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:

Redifferentiated cardiomyocytes retain residual dedifferentiation signatures and are protected against ischemic injury

Abstract

Cardiomyocyte proliferation and dedifferentiation have fueled the field of regenerative cardiology in recent years, whereas the reverse process of redifferentiation remains largely unexplored. Redifferentiation is characterized by the restoration of function lost during dedifferentiation. Previously, we showed that ERBB2-mediated heart regeneration has these two distinct phases: transient dedifferentiation and redifferentiation. Here we survey the temporal transcriptomic and proteomic landscape of dedifferentiation–redifferentiation in adult mouse hearts and reveal that well-characterized dedifferentiation features largely return to normal, although elements of residual dedifferentiation remain, even after the contractile function is restored. These hearts appear rejuvenated and show robust resistance to ischemic injury, even 5 months after redifferentiation initiation. Cardiomyocyte redifferentiation is driven by negative feedback signaling and requires LATS1/2 Hippo pathway activity. Our data reveal the importance of cardiomyocyte redifferentiation in functional restoration during regeneration but also protection against future insult, in what could lead to a potential prophylactic treatment against ischemic heart disease for at-risk patients.

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: Hundreds of genes and proteins show differential expression after redifferentiation.
Fig. 2: Dedifferentiated phenotypes are largely reversed in functionally redifferentiated hearts.
Fig. 3: DR cycle confers robust protection against ischemic injury.
Fig. 4: ERBB2 signaling promotes a multifaceted negative feedback response.
Fig. 5: LATS1/2 negative feedback signaling is required for redifferentiation.

Similar content being viewed by others

Data availability

RNA-seq reads were aligned to the mm10 mouse genome (https://www.ncbi.nlm.nih.gov/data-hub/genome/GCF_000001635.20/). The RNA-seq data discussed in this publication have been deposited in the National Center for Biotechnology Information’s Gene Expression Omnibus (GEO)71 and are accessible through GEO Series accession number GSE220612. The mass spectrometry proteomics data have been deposited to the ProteomeXchange Consortium via the PRIDE74 partner repository with the dataset identifier PXD038450. Data reanalyzed from O’Meara et al.40 (https://doi.org/10.1161/CIRCRESAHA.116.304269) are readily accessible from the online article. Data reanalyzed from Quaife-Ryan et al.41 (https://doi.org/10.1161/CIRCULATIONAHA.117.028252) were downloaded from the GEO database (GSE95755). Source data are provided with this paper.

References

  1. Ali, H., Braga, L. & Giacca, M. Cardiac regeneration and remodelling of the cardiomyocyte cytoarchitecture. FEBS J. 287, 417–438 (2019).

    Article  PubMed  Google Scholar 

  2. 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).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Chen, Y. et al. Reversible reprogramming of cardiomyocytes to a fetal state drives heart regeneration in mice. Science 373, 1537–1540 (2021).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Mahmoud, A. I. et al. Meis1 regulates postnatal cardiomyocyte cell cycle arrest. Nature 497, 249–253 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Heallen, T. et al. Hippo signaling impedes adult heart regeneration. Development 140, 4683–4690 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  10. Lin, Z. et al. Cardiac-specific YAP activation improves cardiac function and survival in an experimental murine MI model. Circ. Res. 115, 354–363 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Wu, Y. et al. LRP6 downregulation promotes cardiomyocyte proliferation and heart regeneration. Cell Res. 31, 450–462 (2021).

  12. Yue, Z. et al. PDGFR-β signaling regulates cardiomyocyte proliferation and myocardial regeneration. Cell Rep. 28, 966–978.e4 (2019).

    Article  CAS  PubMed  Google Scholar 

  13. Leach, J. P. et al. Hippo pathway deficiency reverses systolic heart failure after infarction. Nature 550, 260 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Tao, G. et al. Pitx2 promotes heart repair by activating the antioxidant response after cardiac injury. Nature 534, 119–123 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Xiang, F., Guo, M. & Yutzey, K. E. Overexpression of Tbx20 in adult cardiomyocytes promotes proliferation and improves cardiac function after myocardial infarction. Circulation 133, 1081–1092 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Mohamed, T. M. A. et al. Regulation of cell cycle to stimulate adult cardiomyocyte proliferation and cardiac regeneration. Cell 173, 104–116 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Aharonov, A. et al. ERBB2 drives YAP activation and EMT-like processes during cardiac regeneration. Nat. Cell Biol. 22, 1356–135 (2020).

