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:

ERBB2 drives YAP activation and EMT-like processes during cardiac regeneration

Abstract

Cardiomyocyte loss after injury results in adverse remodelling and fibrosis, inevitably leading to heart failure. The ERBB2–Neuregulin and Hippo–YAP signalling pathways are key mediators of heart regeneration, yet the crosstalk between them is unclear. We demonstrate that transient overexpression of activated ERBB2 in cardiomyocytes (OE CMs) promotes cardiac regeneration in a heart failure model. OE CMs present an epithelial–mesenchymal transition (EMT)-like regenerative response manifested by cytoskeletal remodelling, junction dissolution, migration and extracellular matrix turnover. We identified YAP as a critical mediator of ERBB2 signalling. In OE CMs, YAP interacts with nuclear-envelope and cytoskeletal components, reflecting an altered mechanical state elicited by ERBB2. We identified two YAP-activating phosphorylations on S352 and S274 in OE CMs, which peak during metaphase, that are ERK dependent and Hippo independent. Viral overexpression of YAP phospho-mutants dampened the proliferative competence of OE CMs. Together, we reveal a potent ERBB2-mediated YAP mechanotransduction signalling, involving EMT-like characteristics, resulting in robust heart regeneration.

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: ERBB2 induces cardiac regeneration in a heart failure model that involves EMT-like processes.
Fig. 2: YAP is activated downstream of ERBB2 signalling in CMs.
Fig. 3: The Hippo pathway is activated in OE hearts.
Fig. 4: YAP is required for ERBB2-related cardiac phenotypes.
Fig. 5: ERBB2 alters the mechanical state of CMs, enhancing the interaction of YAP with cytoskeletal and nuclear-envelope components.
Fig. 6: YAP phosphorylation on S274 and S352 is required for mitosis and occurs downstream of ERK.

Similar content being viewed by others

Data availability

The RNA-seq data were deposited in the Gene Expression Omnibus under accession code GSE144391. The mass spectrometry data have been deposited in ProteomeXchange under accession code PXD020731. 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. Tzahor, E. & Poss, K. D. Cardiac regeneration strategies: staying young at heart. Science 356, 1035–1039 (2017).

  2. Poss, K. D., Wilson, L. G. & Keating, M. T. Heart regeneration in zebrafish. Science 298, 2188–2190 (2002).

    CAS  Google Scholar 

  3. Porrello, E. R. et al. Transient regenerative potential of the neonatal mouse heart. Science 331, 1078–1080 (2011).

    CAS  Google Scholar 

  4. Jopling, C. et al. Zebrafish heart regeneration occurs by cardiomyocyte dedifferentiation and proliferation. Nature 464, 606–609 (2010).

    CAS  Google Scholar 

  5. Leone, M. & Engel, F. B. Advances in heart regeneration based on cardiomyocyte proliferation and regenerative potential of binucleated cardiomyocytes and polyploidization. Clin. Sci. 133, 1229–1253 (2019).

  6. Itou, J. et al. Migration of cardiomyocytes is essential for heart regeneration in zebrafish. Development 139, 4133–4142 (2012).

    CAS  Google Scholar 

  7. Harvey, R. P., Wystub-Lis, K., del Monte-Nieto, G., Graham, R. M. & Tzahor, E. Cardiac regeneration therapies—targeting Neuregulin 1 signalling. Heart Lung Circ. 25, 4–7 (2016).

    Google Scholar 

  8. D’Uva, G. et al. ERBB2 triggers mammalian heart regeneration by promoting cardiomyocyte dedifferentiation and proliferation. Nat. Cell Biol. 17, 627–638 (2015).

    Google Scholar 

  9. Gemberling, M., Karra, R., Dickson, A. L. & Poss, K. D. Nrg1 is an injury-induced cardiomyocyte mitogen for the endogenous heart regeneration program in zebrafish. eLife 4, e05871 (2015).

