids Res. 22, 3167–3173 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Lindert, U., Cramer, M., Meuli, M., Georgiev, O. & Schaffner, W. Metal-responsive transcription factor 1 (MTF-1) activity is regulated by a nonconventional nuclear localization signal and a metal-responsive transactivation domain. Mol. Cell. Biol. 29, 6283–6293 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Bittel, D., Dalton, T., Samson, S. L., Gedamu, L. & Andrews, G. K. The DNA binding activity of metal response element-binding transcription factor-1 is activated in vivo and in vitro by zinc, but not by other transition metals. J. Biol. Chem. 273, 7127–7133 (1998).

    Article  CAS  PubMed  Google Scholar 

  15. Egli, D. et al. Knockout of ‘metal-responsive transcription factor’ MTF-1 in Drosophila by homologous recombination reveals its central role in heavy metal homeostasis. EMBO J. 22, 100–108 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Gunes, C. et al. Embryonic lethality and liver degeneration in mice lacking the metal-responsive transcriptional activator MTF-1. EMBO J. 17, 2846–2854 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Wang, Y. et al. Metal-responsive transcription factor-1 (MTF-1) is essential for embryonic liver development and heavy metal detoxification in the adult liver. FASEB J. 18, 1071–1079 (2004).

    Article  CAS  PubMed  Google Scholar 

  18. Hua, H. et al. Toxicity of Alzheimer’s disease-associated Aβ peptide is ameliorated in a Drosophila model by tight control of zinc and copper availability. Biol. Chem. 392, 919–926 (2011).

    Article  CAS  PubMed  Google Scholar 

  19. Saini, N., Georgiev, O. & Schaffner, W. The parkin mutant phenotype in the fly is largely rescued by metal-responsive transcription factor (MTF-1). Mol. Cell. Biol. 31, 2151–2161 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Shi, Y. et al. The metal-responsive transcription factor-1 protein is elevated in human tumors. Cancer Biol. Ther. 9, 469–476 (2010).

    Article  CAS  PubMed  Google Scholar 

  21. Yu, F. X., Zhao, B. & Guan, K. L. Hippo pathway in organ size control, tissue homeostasis, and cancer. Cell 163, 811–828 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Piccolo, S., Dupont, S. & Cordenonsi, M. The biology of YAP/TAZ: Hippo signaling and beyond. Physiol. Rev. 94, 1287–1312 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Zanconato, F., Cordenonsi, M. & Piccolo, S. YAP/TAZ at the roots of Cancer. Cancer Cell 29, 783–803 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Zheng, Y. et al. Identification of Happyhour/MAP4K as alternative Hpo/Mst-like kinases in the Hippo kinase cascade. Dev. Cell 34, 642–655 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Li, S., Cho, Y. S., Yue, T., Ip, Y. T. & Jiang, J. Overlapping functions of the MAP4K family kinases Hppy and Msn in Hippo signaling. Cell Discov. 1, 15038 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Meng, Z. et al. MAP4K family kinases act in parallel to MST1/2 to activate LATS1/2 in the Hippo pathway. Nat. Commun. 6, 8357 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Li, Q. et al. The conserved Misshapen–Warts–Yorkie pathway acts in enteroblasts to regulate intestinal stem cells in Drosophila. Dev. Cell 31, 291–304 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  31. Li, Q. et al. Ingestion of food particles regulates the mechanosensing Misshapen–Yorkie pathway in Drosophila intestinal growth. Dev. Cell 45, 433–449 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Boggiano, J. C., Vanderzalm, P. J. & Fehon, R. G. Tao-1 phosphorylates Hippo/MST kinases to regulate the Hippo–Salvador–Warts tumor suppressor pathway. Dev. Cell 21, 888–895 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Poon, C. L., Lin, J. I., Zhang, X. & Harvey, K. F. The sterile 20-like kinase Tao-1 controls tissue growth by regulating the Salvador–Warts–Hippo pathway. Dev. Cell 21, 896–906 (2011).