  18. 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).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Puente, B. N. et al. The oxygen-rich postnatal environment induces cardiomyocyte cell-cycle arrest through DNA damage response. Cell 157, 565–579 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Rodríguez-Sinovas, A., Sánchez, J. A., Valls-Lacalle, L., Consegal, M. & Ferreira-González, I. Connexins in the heart: regulation, function and involvement in cardiac disease. Int. J. Mol. Sci. 22, 4413 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  21. van Rijen, H. V. M. et al. Slow conduction and enhanced anisotropy increase the propensity for ventricular tachyarrhythmias in adult mice with induced deletion of connexin43. Circulation 109, 1048–1055 (2004).

    Article  PubMed  Google Scholar 

  22. Gourdie, R. G. The cardiac gap junction has discrete functions in electrotonic and ephaptic coupling. Anat. Rec. 302, 93–100 (2019).

    Article  CAS  Google Scholar 

  23. Richardson, W. J., Clarke, S. A., Quinn, T. A. & Holmes, J. W. Comprehensive physiology. Compr. Physiol. 5, 1877–1909 (2017).

    Google Scholar 

  24. Liu, X. et al. Lymphoangiocrine signals promote cardiac growth and repair. Nature 588, 705–711 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Garcia-Dorado, D. et al. Gap junction uncoupler heptanol prevents cell-to-cell progression of hypercontracture and limits necrosis during myocardial reperfusion. Circulation 96, 3579–3586 (1997).

    Article  CAS  PubMed  Google Scholar 

  26. Gadicherla, A. K. et al. Mitochondrial Cx43 hemichannels contribute to mitochondrial calcium entry and cell death in the heart. Basic Res. Cardiol. 112, 27 (2017).

    Article  PubMed  Google Scholar 

  27. Wang, N. et al. Selective inhibition of Cx43 hemichannels by Gap19 and its impact on myocardial ischemia/reperfusion injury. Basic Res. Cardiol. 108, 309 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Doumont, G., Martoriati, A. & Marine, J.-C. PTPRV is a key mediator of p53-induced cell cycle exit. Cell Cycle 4, 1703–1705 (2005).

    Article  CAS  PubMed  Google Scholar 

  29. Tian, X., Yang, C., Yang, L., Sun, Q. & Liu, N. PTPRF as a novel tumor suppressor through deactivation of ERK1/2 signaling in gastric adenocarcinoma. OncoTargets Ther. 11, 7795–7803 (2018).

    Article  CAS  Google Scholar 

  30. Yuan, T., Wang, Y., Zhao, Z. J. & Gu, H. Protein-tyrosine phosphatase PTPN9 negatively regulates ErbB2 and epidermal growth factor receptor signaling in breast cancer cells. J. Biol. Chem. 285, 14861–14870 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. van der Lely, L. et al. Loss of PTPN23 promotes proliferation and epithelial-to-mesenchymal transition in human intestinal cancer cells. Inflamm. Intest. Dis. 4, 161–173 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Fearon, A. E. et al. PHLDA1 mediates drug resistance in receptor tyrosine kinase-driven cancer. Cell Rep. 22, 2469–2481 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Kawase, T. et al. PH domain-only protein PHLDA3 is a p53-regulated repressor of Akt. Cell 136, 535–550 (2009).

    Article  CAS  PubMed  Google Scholar 

  34. Habibian, J. S. et al. DUSP5 functions as a feedback regulator of TNFα-induced ERK1/2 dephosphorylation and inflammatory gene expression in adipocytes. Sci Rep. 7, 12879 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Okudela, K. et al. Down-regulation of DUSP6 expression in lung cancer its mechanism and potential role in carcinogenesis. Am. J. Pathol. 175, 867–881 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Zhang, Z. et al. Dual specificity phosphatase 6 (DUSP6) is an ETS-regulated negative feedback mediator of oncogenic ERK signaling in lung cancer cells. Carcinogenesis 31, 577–586 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Li, M. et al. SPRY4-mediated ERK1/2 signaling inhibition abolishes 17β-estradiol-induced cell growth in endometrial adenocarcinoma cell. Gynecol. Endocrinol. 30, 600–604 (2014).