    Google Scholar 

  10. Polizzotti, B. D. et al. Neuregulin stimulation of cardiomyocyte regeneration in mice and human myocardium reveals a therapeutic window. Sci. Transl. Med. 7, 281ra45 (2015).

    Google Scholar 

  11. Heallen, T. et al. Hippo pathway deficiency reverses systolic heart failure after infarction. Nature 550, 260–264 (2017).

    Google Scholar 

  12. von Gise, A. et al. YAP1, the nuclear target of Hippo signaling, stimulates heart growth through cardiomyocyte proliferation but not hypertrophy. Proc. Natl Acad. Sci. USA 109, 2394–2399 (2012).

    Google Scholar 

  13. Morikawa, Y. et al. Actin cytoskeletal remodeling with protrusion formation is essential for heart regeneration in Hippo-deficient mice. Sci. Signal. 8, ra41 (2015).

    Google Scholar 

  14. Monroe, T. O. et al. YAP partially reprograms chromatin accessibility to directly induce adult cardiogenesis in vivo. Dev. Cell 48, 765–779 (2019).

    CAS  Google Scholar 

  15. Xin, M. et al. Hippo pathway effector Yap promotes cardiac regeneration. Proc. Natl Acad. Sci. USA 110, 13839–13844 (2013).

    CAS  Google Scholar 

  16. Panciera, T., Azzolin, L., Cordenonsi, M. & Piccolo, S. Mechanobiology of YAP and TAZ in physiology and disease. Nat. Rev. Mol. Cell Biol. 18, 758–770 (2017).

    CAS  Google Scholar 

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

    CAS  Google Scholar 

  18. Ragni, C. V. et al. Amotl1 mediates sequestration of the Hippo effector Yap1 downstream of Fat4 to restrict heart growth. Nat. Commun. 8, 14582 (2017).

    CAS  Google Scholar 

  19. Hirai, M. et al. Adaptor proteins NUMB and NUMBL promote cell cycle withdrawal by targeting ERBB2 for degradation. J. Clin. Invest. 127, 569–582 (2017).

  20. Li, J. et al. Alpha-catenins control cardiomyocyte proliferation by regulating yap activity. Circ. Res. 116, 70–79 (2015).

    CAS  Google Scholar 

  21. Nakada, Y. et al. Hypoxia induces heart regeneration in adult mice. Nature 54, 222–227 (2016).

  22. Marín-juez, R. et al. Fast revascularization of the injured area is essential to support zebrafish heart regeneration. Proc. Natl. Acad. Sci. USA 113, 11237–11242 (2016).

    Google Scholar 

  23. Das, S. et al. A unique collateral artery development program promotes neonatal heart regeneration. Cell 176, 1128–1142 (2019).

    CAS  Google Scholar 

  24. Honkoop, H. et al. Single-cell analysis uncovers that metabolic reprogramming by ErbB2 signaling is essential for cardiomyocyte proliferation in the regenerating heart. eLife 8, e50163 (2019).

  25. Fukuda, R. et al. Metabolic modulation regulates cardiac wall morphogenesis in zebrafish. eLife 8, e50161 (2019).

  26. Lamouille, S., Xu, J. & Derynck, R. Molecular mechanisms of epithelial–mesenchymal transition. Nat. Rev. Mol. Cell Biol. 15, 178–196 (2014).

    CAS  Google Scholar 

  27. 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  Google Scholar 

  28. 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  Google Scholar 

  29. Ni, L., Zheng, Y., Hara, M., Pan, D. & Luo, X. Structural basis for Mob1-dependent activation of the core Mst–Lats kinase cascade in Hippo signaling. Genes Dev. 29, 1416–1431 (2015).

    CAS  Google Scholar 

  30. Moroishi, T. et al. A YAP/TAZ-induced feedback mechanism regulates Hippo pathway homeostasis. Genes Dev. 29, 1271–1284 (2015).