    Article  CAS  PubMed  Google Scholar 

  34. Plouffe, S. W. et al. Characterization of Hippo pathway components by gene inactivation. Mol. Cell 64, 993–1008 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Hamaratoglu, F. et al. The tumour-suppressor genes NF2/Merlin and Expanded act through Hippo signalling to regulate cell proliferation and apoptosis. Nat. Cell Biol. 8, 27–36 (2006).

    Article  CAS  PubMed  Google Scholar 

  36. Yin, F. et al. Spatial organization of Hippo signaling at the plasma membrane mediated by the tumor suppressor Merlin/NF2. Cell 154, 1342–1355 (2013).

    Article  CAS  PubMed  Google Scholar 

  37. Yu, F. X., Meng, Z., Plouffe, S. W. & Guan, K. L. Hippo pathway regulation of gastrointestinal tissues. Annu Rev. Physiol. 77, 201–227 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Wu, S., Liu, Y., Zheng, Y., Dong, J. & Pan, D. The TEAD/TEF family protein Scalloped mediates transcriptional output of the Hippo growth-regulatory pathway. Dev. Cell 14, 388–398 (2008).

    Article  CAS  PubMed  Google Scholar 

  39. Goulev, Y. et al. SCALLOPED interacts with YORKIE, the nuclear effector of the hippo tumor-suppressor pathway in Drosophila. Curr. Biol. 18, 435–441 (2008).

    Article  CAS  PubMed  Google Scholar 

  40. Zhang, L. et al. The TEAD/TEF family of transcription factor Scalloped mediates Hippo signaling in organ size control. Dev. Cell 14, 377–387 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. 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).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. 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).

    Article  CAS  PubMed  Google Scholar 

  44. Wang, Z. et al. Solution structure of a Zap1 zinc-responsive domain provides insights into metalloregulatory transcriptional repression in Saccharomyces cerevisiae. J. Mol. Biol. 357, 1167–1183 (2006).

    Article  CAS  PubMed  Google Scholar 

  45. Li, Y., Kimura, T., Huyck, R. W., Laity, J. H. & Andrews, G. K. Zinc-induced formation of a coactivator complex containing the zinc-sensing transcription factor MTF-1, p300/CBP, and Sp1. Mol. Cell. Biol. 28, 4275–4284 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Datta, J. et al. Metallothionein expression is suppressed in primary human hepatocellular carcinomas and is mediated through inactivation of CCAAT/enhancer binding protein α by phosphatidylinositol 3-kinase signaling cascade. Cancer Res. 67, 2736–2746 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Fan, F. et al. Pharmacological targeting of kinases MST1 and MST2 augments tissue repair and regeneration. Sci. Transl. Med. 8, 352ra108 (2016).

    Article  PubMed  Google Scholar 

  48. Chen, X. et al. Copper sensing function of Drosophila metal-responsive transcription factor-1 is mediated by a tetranuclear Cu(I) cluster. Nucleic Acids Res. 36, 3128–3138 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Selvaraj, A. et al. Metal-responsive transcription factor (MTF-1) handles both extremes, copper load and copper starvation, by activating different genes. Genes Dev. 19, 891–896 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Han, H. et al. Regulation of the Hippo pathway by phosphatidic acid-mediated lipid–protein interaction. Mol. Cell 72, 328–340 e328 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Lv, X. B. et al. PARD3 induces TAZ activation and cell growth by promoting LATS1 and PP1 interaction. EMBO Rep. 16, 975–985 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Rahmat, M. B., Zhang, S. & Koh, C. G. POPX2 is a novel LATS phosphatase that regulates the Hippo pathway. Oncotarget 10, 1525–1538 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Rottenberg, S., Disler, C. & Perego, P. The rediscovery of platinum-based cancer therapy. Nat. Rev. Cancer 21, 37–50 (2021).

    Article  CAS  PubMed  Google Scholar 

  54. Chen, P. H. et al. Kinome screen of ferroptosis reveals a novel role of ATM in regulating iron metabolism. Cell Death Differ. 27, 1008–1022 (2020).

    Article  CAS  PubMed  Google Scholar 

  55. LaRochelle, O., Gagne, V., Charron, J., Soh, J. W. & Seguin, C. Phosphorylation is involved in the activation of metal-regulatory transcription factor 1 in response to metal ions. J. Biol. Chem. 276, 41879–41888 (2001).