    Article  PubMed  Google Scholar 

  38. 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).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Kumbrink, J., Gerlinger, M. & Johnson, J. P. Egr-1 induces the expression of its corepressor Nab2 by activation of the Nab2 promoter thereby establishing a negative feedback loop. J. Biol. Chem. 280, 42785–42793 (2005).

    Article  CAS  PubMed  Google Scholar 

  40. O’Meara, C. C. et al. Transcriptional reversion of cardiac myocyte fate during mammalian cardiac regeneration. Circ. Res. 116, 804–815 (2015).

    Article  PubMed  Google Scholar 

  41. Quaife-Ryan, G. A. et al. Multicellular transcriptional analysis of mammalian heart regeneration. Circulation 136, 1123–1139 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Lawler, P. R. & Lawler, J. Molecular basis for the regulation of angiogenesis by thrombospondin-1 and -2. Cold Spring Harb. Perspect. Med. 2, a006627 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Mundel, T. M. & Kalluri, R. Type IV collagen-derived angiogenesis inhibitors. Microvasc. Res. 74, 85–89 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  45. Kastan, N. et al. Small-molecule inhibition of Lats kinases may promote Yap-dependent proliferation in postmitotic mammalian tissues. Nat. Commun. 12, 3100 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Heallen, T. et al. Hippo pathway inhibits Wnt signaling to restrain cardiomyocyte proliferation and heart size. Science 332, 458–461 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Vivien, C. J., Hudson, J. E. & Porrello, E. R. Evolution, comparative biology and ontogeny of vertebrate heart regeneration. NPJ Regen. Med. 1, 16012 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Liu, S. et al. Gene therapy knockdown of Hippo signaling induces cardiomyocyte renewal in pigs after myocardial infarction. Sci. Transl. Med. 13, eabd6892 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Abouleisa, R. R. E. et al. Transient cell cycle induction in cardiomyocytes to treat subacute ischemic heart failure. Circulation 145, 1339–1355 (2022).

    Article  CAS  PubMed  Google Scholar 

  50. Nakada, Y. et al. Hypoxia induces heart regeneration in adult mice. Nature 541, 222–227 (2017).

    Article  CAS  PubMed  Google Scholar 

  51. Ocampo, A. et al. In vivo amelioration of age-associated hallmarks by partial reprogramming. Cell 167, 1719–1733 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Zhao, M. et al. Apical resection prolongs the cell cycle activity and promotes myocardial regeneration after left ventricular injury in neonatal pig. Circulation 142, 913–916 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Cardoso, A. C. et al. Mitochondrial substrate utilization regulates cardiomyocyte cell-cycle progression. Nat. Metab. 2, 167–178 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Coburn, C. et al. Anthranilate fluorescence marks a calcium-propagated necrotic wave that promotes organismal death in C. elegans. PLoS Biol. 11, e1001613 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Marsh, S. R., Williams, Z. J., Pridham, K. J. & Gourdie, R. G. Peptidic connexin43 therapeutics in cardiac reparative medicine. J. Cardiovasc. Dev. Dis. 8, 52 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Stokfisz, K., Ledakowicz-Polak, A., Zagorski, M. & Zielinska, M. Ischaemic preconditioning—current knowledge and potential future applications after 30 years of experience. Adv. Med. Sci. 62, 307–316 (2017).

    Article  PubMed  Google Scholar 

  57. Yan, R. et al. An enhancer-based gene-therapy strategy for spatiotemporal control of cargoes during tissue repair. Cell Stem Cell 30, 96–111 (2023).