    CAS  Google Scholar 

  31. Hertig, V. et al. Nestin expression is dynamically regulated in cardiomyocytes during embryogenesis. J. Cell. Physiol. 233, 3218–3229 (2018).

    CAS  Google Scholar 

  32. Stroud, M. J., Banerjee, I., Veevers, J. & Chen, J. LINC complex proteins in cardiac structure, function, and disease. Circ. Res. 31, 538–548 (2014).

    Google Scholar 

  33. Kirby, T. J. & Lammerding, J. Emerging views of the nucleus as a cellular mechanosensor. Nat. Cell Biol. 20, 373–381 (2018).

    CAS  Google Scholar 

  34. Swift, J. et al. Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation. Science 341, 965–966 (2013).

    Google Scholar 

  35. Cho, S., Irianto, J. & Discher, D. E. Mechanosensing by the nucleus: from pathways to scaling relationships. J. Cell Biol. 216, 305–315 (2017).

    CAS  Google Scholar 

  36. Torvaldson, E., Kochin, V. & Eriksson, J. E. Phosphorylation of lamins determine their structural properties and signaling functions. Nucleus 6, 166–171 (2015).

    CAS  Google Scholar 

  37. Robison, P. et al. Detyrosinated microtubules buckle and bear load in contracting cardiomyocytes. Science 352, aaf0659 (2016).

    Google Scholar 

  38. Delmar, M. & Liang, F.-X. Connexin43, and the regulation of intercalated disc function. Heart Rhythm 9, 835–838 (2012).

    Google Scholar 

  39. Yang, S. et al. CDK1 phosphorylation of YAP promotes mitotic defects and cell motility and is essential for neoplastic transformation. Cancer Res. 73, 6722–6733 (2013).

    CAS  Google Scholar 

  40. Lin, Z. & Pu, W. T. Releasing YAP from an α-catenin trap increases cardiomyocyte proliferation. Circ. Res. 116, 9–11 (2015).

    CAS  Google Scholar 

  41. Morikawa, Y., Heallen, T., Leach, J., Xiao, Y. & Martin, J. F. Dystrophin–glycoprotein complex sequesters Yap to inhibit cardiomyocyte proliferation. Nature 547, 227–231 (2017).

    CAS  Google Scholar 

  42. Zhang, R. et al. In vivo cardiac reprogramming contributes to zebrafish heart regeneration. Nature 498, 497–501 (2013).

    CAS  Google Scholar 

  43. del Monte-Nieto, G. et al. Control of cardiac jelly dynamics by NOTCH1 and NRG1 defines the building plan for trabeculation. Nature 557, 439–471 (2018).

    Google Scholar 

  44. Von Gise, A. & Pu, W. T. Endocardial and epicardial epithelial to mesenchymal transitions in heart development and disease. Circ. Res. 110, 1628–1645 (2012).

    Google Scholar 

  45. Nieto, M. A., Huang, R. Y. Y. J., Jackson, R. A. A. & Thiery, J. P. P. EMT: 2016. Cell 166, 21–45 (2016).

    CAS  Google Scholar 

  46. Combs, M. D. & Yutzey, K. E. Heart valve developement: regulatory networks in development and disease. Circ. Res. 105, 408–421 (2009).

  47. Gabisonia, K. et al. MicroRNA therapy stimulates uncontrolled cardiac repair after myocardial infarction in pigs. Nature 569, 418–422 (2019).

  48. Aragona, M. et al. A mechanical checkpoint controls multicellular growth through YAP/TAZ regulation by actin-processing factors. Cell 154, 1047–1059 (2013).

    CAS  Google Scholar 

  49. Elosegui-Artola, A. et al. Force triggers YAP nuclear entry by regulating transport across nuclear pores. Cell 171, 1397–1410 (2017).