    Article  CAS  PubMed  Google Scholar 

  56. Murphy, J. T. et al. Histidine protects against zinc and nickel toxicity in Caenorhabditis elegans. PLoS Genet. 7, e1002013 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Bafaro, E., Liu, Y., Xu, Y. & Dempski, R. E. The emerging role of zinc transporters in cellular homeostasis and cancer. Signal Transduct. Target. Ther. 2, 17029 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  58. Yang, H. & Shu, Y. Cadmium transporters in the kidney and cadmium-induced nephrotoxicity. Int. J. Mol. Sci. 16, 1484–1494 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Gupta, A. & Lutsenko, S. Human copper transporters: mechanism, role in human diseases and therapeutic potential. Future Med. Chem. 1, 1125–1142 (2009).

    Article  CAS  PubMed  Google Scholar 

  60. Garrick, M. D. Human iron transporters. Genes Nutr. 6, 45–54 (2011).

    Article  CAS  PubMed  Google Scholar 

  61. Tsang, T., Davis, C. I. & Brady, D. C. Copper biology. Curr. Biol. 31, R421–R427 (2021).

    Article  CAS  PubMed  Google Scholar 

  62. Zhang, B. et al. Activity of metal-responsive transcription factor 1 by toxic heavy metals and H2O2 in vitro is modulated by metallothionein. Mol. Cell. Biol. 23, 8471–8485 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Pace, N. J. & Weerapana, E. Zinc-binding cysteines: diverse functions and structural motifs. Biomolecules 4, 419–434 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Srivastava, A. & Kumar, M. Prediction of zinc binding sites in proteins using sequence derived information. J. Biomol. Struct. Dyn. 36, 4413–4423 (2018).

    Article  CAS  PubMed  Google Scholar 

  65. Shu, N., Zhou, T. & Hovmoller, S. Prediction of zinc-binding sites in proteins from sequence. Bioinformatics 24, 775–782 (2008).

    Article  CAS  PubMed  Google Scholar 

  66. Beyersmann, D. & Haase, H. Functions of zinc in signaling, proliferation and differentiation of mammalian cells. Biometals 14, 331–341 (2001).

    Article  CAS  PubMed  Google Scholar 

  67. Maret, W. Zinc in cellular regulation: the nature and significance of “Zinc Signals”. Int. J. Mol. Sci. 18, 2285 (2017).

    Article  PubMed Central  Google Scholar 

  68. Brady, D. C. et al. Copper is required for oncogenic BRAF signalling and tumorigenesis. Nature 509, 492–496 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Tsang, T. et al. Copper is an essential regulator of the autophagic kinases ULK1/2 to drive lung adenocarcinoma. Nat. Cell Biol. 22, 412–424 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Turski, M. L. et al. A novel role for copper in Ras/mitogen-activated protein kinase signaling. Mol. Cell. Biol. 32, 1284–1295 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Weinman, S. Calcium-binding proteins: an overview. J. Biol. Buccale 19, 90–98 (1991).

    CAS  PubMed  Google Scholar 

  72. Bialasek, M. et al. Exploiting iron-binding proteins for drug delivery. J. Physiol. Pharmacol. 70, 675–685 (2019).

  73. Karlsson, M. & Kurz, T. Attenuation of iron-binding proteins in ARPE-19 cells reduces their resistance to oxidative stress. Acta Ophthalmol. 94, 556–564 (2016).

    Article  CAS  PubMed  Google Scholar 

  74. Tristao, G. B. et al. Predicting copper-, iron-, and zinc-binding proteins in pathogenic species of the Paracoccidioides genus. Front. Microbiol. 5, 761 (2014).

    PubMed  Google Scholar 

  75. Sanchez, M. et al. Iron regulatory protein-1 and -2: transcriptome-wide definition of binding mRNAs and shaping of the cellular proteome by iron regulatory proteins. Blood 118, e168-79 (2011).

    Article  PubMed  Google Scholar 

  76. Piovesan, D., Profiti, G., Martelli, P. L. & Casadio, R. The human “magnesome”: detecting magnesium binding sites on human proteins. BMC Bioinform. 13, S10 (2012).