    Article  CAS  PubMed  Google Scholar 

  58. Scuderi, G. J. & Butcher, J. Naturally engineered maturation of cardiomyocytes. Front. Cell Dev. Biol. 5, 50 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Bogush, N. et al. DUSP5 expression in left ventricular cardiomyocytes of young hearts regulates thyroid hormone (T3)-induced proliferative ERK1/2 signaling. Sci Rep. 10, 21918 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Maillet, M. et al. DUSP6 (MKP3) null mice show enhanced ERK1/2 phosphorylation at baseline and increased myocyte proliferation in the heart affecting disease susceptibility. J. Biol. Chem. 283, 31246–31255 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Missinato, M. A. et al. Dusp6 attenuates Ras/MAPK signaling to limit zebrafish heart regeneration. Development 145, dev157206 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Han, P. et al. Hydrogen peroxide primes heart regeneration with a derepression mechanism. Cell Res. 24, 1091–1107 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Sadek, H. & Olson, E. N. Toward the goal of human heart regeneration. Cell Stem Cell 26, 7–16 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Tzahor, E. & Poss, K. D. Cardiac regeneration strategies: staying young at heart. Science 356, 1035–1039 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Yu, Z., Redfern, C. S. & Fishman, G. I. Conditional transgene expression in the heart. Circ. Res. 79, 691–697 (1996).

    Article  CAS  PubMed  Google Scholar 

  66. Agah, R. et al. Gene recombination in postmitotic cells. Targeted expression of Cre recombinase provokes cardiac-restricted, site-specific rearrangement in adult ventricular muscle in vivo. J. Clin. Invest. 100, 169–179 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10–12 (2011).

    Article  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  69. Anders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).

    Article  CAS  PubMed  Google Scholar 

  70. 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 

  71. Edgar, R., Domrachev, M. & Lash, A. E. Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res. 30, 207–210 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Tamary, E. et al. Chlorophyll catabolism precedes changes in chloroplast structure and proteome during leaf senescence. Plant Direct 3, e00127 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  73. 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).

    Article  CAS  PubMed  Google Scholar 

  74. Perez-Riverol, Y. et al. The PRIDE database resources in 2022: a hub for mass spectrometry-based proteomics evidences. Nucleic Acids Res. 50, D543–D552 (2021).

    Article  PubMed Central  Google Scholar 

  75. Mi, H. et al. PANTHER version 16: a revised family classification, tree-based classification tool, enhancer regions and extensive API. Nucleic Acids Res. 49, gkaa1106 (2020).

    Google Scholar 

  76. Nussinovitch, U. & Gepstein, L. Optogenetics for in vivo cardiac pacing and resynchronization therapies. Nat. Biotechnol. 33, 750–754 (2015).

    Article  CAS  PubMed  Google Scholar 

  77. Lang, D. & Glukhov, A. V. High-resolution optical mapping of the mouse sino-atrial node. J. Vis. Exp. https://doi.org/10.3791/54773 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  78. Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics 26, 139–140 (2010).

    Article  CAS  PubMed  Google Scholar 

  79. Westerfield, M. The Zebrafish Book. A Guide for the Laboratory Use of Zebrafish (Danio rerio), 4th edn (Univ. Oregon Press, 2000).

  80. Aleström, P. et al. Zebrafish: housing and husbandry recommendations. Lab. Anim. 54, 213–224 (2020).

    Article  PubMed  Google Scholar 

  81. Schnabel, K., Wu, C.-C., Kurth, T. & Weidinger, G. Regeneration of cryoinjury induced necrotic heart lesions in zebrafish is associated with epicardial activation and cardiomyocyte proliferation. PLoS ONE 6, e18503 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank the Benoziyo Endowment Fund for the Advancement of Science, the Head of the Yad Abraham Research Center for Cancer Diagnostics and Therapy, the Zuckerman STEM Leadership Program, the Dr. Dvora and Haim Teitelbaum Endowment Fund and the Daniel S. Shapiro Cardiovascular Research Fund. G.F. is the Incumbent of the David and Stacey Cynamon Research Fellow Chair in Genetics and Personalized Medicine. We thank O. Golani for assistance with image processing (Weizmann Institute of Science), M. Oren (Weizmann Institute of Science) and J. Martin (Baylor College of Medicine) for the LATS1/2 cKO mice and A. Genzelinakh (Weizmann Institute of Science) for providing the Sirius Red quantification script. Illustrations in figures were created with BioRender. Funding: European Research Council StG grant 281289 (E.T.); European Research Council AdG grant 788194 (E.T.); ERA-CVD CARDIO-PRO (J.B. and E.T.); U.S.–Israel Binational Science Foundation (E.T.); Israel Science Foundation (E.T.); Foundation Leducq Transatlantic Network of Excellence (E.T.); Minerva Foundation, with funding from the Federal German Ministry for Education and Research (E.T.); EU Horizon 2020 Research and Innovation Programme REANIMA (E.T.); EMBO Long Term Fellowship grant ALTF1129-2015 (P.D.N.); HFSPO Fellowship grant LT001404/2017-L (P.D.N.); NWO-ZonMW Veni grant 016.186.017-3 (P.D.N.); Netherlands Cardiovascular Research Initiative: an initiative with support from the Dutch Heart Foundation and Hartekind, CVON2019-002 OUTREACH (J.B.); and National Institutes of Health grant T32GM007739 (N.K.).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: A. Shakked and E.T. Methodology: A. Shakked, A.A. and K.-B.U. Data curation: G.F., A. Savidor and Y.L. Investigation: A. Shakked, A.A., K.-B.U., Z.P., M.G., P.D.N., D.K., D.L., J.E., Y.D., S.M., L.Z., H.B.-L. and D.E.P. Resources: N.K. Advice, guidance and discussion of results and strategy: R.S. Supervision: E.T., J.B. and L.G. Writing—original draft: A. Shakked and E.T. Writing—review and editing: A. Shakked, E.T., A.A. and S.M.