    CAS  Google Scholar 

  50. Shiu, J. Y., Aires, L., Lin, Z. & Vogel, V. Nanopillar force measurements reveal actin-cap-mediated YAP mechanotransduction. Nat. Cell Biol. 20, 262–270 (2018).

    CAS  Google Scholar 

  51. Komuro, A., Nagai, M., Navin, N. E. & Sudol, M. WW domain-containing protein YAP associates with ErbB-4 and acts as a co-transcriptional activator for the carboxyl-terminal fragment of ErbB-4 that translocates to the nucleus. J. Biol. Chem. 278, 33334–33341 (2003).

    CAS  Google Scholar 

  52. Haskins, J. W., Nguyen, D. X. & Stern, D. F. Neuregulin 1–activated ERBB4 interacts with YAP to induce Hippo pathway target genes and promote cell migration. Sci. Signal. 355, ra116 (2014).

    Google Scholar 

  53. Bui, D. A. et al. Cytokinesis involves a nontranscriptional function of the Hippo pathway effector YAP. Sci. Signal. 9, ra23 (2016).

    Google Scholar 

  54. Zhao, Y. et al. YAP-induced resistance of cancer cells to antitubulin drugs is modulated by a hippo-independent pathway. Cancer Res. 74, 4493–4503 (2014).

    CAS  Google Scholar 

  55. Plouffe, S. W. et al. The Hippo pathway effector proteins YAP and TAZ have both distinct and overlapping functions in the cell. J. Biol. Chem. 293, 11230–11240 (2018).

    CAS  Google Scholar 

  56. Reddy, B. V. V. G. & Irvine, K. D. Regulation of hippo signaling by EGFR–MAPK signaling through ajuba family proteins. Dev. Cell 24, 451–471 (2013).

    Google Scholar 

  57. Hino, N. et al. ERK-mediated mechanochemical waves direct collective cell polarization. Dev. Cell 53, 646–660 (2020).

    CAS  Google Scholar 

  58. Sohal, D. S. et al. Temporally regulated and tissue-specific gene manipulations in the adult and embryonic heart using a tamoxifen-inducible Cre protein. Circ. Res. 89, 20–25 (2001).

    CAS  Google Scholar 

  59. Zhang, N. et al. The Merlin/NF2 tumor suppressor functions through the YAP oncoprotein to regulate tissue homeostasis in mammals. Dev. Cell 19, 27–38 (2010).

    CAS  Google Scholar 

  60. Martin, M. Cutadapt removes adapter sequesnces from high-throughput sequencing reads. EMBnet 17, 5–7 (2011).

    Google Scholar 

  61. Kim, D. et al. TopHat2: accurate alignment of transcriptomes in the presence of insertions, deletions and gene fusions. Genome Biol. 14, 1–13 (2013).

    Google Scholar 

  62. Anders, S. et al. Count-based differential expression analysis of RNA sequencing data using R and Bioconductor. Nat. Protoc. 8, 1765–1786 (2013).

    Google Scholar 

  63. 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).

  64. Wisniewski, J., Zougman, A., Nagaraj, N. & Mann, M. Universal sample preparation method for proteome analysis. Nat. Methods 6, 359–362 (2009).

    CAS  Google Scholar 

  65. Tamary, E. et al. Chlorophyll catabolism precedes changes in chloroplast structure and proteome during leaf senescence. Plant Direct 3, 1–18 (2019).

    CAS  Google Scholar 

  66. Milo-Cochavi, S. et al. The response to the DNA damaging agent methyl methanesulfonate in a fungal plant pathogen. Fungal Biol. 123, 408–422 (2019).

    CAS  Google Scholar 

  67. Shalit, T., Elinger, D., Savidor, A., Gabashvili, A. & Levin, Y. MS1-based label-free proteomics using a quadrupole orbitrap mass spectrometer. J. Proteome Res. 14, 1979–1986 (2015).