    Article  CAS  Google Scholar 

  77. Wang, W., Huang, J. & Chen, J. Angiomotin-like proteins associate with and negatively regulate YAP1. J. Biol. Chem. 286, 4364–4370 (2011).

    Article  CAS  PubMed  Google Scholar 

  78. Wang, W., Chen, L., Ding, Y., Jin, J. & Liao, K. Centrosome separation driven by actin-microfilaments during mitosis is mediated by centrosome-associated tyrosine-phosphorylated cortactin. J. Cell Sci. 121, 1334–1343 (2008).

    Article  CAS  PubMed  Google Scholar 

  79. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    Article  CAS  PubMed  Google Scholar 

  80. Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  81. Lee, T. I., Johnstone, S. E. & Young, R. A. Chromatin immunoprecipitation and microarray-based analysis of protein location. Nat. Protoc. 1, 729–748 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Wang, X. et al. The proteasome-interacting Ecm29 protein disassembles the 26S proteasome in response to oxidative stress. J. Biol. Chem. 292, 16310–16320 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Maier, J. A. et al. ff14SB: improving the accuracy of protein side chain and backbone parameters from ff99SB. J. Chem. Theory Comput. 11, 3696–3713 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Peters, M. B. et al. Structural survey of zinc containing proteins and the development of the zinc AMBER Force Field (ZAFF). J. Chem. Theory Comput. 6, 2935–2947 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Roe, D. R. & Cheatham, T. E.III PTRAJ and CPPTRAJ: software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 9, 3084–3095 (2013).

    Article  CAS  PubMed  Google Scholar 

  86. Johnson, M. S., Sutcliffe, M. J. & Blundell, T. L. Molecular anatomy: phyletic relationships derived from three-dimensional structures of proteins. J. Mol. Evol. 30, 43–59 (1990).

    Article  CAS  PubMed  Google Scholar 

  87. Krissinel, E. & Henrick, K. Secondary-structure matching (SSM), a new tool for fast protein structure alignment in three dimensions. Acta Crystallogr. D 60, 2256–2268 (2004).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank K.-L. Guan (University of California, San Diego), D. Pan (The University of Texas Southwestern Medical Center), D. Fruman (University of California, Irvine) and A. Edinger (University of California, Irvine) for comments on this study. We thank J. W. C. Leung (University of Arkansas for Medical Sciences), R. Edwards (University of California, Irvine), G. MacGregor (University of California, Irvine) and R. Geertsema (University of California, Irvine) for their technical help. This work was supported by NIH grants (grant nos R01GM126048 and R21ES031642), an American Cancer Society Research Scholar grant (grant no. RSG-18-009-01-CCG), a University of California Cancer Research Coordinating Committee Research Award (grant no. CRN-19-585568), and a Tower Cancer Research Foundation Cancer Free Generation Career Development Grant to W.W.; and was also supported in part by NIH grants to R.L. (grant no. R35GM130367), K.W.Y.C. (grant no. R35GM139617) and L.H. (grant no. R01GM130144). W.W. is a member of the Chao Family Comprehensive Cancer Center (award no. P30CA062203) at the University of California, Irvine.

Author information

Authors and Affiliations

Authors

Contributions

H.H. and W.W. conceived and designed the study. W.W. supervised the study. H.H. and H.J.N. performed all of the experiments with the assistance from G.S., R.E.V. and B.Y. L.H. and R.W. performed the Drosophila survival study. C.Y. and L.H. performed the mass spectrometry analysis. L.B. helped with the in vitro kinase assay. J.N., L.Z., R.Q. and R.L. performed the simulation analysis. J.J.Z. and K.W.Y.C. helped with the RNAseq analysis. D.A.F. helped with the circular dichroism analysis. W.W. wrote the manuscript.

Corresponding authors

Correspondence to Rahul Warrior or Wenqi Wang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review information

Nature Cell Biology thanks the anonymous reviewers for their contribution to the peer review of this work. Peer reviewer reports are available.

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 Inhibition of the Hippo pathway activates heavy metal response.