Corresponding authors

Correspondence to Avraham Shakked or Eldad Tzahor.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cardiovascular Research thanks the anonymous reviewers for their contribution to the peer review of this work.

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 Hundreds of genes and proteins show differential expression after redifferentiation.

ah, Left Ventricular Posterior Diastolic Wall thickness (LVPW;d) (a,f), Left Ventricular Anterior Diastolic Wall thickness (LVAW;d) (b,e), Ejection Fraction (c,g), and long-axis Fractional Shortening (d,h) of WT MI, tOE MI, pOE MI, WT Sham and tOE Sham mice, measured by echocardiography. Gray shaded area represents time period of caERBB2 activation (dedifferentiation). Blue shaded area represents the redifferentiation phase. Overall, n = 3 - 8 per group. For (a), WT MI vs pOE MI p = 0.0002, tOE MI vs pOE MI p = 0.0005, for (b), WT MI vs pOE MI p = 0.0018, tOE MI vs pOE MI p = 0.0461, for (c), WT MI vs pOE MI p = <0.0001, tOE MI vs pOE MI p = 0.0161, WT MI vs tOE MI p = 0.0001, for (d), WT MI vs pOE MI p = <0.0001, tOE MI vs pOE MI p = 0.0003, WT MI vs tOE MI p = 0.0005. i, RT-qPCR analysis of Erbb2 from sham WT and tOE adult heart lysates for each time point. All values are normalised to their in-time point average WT value (black dashed line). All groups had minimum n = 3. For WT vs tOE, Dediff p = 0.004606, Intermediate p = 0.016849. j, Western blot quantification of tOE/WT FC of ERBB2 and pERBB2 (Tyr-1248) protein. All values are normalised to their in-time point average WT value (black dashed line). n = 4 - 8 per group. For WT vs tOE for ERBB2, Dediff p = 0.00189, Intermediate p = 0.00551, for pERBB2, Dediff p = 0.00016. k, Representative western blot images of data from (j). l,m, Heatmaps of differentially expressed genes (l) and proteins (m) from MI-injured samples, compiled as described in Fig. 1g,h. n–q, Principal component analysis and dendrograms of sham RNAseq (n,o), and sham proteomics groups (p,q). In all panels numerical data are presented as mean ± SEM; statistical significance was calculated using one-way ANOVA with Sidak’s multiple comparison test at the 7WPMI time point in (ah), two-tailed unpaired Student’s t-test between tOE and WT of each time point in (i,j). *p ≤ 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001. Uncropped blots for (k) are provided in supplementary source data.