    CAS  Google Scholar 

  68. Cox, Jurgen & Mann, M. MaxQuant enables high peptide identification rates, individualized p.p.b.-range mass accuracies and proteome-wide protein quantification. Nat. Biotechnol. 26, 1367–1372 (2008).

    CAS  Google Scholar 

Download references

Acknowledgements

This study has been supported by grants to E.T. from the European Research Council (ERC StG grant no. 281289, CM turnover, and ERC AdG grant no. 788194, CardHeal), ERA-CVD CARDIO-PRO, EU Horizon 2020 research and innovation programme REANIMA, the U.S.–Israel Binational Science Foundation (BSF; to both E.T. and J.F.M.), the Israel Science Foundation (ISF), Foundation Leducq Transatlantic Network of Excellence and Minerva foundation, with funding from the Federal German Ministry for Education and Research. We thank the Benoziyo Endowment Fund for the Advancement of Science, Head of the Yad Abraham Research Center for Cancer Diagnostics and Therapy, Zuckerman STEM Leadership Program, Dr. Dvora and Haim Teitelbaum Endowment Fund and Daniel S. Shapiro Cardiovascular Research Fund. This work was supported by grants from the National Institutes of Health (grant nos HL127717, HL130804 and HL118761 (J.F.M.)), Vivian L. Smith Foundation (J.F.M.) and State of Texas funding (J.F.M.). We thank O. Singer for help with the AAV preparations, G. Friedlander for RNA-seq analysis and input, and N. Priel for teaching ‘trackmate’ software usage.

Author information

Authors and Affiliations

Authors

Contributions

A.A. and E.T. conceived and designed the experiments. A.A., with help from A.Shakked, carried out most of the experiments and analysed the data. Specifically, A.Shakked helped with the animal studies, tissue culture work, RT–qPCR and cloning of YAP mutants into AAV viruses and all associated work and analysis. K.B.U. helped with the animal studies, RNA-seq preparation and analysis. A.Savidor and Y.L. performed the proteomic analysis. A.G. helped with the image analysis. D.K. and D.L. performed the myocardial infarction and echocardiographic analysis. O-Y.R. helped with the functional gelatin-degradation assays and migration time-lapse microscopy. Y.M. helped with the YAP5SA heart sections. J.D. provided custom-made and pYAP S274 antibody. B.G. and J.F.M. contributed to the planning and discussion of the project. E.T. supervised the entire project. A.A. and E.T. wrote the manuscript with contributions from all of the authors.

Corresponding author

Correspondence to Eldad Tzahor.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Peer reviewer reports are available.

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

Extended data

Extended Data Fig. 1 Transient ERBB2 activation in injured hearts (HF model) promotes functional improvement.

a, % Scar quantification for the 3WPMI time point (relates to Fig. 1a, orange arrow) for adult WT (n=5) and OE (n=4) hearts, p=0.180. Statistics were determined by two-tailed t-test. (b-d) Cardiac parameters derived from echocardiographic analysis. Green window represents duration of ERBB2 activation. b, Left ventricular anterior wall (LVAW) thickness, p<0.0001. c, Stroke volume, p=0.0002 (d) Cardiac output, p=0.04, in WT (n=16), OE (n=11) and Sham (n=5) hearts. Statistics, one-way ANOVA followed by Tukey’s multiple comparisons test. Data represent mean ± s.e.m. For statistical source data, see Source Data Extended Data Figure 1.

Source data

Extended Data Fig. 2 EMT hallmarks are upregulated in OE hearts.

a, Analysis of bulk RNA- seq of adult WT–MI (n=4) and OE–MI (n=4) hearts using GSEA hallmark module (relates to Fig. 1h). Bar plot depict the normalized enrichment scores (NES). b-e, RNA- seq derived heat map of the indicated genes belonging to the EMT hallmark (as shown in (a) and Fig. 1h). Genes were annotated as (b) ECM composition (c) ECM modulators (d) Transcription factors and (e) Transmembranal and secreted proteins. FC to the right indicates OE/WT fold change of the detected transcripts by RNA- seq using a threshold FC≥1.5 and p adjusted ≤0.05. Statistics were determined by DESeq2, and adjusted for multiple testing using the procedure of Benjamini and Hochberg. f, Representative image in adult WT and OE hearts. Scale bar, 50 µm.