(a) Validation of the YAP-5SA overexpressed HEK239A cells. Western blot was performed using the indicated antibodies. (b) Validation of the YAP/TAZ shRNAs-transduced LATS1/2 DKO HEK293A cells. Western blot was performed using the indicated antibodies. (c,d) Western blot was performed using the indicated antibodies (c). MT1A gene transcription was examined in the indicated cells treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours by qPCR (mean ± s.d., n = 3 biological replicates) (d). *** p < 0.001 (two-tailed Student’s t-test). Relative expression fold change numbers are shown. (e,f) Transcription of MT1F (e) and CTGF (f) was examined in the HEK293A cells cultured under serum starvation and serum stimulation by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). (g,h) Transcription of MT1F (g) and CTGF (h) was examined in the HEK293A cells cultured under high density and low density by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). (i,j) Serum-starved wild-type and the LATS1/2 DKO HEK293A cells were treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours. Transcription of MT1A (i) and MT1F (j) was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). Relative expression fold change numbers are shown. (k-l) Wild-type and the LATS1/2 DKO HEK293A cells were treated with 250 μM Zn with the indicated anions for 4 hours. The transcription of MT1A (i) and MT1F (j) was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). (m) Validation of the indicated HEK293A cells stably expressing SFB-MTF1. Western blot was performed using the indicated antibodies. Data shown represent 2 independent experiments in a-c, m.

Source data

Extended Data Fig. 2 LATS1 binds and phosphorylates MTF1.

(a) HEK293A and Hep3B cells stably expressing SFB-tagged MTF1 were subjected to pulldown assay using S protein beads. Western blot was performed using the indicated antibodies. (b) HEK293A cells were transfected with the construct encoding SFB-tagged MTF1, treated with CdCl2 (50 µM) and ZnCl2 (250 µM) for 1 hour, and subjected to pulldown assay using S protein beads. Western blot was performed using the indicated antibodies. (c) SFB-tagged LATS1 and its kinase dead (K734R) mutant were expressed in HEK293T cells, purified using S protein beads, and subjected to the ATP-γ-S-based in vitro kinase assay. Bacterially purified MBP-MTF1-R2, MBP-R2-5A mutant (S152A/T153A/T241A/S301A/T302A) and GST-YAP proteins were used as substrates. Western blot was performed using the indicated antibodies. (d) HEK293A cells stably expressing SFB-tagged MTF1 were transfected with the constructs encoding HA-tagged LATS1 and its kinase dead (K734R) mutant, and subjected to pulldown assay using S protein beads. Western blot was performed using the indicated antibodies. Arrow indicates the exogenously expressed SFB-MTF1. Pound sign indicates endogenous MTF1. (e) SFB-tagged LATS1, STK38 and STK38L were expressed in HEK293T cells, purified using S protein beads, and subjected to in vitro kinase assay. Bacterially purified MBP-tagged MTF1-R2 protein was used as substrate. Western blot was performed using the indicated antibodies. Data shown represent 2 independent experiments in a-e.

Source data

Extended Data Fig. 3 MTF1 S152 phosphorylation inhibits heavy metal response.

(a,b) The indicated HEK293A cells were treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours. The transcription of MT1A (a) and MT1F (b) was examined by qPCR (mean ± s.d., n = 3 biological replicates). * p < 0.05, ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). ns, no significance. Relative expression fold change numbers are shown. (c) Validation of the MTF1 KO HEK293A cells reconstituted with the indicated SFB-tagged MTF1 and its mutants. Western blot was performed using the indicated antibodies. Arrow indicates the exogenously expressed SFB-tagged MTF1 and its mutants. (d,e) The MTF1KO HEK293A cells were reconstituted with the indicated SFB-tagged MTF1 and its mutants, treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 12 hours, visualized by crystal violet staining (d), and quantified for relative viability (mean ± s.d., n = 4 biological replicates) (e). *** p < 0.001 (two-tailed Student’s t-test). (f,g) The transcription of MT1A (f) and MT1F (g) was examined in wild-type and the MTF1 KO HEK293A cells reconstituted with the indicated SFB-tagged MTF1 and its mutants by qPCR (mean ± s.d., n = 3 biological replicates). Cells were treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours. * p < 0.05, ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). ns, no significance. (h) The structure comparison of the ZAP1-ZNF1 in khaki with the phosphorylated ZAP1-ZNF1 (p-S591) in pink (left) and with ZAP1-ZNF1 S591D mutant in blue (right). The overall RMSD values of wild-type ZAP1-ZNF1 with its phosphorylated form (p-S591) and S591D mutant are shown. Arrows indicate the S591 site. (i) The MTF1 KO HEK293A cells were reconstituted with the indicated SFB-MTF1 and its mutants, and treated with ZnCl2 (250 μM) for 1 hour. Immunofluorescent staining was performed. Scale bar, 40 μm. Data shown represent 3 independent experiments. (j) The MTF1 KO HEK293A cells were reconstituted with the indicated SFB-MTF1 and its mutants, treated with ZnCl2 (250 μM) for 1 hour, and subjected to pulldown assay using S protein beads. Western blot was performed using the indicated antibodies. Data shown represent 2 independent experiments in c, j.