Source data

Extended Data Fig. 2 Dedifferentiated phenotypes are largely reversed in functionally redifferentiated hearts.

a,b, tOE/WT at all timepoints (a) and pOE/WT at Rediff (b) RNA expression fold change of ‘return to normal’ genes, involved in metabolism, proliferation and EMT-like features, determined by RT-qPCR. All values are normalised to their in-time point average WT value (black dashed line). n = 3 - 4 mice per group. c, Western blot quantification for data in Fig. 2b in order to validate the ‘return to normal’ behaviour of proteins involved in proliferation, EMT-like features and metabolism. n = 4 – 7 mice per group. d, Representative immunofluorescence images of isotype controls (WT) for the tOE hearts at Int. and Rediff timepoints shown in Fig. 2c, stained for Ki67, Nestin and Tomm20. Full quantification is provided in Fig. 2c. Scale bars = 50μm for Ki67 and NESTIN, 100μm for TOMM20. e’–f, Metabolic analysis of cultured P7 WT (n = 3) and OE (n = 5) CMs using an XFe96 Seahorse analyser. OCR (oxygen consumption rate) during the Cell Mito Stress Test (e’) and ECAR (extracellular acidification rate (glycolysis proxy)) during the Glycolysis Stress Test (f’). Maximal respiration/OCR (e) and Glycolysis (f). For (e) WT vs OE p = 0.0320. g, H&E-stained histological sections of WT Dediff and tOE Dediff, Int. and Rediff hearts. Images were acquired in the remote zones of MI injured hearts as a proxy for sham injury. Scale bars = 50μm. n = 3 for each group. h,i, Scatter plot of immune-related (h) and angiogenesis-related (i) GO term z-scores against enrichment significance (log10 p-value) for tOE/WT across all timepoints, based on Ingenuity Canonical Pathway Analysis of RNAseq data using a threshold fold change (FC) ≥ 1.5; adjusted p ≤ 0.05. Arrows on each line indicate the direction of the GO term from Dediff to Int. to Rediff. Z-scores below −2 are predictive of pathway inactivation and above +2 are predictive of pathway activation. Values above the horizontal dashed black line represent statistically significant enrichment. j,k, Heatmaps of differentially expressed genes corresponding to the IPA analysis in (i), (j) and an independently curated list of angiogenesis genes (k), compiled as described in Fig. 1g. l, Western blot quantification for data in Fig. 2g in order to validate the proteins involved in metabolism, cytoskeletal signalling and heart function that remain differentially expressed at Rediff. In all panels numerical data are presented as mean ± SEM; statistical significance was calculated using two-tailed unpaired Student’s t-test in (a–c,e,f,l) between the in-time point WT and tOE values. *p ≤ 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001.

Source data

Extended Data Fig. 3 Expanded RNAseq evidence of effects for EMT-like GO terms that ‘return to normal’.

Heatmaps based on log2 transformed normalised counts from Sham RNAseq data of differentially expressed genes that appear within at least one of the listed EMT-like category GO terms (indicated by a tick). Rows represent genes. Columns represent each biological sample. Colour bars represent z-score for each timepoint.

Extended Data Fig. 4 Expanded RNAseq evidence of effects for mitochondrial metabolism GO terms that ‘return to normal’.

Heatmaps based on log2 transformed normalised counts from Sham RNAseq data of differentially expressed genes that appear within at least one of the listed mitochondrial metabolism GO terms (indicated by a tick). Rows represent genes. Columns represent each biological sample. Colour bars represent z-score for all timepoints.

Extended Data Fig. 5 Expanded proteomics evidence of effects for mitochondrial metabolism GO terms that partially ‘return to normal’.

Heat maps based on log10 transformed intensity values from Sham proteomics data of differentially expressed proteins that appear within at least one of the listed mitochondrial metabolism GO terms (indicated by a tick). Rows represent proteins. Columns represent each biological sample. Colour bars represent z-score for each row across all timepoints.

Extended Data Fig. 6 Expanded proteomics evidence of effects for EMT-like category GO terms that partially ‘return to normal’.

Heat maps based on log10 transformed intensity values from Sham proteomics data of differentially expressed proteins that appear within at least one of the listed EMT-like category GO terms (indicated by a tick). Rows represent proteins. Columns represent each biological sample. Colour bars represent z-score for each row across all timepoints.

Extended Data Fig. 7 Expanded proteomics evidence of effects for heart related category GO terms that partially ‘return to normal’.

Heat maps based on log10 transformed intensity values from Sham proteomics data of differentially expressed proteins that appear within at least one of the listed heart related category GO terms (indicated by a tick). Rows represent proteins. Columns represent each biological sample. Colour bars represent z-score for each row across all timepoints.