Extended Data Fig. 3 ERBB2 activation results in reduced CM connectivity and sarcomeric organization, partially depending on YAP.

a, Representative IF images and (b) Representative Haematoxylin & Eosin stain for adult WT, OE and OE-KD heart sections. Scale bar 50µm. c, Quantification of normalized sarcomeric fluorescence intensity for WT (n=3), OE (n=3) and OE-KD (n=3) adult hearts, p=0.047. d, Quantification of normalized interstitial gaps between CMs, for WT (n=3), OE (n=3) and OE-KD (n=3) adult hearts. WT compared to OE, p=0.0018. OE compared to OE-KD, p=0.007. e, Representative IF images of WT (n=3), OE (n=3) and OE-KD (n=3) adult hearts. Scale bar 100µm. Insets highlight separate channels as indicated, white and yellow arrow heads point to connections between CMs with intact and mispatterened CDH2, correspondingly, at the termini of CMs. Inset Scale bar 10µm. Data represent mean ± s.e.m. All figure statistics were determined by one-way ANOVA followed by Tukey’s multiple comparisons test. For statistical source data, see Source Data Extended Data Figure 3.

Source data

Extended Data Fig. 4 YAP Co-IP–MS binding partners.

a, Heat map of YAP enrichment in Yap Co-IP and IgG reactions for WT (n=4) and OE (n=4) adult hearts, as described in Fig. 5a. FC indicates fold change between IP and IgG reactios. b, Heat map of differential YAP binding partners between WT (n=4) and OE (n=4) adult hearts. FC indicates fold change between OE to WT IP reactions of the detected protein by Mass Spec. All displayed results are of statistical significance p ≤ 0.05 between OE and WT IP reactions, which were non-significantly different between OE and WT IgG reactions (similar background). All figure statistics were determined by two-tailed t-test.

Extended Data Fig. 5 Phosphorylation of YAP on S274 and S352 is prevalent in OE hearts and peaks during mitosis.

a, WB analysis of YAP on the soluble fraction of adult WT/OE hearts. b, WB analysis of pYAP S274 on the soluble fraction of adult WT/OE hearts. c, WB analysis of pYAP S352 on the soluble fraction of adult WT/OE hearts. d, WB analysis of YAP on whole cell lysates (includes cytoskeletal components underrepresented in soluble extracts) of adult WT/OE hearts. Blue and red arrow heads point to the usual (“upper”) and below 63kDa (“lower”) bands. e, YAP IP (and IgG) from WT/OE hearts blotted for YAP and pYAP S274. Purple circles represent cut-out bands analysed by Mass-Spec in (f). f, Peptide-spectrum match (#PSM) for YAP analysed by Mass-Spec of (e), an experiment performed once. g, h, Diagram of murine YAP (YAP2L, 488aa) (Plus Myc-DDK tag, as purchased) in (g) and human YAP (YAP2L, 504aa) in (h). Domains indicated above, Hippo phosphorylation sites denoted in grey, and S274/S352 (S289/367 in human) phosphorylation sites circled in red. i, j, Representative images of P7 WT/OE cardiac cultures. Scale bar 50µm. WT and OE insets highlight with an arrow a non-CM, and a CM, at metaphase with peaking pYAP S274/S352 stain. Scale bar 10 µm. k, Quantification of P7 OE Aurkb+ CMs (blue curve) and pYAP S352 mitotic localization (orange curve) at stages of mitosis. Lower panel shows pYAP S352-Aurkb mitotic overlap. n=2274 CMs, from 3 OE hearts. l, Representative image at metaphase of P7 OE cardiac cultures Scale bar 50µm. m, Representative image of embryonic WT E16.5 heart sections. Scale bar 50µm. Inset arrows point to CMs in metaphase. Inset scale bar 10µm. n, Representative image of OE and 5SA adult hearts. White, yellow arrows point to metaphase, pre-metaphase CMs, correspondingly. Scale bar, 10µm. Uncropped blots in panels a, b, c, d, e and numerical source data in panel k are provided in Source Data Extended Data Figure 5.