Source data

Extended Data Fig. 4 Characterization of XMU-MP-1 in regulating the Hippo pathway and heavy metal response.

(a) Immunohistochemical analyses of Mtf1 and its S151 phosphorylation were performed in the liver and kidney tissues of the mice treated with DMSO, XMU-MP-1 alone or combined with VP post CdCl2 intoxication. The indicated regions in the box are shown 3 times enlarged. Scale bar, 150 μm. Data shown represent 3 independent experiments. (b) Wild-type and the MST1/2 DKO Hep3B and HEK293A cells were treated with DMSO and XMU-MP-1 (10 μM) for 4 hours, and subjected to Western blot analysis using the indicated antibodies. (c,d) The indicated Hep3B cells (c) and HEK293A cells (d) were pre-incubated with XMU-MP-1 (10 μM) for 2 hours, and treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours. The transcription of MT1F was examined by qPCR (mean ± s.d., n = 3 biological replicates). * p < 0.05, ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). ns, no significance. (e-h) Hep3B and HEK293A cells stably expressing SFB-tagged TAOK1 (e), MAP4K2 (f), MAP4K4 (g) and MAP4K6 (h) were serum starved for 24 hours, and treated with XMU-MP-1 (10 μM) for 4 hours. The indicated SFB-tagged kinases were purified using S protein beads and subjected to in vitro kinase assay using the bacterially purified MBP-tagged LATS1-C3 protein as substrate. Western blot was performed using the indicated antibodies Data shown represent 2 independent experiments in b, e-h.

Source data

Extended Data Fig. 5 The Hippo pathway-mediated regulation of MTF1 and heavy metal response is conserved in Drosophila.

(a) Heterozygous mats Drosophila mutant and homozygous mtf1 Drosophila mutant were used for adult viability assay. Adult viability was assayed by feeding adult Drosophila (~7-day old) with normal food supplemented with 5 mM CuSO4 and determining the number of survivors daily. **, p < 0.01; ***, p < 0.001 (Log-rank test). n = 19 for wild-type Drosophila, n = 14 for mtf-/- Drosophila, n = 24 for mats−/+ Drosophila. (b) The MTF1 KO HEK293A cells were transfected with the constructs encoding the SFB-tagged human MTF1 and fly-mtf1. Western blot was performed using the indicated antibodies. (c) The MTF1 KO HEK293A cells were transfected with the constructs encoding the SFB-tagged human MTF1 and fly-mtf1, and treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours. MT1A gene transcription was examined by qPCR (mean ± s.d., n = 3 biological replicates). *** p < 0.001 (two-tailed Student’s t-test). (d) HEK293A cells were transfected with the constructs encoding the SFB-tagged fly-mtf1 and its S126D mutant, treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 1 hour, and subjected to immunofluorescent staining. Scale bar, 40 μm (e) The LATS/MTF1 TKO HEK293A cells were transduced with the constructs encoding the SFB-tagged fly-mtf1 and its S126D mutant. Western blot was performed using the indicated antibodies. (f) The LATS/MTF1 TKO HEK293A cells stably expressing the SFB-tagged fly-mtf1 and its S126D mutant were treated with CdCl2 (50 μM), ZnCl2 (250 μM) and CuSO4 (250 μM) for 4 hours. MT1F gene transcription was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). Data shown represent 2 independent experiments in b, d, e.