Extended Data Fig. 8 Dedifferentiation-Redifferentiation cycle confers robust protection against ischaemic injury.

a-d, Relative stroke volume (a), Left Ventricular Anterior Diastolic Wall thickness (LVAW;d) (b), Left Ventricular Posterior Diastolic Wall thickness (LVPW;d) (c) and Left ventricular diastolic volume (LV Volume;d) (d) of WT and tOE-DR, measured by echocardiography. n = 14 for each group. For WT vs tOE-DR in (a), 2DPMI p = 0.000369, 2WPMI p = 0.002282, 4WPMI p = 0.00133. For WT vs tOE-DR in (b), 4WPMI p = 0.010216. e, Scar classification quantification for WT and tOE-DR mice. n = 14 for each group. For WT vs tOE-DR, No Scar p = 0.0442, Transmural scar p = 0.0051. f, Average perimeter and area per aSMA+ vessel. n = 5 for each group. g,h, Average perimeter (g) and area (h) per blood and lymphatic vessel, as measured by CD31 and LYVE1 immunofluorescence. n = 3 for each group. i-l, rSV (i), LVAW;d (j), LVPW;d (k) and LV Volume;d (l) of WT and tOE-DR 5 months after ERBB2 shut-off, measured by echocardiography. n = 21 for WT, n = 16 for tOE-DR. For WT vs tOE-DR in (i), 2WPMI p = 0.013785, 4WPMI p = 0.001556, (j), 2DPMI p = 0.021647. m, Scar classification quantification for WT and tOE-DR mice, 5 months after ERBB2 shut-off. n = 19 for WT, n = 15 for tOE-DR. Data are presented as mean ± SEM; statistical significance was calculated using a one-tailed Mann-Whitney test in (e) and (m) for ‘No scar’, a two-tailed Mann-Whitney test in (m) for transmural scar, and a two-tailed unpaired Student’s t-test in (e) for ‘non-transmural’ and ‘transmural’ scar and (m) for ‘non-transmural’ scar counts and (f to h). *p ≤ 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001

Source data

Extended Data Fig. 9 ERBB2 signalling promotes a multi-faceted negative feedback response.

a, RNAseq FPKM values for negative feedback regulators from ventricular lysate of various ages. Data re-analysed from O’meara et al.40. b, RNAseq normalised counts for negative feedback regulators from bulk-RNAseq of purified CMs from sham injured P1 and P56 mice. Data re-analysed from Quaife-Ryan et al.41. c,d, In situ hybridisation for Hippo pathway genes Lats1, Lats2 and Sav1 (c) and ERK negative feedback regulators Dusp6 & Spry4 (d) mRNA in 7DPI (days post injury) adult zebrafish hearts. RZ = Remote Zone. BZ = Border Zone. Zones are delineated by the black dotted lines. Middle and bottom panels show higher magnification images of the corresponding dashed black boxes in the top panel. Black arrows highlight the presence of detected mRNA. Scale bar in top panels represent 100 µm, scale bars in middle and bottom panels represent 10 µm. n = 3 for each group. e, Quantification of combined FISH and immunofluorescence staining for PCNA and either Dusp6 (top panel) or Spry4 (bottom panel) positive CM nuclei in RZ and BZ of 7DPI adult zebrafish hearts (n = 4). For Dusp6 Border zone vs Remote zone, PCNA + CMs p = 0.0033, Dusp6+ CMs p = 0.0371, PCNA + /Dusp6+ CMs p = 0.0421. For Spry4 Border zone vs Remote zone, PCNA + CMs p = 0.0241, Dusp6+ CMs p = 0.0240, PCNA + /Dusp6+ CMs p = 0.0369. f, Representative images of remote and border zones for Dusp6 (upper) and Spry4 (lower) stained sections in (e). White arrows indicate double-positive CM nuclei. Scale bars represent 20μm. g, To-scale Venn diagram for Dusp6 (upper) and Spry4 (lower) positive nuclei overlapping with PCNA positive nuclei between the RZ and BZ. In all graph panels numerical data are presented as mean ± SEM; statistical significance was calculated using a two-tailed unpaired Student’s t-test in (b) between the P1 and P56 values, and in (e) between the corresponding RZ and BZ values. *p ≤ 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001.