Source data

Extended Data Fig. 6 ERK interacts with YAP and drives YAP S274 phopshorylation and morphological changes in ERBB2-OE CMs.

a, b, Representative IF images of P7 WT/OE cardiac cultures upon vehicle (DMSO) or ERK-i treatment. Scale bar 50µm. ce, quantification of morphological features of circularity (p), Axial ratio (q), and solidity, (r) upon vehicle (DMSO) or ERK-i (n=127) treatment, of P7 WT (n=123) or OE (n=126) CMs derived from 4 WT and 4 OE P7 hearts. For circularity, WT compared to OE, p=0.0220, OE compared to OE-ERK-i, p=0.0498. For axial ratio, WT compared to OE, p=0.0029, OE compared to OE-ERK-i, p=0.0199. For solidity, WT compared to OE, p=0.0008. OE compared to OE-ERK-i, p=0.0001. f, Quantification of PLA foci in adult cardiac sections, with individual corresponding technical controls for WT (n=6) and OE (n=6) hearts. Data represent mean ± s.e.m. All figure statistics were determined by one-way ANOVA followed by Tukey’s multiple comparisons test. () if P ≤0.05, () if P <0.01, () if P <0.001 and () if P <0.0001. For statistical source data, see Source Data Extended Data Figure 6.

Source data

Supplementary information

Reporting Summary

Peer Review Information

Supplementary Video 1

Time-lapse movie (96 h) of a P7 WT cardiac culture showing CMs tagged with tdTomato fluorescent protein and non-CMs in phase. n = 6 independent experiments.

Supplementary Video 2

Time-lapse movie (96 h) of a P7 OE cardiac culture showing CMs tagged with tdTomato fluorescent protein and non-CMs in phase. n = 4 independent experiments.

Supplementary Table

Table of supplementary information regarding the primer sequences and antibodies used in the study.

Source data

Source Data Fig. 1

Statistical source data of Fig. 1

Source Data Fig. 2

Unprocessed western blots of Fig. 2

Source Data Fig. 2

Statistical source data of Fig. 2

Source Data Fig. 3

Unprocessed western blots of Fig. 3

Source Data Fig. 3

Statistical source data of Fig. 3

Source Data Fig. 4

Unprocessed western blots of Fig. 4

Source Data Fig. 4

Statistical source data of Fig. 4

Source Data Fig. 5

Unprocessed western blots of Fig. 5

Source Data Fig. 5

Statistical source data of Fig. 5

Source Data Fig. 6

Unprocessed western blots of Fig. 6

Source Data Fig. 6

Statistical source data of Fig. 6

Source Data Extended Data Fig. 1

Statistical source data of Extended Data Fig. 1

Source Data Extended Data Fig. 3

Statistical source data of Extended Data Fig. 3

Source Data Extended Data Fig. 5

Unprocessed western blots of Extended Data Fig. 5

Source Data Extended Data Fig. 5

Statistical source data of Extended Data Fig. 5

Source Data Extended Data Fig. 6

Statistical source data of Extended Data Fig. 6

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aharonov, A., Shakked, A., Umansky, K.B. et al. ERBB2 drives YAP activation and EMT-like processes during cardiac regeneration. Nat Cell Biol 22, 1346–1356 (2020). https://doi.org/10.1038/s41556-020-00588-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-020-00588-4

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