Source data

Extended Data Fig. 6 Heavy metals inhibit the Hippo pathway in different cells and tissues.

(a) Hep3B cells were serum starved for 24 hours, and treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for the indicated time points. Western blot was performed using the indicated antibodies. (b) MCF10A and HeLa cells were serum starved for 24 hours, treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 1 hour, and subjected to immunofluorescent staining. Scale bar, 40 μm (c) MCF10A and HeLa cells were serum starved for 24 hours, and treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 4 hours. The transcription of YAP downstream genes CTGF and CYR61 was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). (d) Hep3B cells stably expressing SFB-tagged MTF1 were serum starved for 24 hours, treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for the indicated time points, and subjected to pulldown assay using S protein beads. Western blot was performed using the indicated antibodies. (e) Hep3B and HEK293A cells were serum starved for 24 hours, treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 1 hour, and subjected to immunoprecipitation using LATS1 antibody. Western blot was performed using the indicated antibodies. Arrow indicates the correct p-LATS1 S909 signal. Asterisk indicates the non-specific signal. (f) The transcription of heavy metal response gene Mt1 was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). Relative expression fold change numbers are shown. (g) The transcription of Yap downstream gene Ctgf was examined by qPCR (mean ± s.d., n = 3 biological replicates). * p < 0.05, ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). Relative expression fold change numbers are shown. (h-j) The Human Protein Atlas (https://www.proteinatlas.org/) was used to examine the expression of MTF1 and LATS1 in different human tissues (h), human cells (i) and commonly used human cell lines (j). Data shown represent 2 independent experiments in a, b, d, e.

Source data

Extended Data Fig. 7 Zn binds and inhibits LATS.

(a) Illustration of the metal beads. (b) HEK293T cells were transfected with the constructs encoding the indicated Hippo pathway components and subjected to pulldown assay using Cu beads. The input blot is shared with the one in Fig. 8c. (c,d) HEK293T cells were transfected with the constructs encoding SFB-tagged LATS1 (c) and LATS2 (d), and subjected to pulldown assay using Zn beads in the presence of TPEN at the indicated concentrations. (e-f) SFB-tagged MST1 was expressed in HEK293T cells, purified using S protein beads, and subjected to in vitro kinase assay in the presence of CaCl2 (e) and ZnCl2 (f) at the indicated concentrations. MBP-tagged LATS1-C3 protein was used as substrate. (g) SFB-tagged MAP4K2 was expressed in HEK293T cells, purified using S protein beads, and subjected to in vitro kinase assay in the presence of ZnCl2 at the indicated concentrations. MBP-tagged LATS1-C3 protein was used as substrate. (h) SFB-tagged MST1 was expressed in HEK293T cells, purified using S protein beads, and subjected to in vitro kinase assay in the presence of ZnCl2 at the indicated concentrations. GST-tagged MOB1 protein was used as substrate. (i-j) HEK293A cells were transfected with the constructs encoding the SFB-tagged LATS1 with HA-tagged PPP1CA (i) and POPX2 (j), treated with ZnCl2 (250 μM) and CaCl2 (250 μM) for 1 hour, and subjected to pulldown assay using S protein beads. (k-l) SFB-tagged LATS1 was expressed in HEK293T cells, purified using S protein beads, and subjected to in vitro kinase assay in the presence of CaCl2 (k) or ZnCl2 (l) at the indicated concentrations. GST-tagged YAP protein was used as substrate. (m) SFB-tagged LATS2 was expressed in HEK293T cells, purified using S protein beads, and subjected to in vitro kinase assay in the presence of ZnCl2 at the indicated concentrations. GST-tagged YAP protein was used as substrate. (n) The potential Zn-binding Cys, His, Glu and Asp residues in the LATS1-C3 region are shown. The conserved Cys and His residues are highlighted in yellow. The conserved Glu and Asp residues are highlighted in blue. LATS1 T1079 site is indicated by arrow. Data shown represent 2 independent experiments in b-m.