Source data

Extended Data Fig. 10 LATS1/2 negative feedback signalling is required for redifferentiation.

a, Representative immunofluorescence images of Ki67 in LATSi or DMSO treated WT and OE P7 cardiac cultures, with full quantification. All groups had minimum n = 3. Scale bars = 50μm. b,c, Representative western blot of whole-heart lysates for general Yap (gYAP) and pYAP S112 (a target residue of LATS1/2) from WT, tOE, WT LATS1/2 cKO and tOE LATS1/2 cKO mice (b), with quantification, normalised to the average WT value (c). WT n = 3, tOE n = 3, WT cKO n = 4, OE cKO n = 3. In all panels numerical data are presented as mean ± SEM; statistical significance was calculated using a paired two-way ANOVA followed by Sidak’s test in (a) and a one-way ANOVA followed by Tukey’s test in (c). *p ≤ 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001. Uncropped blots for (b) are provided in supplementary source data.

Source data

Supplementary information

Supplementary Information

Captions for Supplementary Movies 1–5 and legends for Extended Data Figs. 1–10.

Reporting Summary

Supplementary Tables 1–5

1. Genotyping primers used for transgenic lines. 2. Primers for RT–qPCR. 3. Antibodies used for western blots. 4. Antibodies used for immunofluorescence. 5. In situ probes.

Functional improvement follows caERBB2 shut-off. Short-axis echocardiography recording of the 4WPMI timepoint (left video) and the 7WPMI timepoint (right video) of the same mouse, demonstrating the functional improvement seen from dedifferentiation to redifferentiation.

DR cycle confers cardioprotection 1 month after ERBB2 shut-off. The first video sequence contains two panels of representative long-axis echocardiography recordings of the same WT mouse at baseline (before injury) and 4 weeks after permanent LAD ligation. The second video sequence is the same format but for a tOE-DR (1 month) mouse. These videos correspond to the data shown in Fig. 3b,c.

DR cycle confers cardioprotection 5 months after ERBB2 shut-off. The first video sequence contains two panels of representative long-axis echocardiography recordings of the same WT mouse at baseline (before injury) and 4 weeks after permanent LAD ligation. The second video sequence is the same format but for a tOE-DR (5 month) mouse. These videos correspond to the data shown in Fig. 3o,p.

Hippo inhibition of ERBB2-overexpressing CMs increases proliferation. Time-lapse microscopy of tdTomato-labeled CMs. Left video shows CMs treated with DMSO, and right video shows CMs from the same mouse treated with TRULI (corresponding to the 2nd and 4th panel of Fig. 5b, respectively). Mitosis counter reflects the number of mitotic events that occurred in the field of view as the video progresses. White arrows appear shortly before a mitotic event. Most, but not all, mitotic events resulted in cytokinesis. Only cytokinetic events were counted during the quantification shown in Fig. 5b. Time counter in top-left corner of each video represents time elapsed (hh:mm).

LATS1/2 cKO in CMs after transient ERBB2 overexpression prevents redifferentiation. From left to right, video panels show short-axis echocardiography recordings of mice at the 7WPI timepoint (see Fig. 5d) for WT, tOE and tOE-LATS1/2 cKO mice to highlight the failure to redifferentiate in the absence of LATS1/2. Each panel is accompanied (below) by a representative immunofluorescent stain of the heart for cardiac troponin T (green) and DAPI (blue) (also shown in Fig. 5g), to reflect the gross loss of sarcomeric integrity that accompanies tOE LATS1/2 cKO hearts. R-diff, redifferentiation.

Supplementary Code

Supplementary Data 1

Supplementary Data 2

Supplementary Data 3

Supplementary Data 4

Supplementary Data 5

Supplementary Data 6

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. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Source Data Fig. 2

Unprocessed western blot.

Source Data Fig. 4

Unprocessed western blot.

Source Data Extended Data Fig. 1

Unprocessed western blot.

Source Data Extended Data Fig. 10

Unprocessed western blot.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shakked, A., Petrover, Z., Aharonov, A. et al. Redifferentiated cardiomyocytes retain residual dedifferentiation signatures and are protected against ischemic injury. Nat Cardiovasc Res 2, 383–398 (2023). https://doi.org/10.1038/s44161-023-00250-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s44161-023-00250-w

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