Source data

Extended Data Fig. 8 The Hippo pathway mediates cellular response to different heavy metals.

(a) HEK293A cells were transfected with the construct encoding SFB-tagged MOB1, treated with CdCl2 (50 μM) and ZnCl2 (250 μM) for 1 hour, and subjected to pulldown assay using S protein beads. Western blot was performed using the indicated antibodies. short, short exposure; long, long exposure. (b) HEK293T cells were transfected with the construct encoding SFB-tagged MOB1 and subjected to pulldown assay using the indicated metal beads. (c) Hep3B and HEK293A cells were serum starved for 24 hours and treated with CuSO4 (250 μM) and ZnCl2 (250 μM) for 4 hours. The transcription of MT1A, MT1F and CYR61 was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.01, *** p < 0.001 (two-tailed Student’s t-test). (d) The indicated cells were treated with CuSO4 (250 μM) and ZnCl2 (250 μM) for 4 hours. The transcription of MT1A was examined by qPCR (mean ± s.d., n = 3 biological replicates). *** p < 0.001 (two-tailed Student’s t-test). (e) SFB-tagged LATS1 was expressed in HEK293T cells, purified using S protein beads, washed thoroughly with high-salt buffer containing 250 mM NaCl, incubated with Zn chelator TPEN (25 μM) or Cu chelator TTM (25 μM), and subjected to in vitro kinase assay. GST-YAP protein used as substrate. (f) Illustration of cisplatin and carboplatin compound structures, where platinum is labelled in red. (g) The indicated cells were treated with cisplatin (10 μM) and carboplatin (100 μM) for 6 hours. The transcription of MT1A was examined by qPCR (mean ± s.d., n = 3 biological replicates). ** p < 0.05, *** p < 0.001 (two-tailed Student’s t-test). ns, no significance. Relative expression fold change numbers are shown. (h-k) The lung squamous cell carcinoma (LUSC) RNAseq datasets with clinic data were downloaded from the Cancer Genome Atlas (TCGA) data portal. The indicated gene expression from a total of 119 patients treated with cisplatin or carboplatin were subjected to Spearman correction analysis. Correlation coefficient R value and p value were calculated by GraphPad prism software. (l) A proposed model for the Hippo pathway-mediated heavy metal response through MTF1. Data shown represent 2 independent experiments in a, b, e.

Source data

Supplementary information

Reporting Summary

Peer Review Information

Supplementary Table 1

Supplementary Table 1. A total of 181 LATS1/2-regulated heavy metal response genes are listed based on clustered groups as shown in Fig. 2c. Supplementary Table 2. Sequence information of the qPCR primers used in this study.

Supplementary Video 1

The H2O control-treated mice are shown (upper left). Mice were pretreated with CdCl2 for 1 h and subjected to treatment with vehicle (upper right), XMU-MP-1 (lower left) or XMU-MP-1 + verteporfin (lower right). This video was taken 4 h post the CdCl2 intoxication.

Supplementary Video 2

The H2O control-treated mice are shown (upper left). Mice were pretreated with CdCl2 for 1 h and subjected to treatment with vehicle (upper right), XMU-MP-1 (lower left) and XMU-MP-1 + verteporfin (lower right). This video was taken 100 h post the CdCl2 intoxication.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 3

Unprocessed western blots.

Source Data Fig. 4

Unprocessed western blots and gel.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 5

Unprocessed western blots.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 6

Unprocessed western blots.

Source Data Fig. 7

Statistical source data.

Source Data Fig. 7

Unprocessed western blots.

Source Data Fig. 8

Statistical source data.

Source Data Fig. 8

Unprocessed western blots and gel.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 5

Unprocessed western blots.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Unprocessed western blots.

Source Data Extended Data Fig. 7

Unprocessed western blots.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 8

Unprocessed western blots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Han, H., Nakaoka, H.J., Hofmann, L. et al. The Hippo pathway kinases LATS1 and LATS2 attenuate cellular responses to heavy metals through phosphorylating MTF1. Nat Cell Biol 24, 74–87 (2022). https://doi.org/10.1038/s41556-021-00813-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-021-00813-8

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