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:

Human embryonic stem cell–derived cardiomyocytes restore function in infarcted hearts of non-human primates

An Erratum to this article was published on 01 October 2018

This article has been updated

Abstract

Pluripotent stem cell–derived cardiomyocyte grafts can remuscularize substantial amounts of infarcted myocardium and beat in synchrony with the heart, but in some settings cause ventricular arrhythmias. It is unknown whether human cardiomyocytes can restore cardiac function in a physiologically relevant large animal model. Here we show that transplantation of 750 million cryopreserved human embryonic stem cell–derived cardiomyocytes (hESC-CMs) enhances cardiac function in macaque monkeys with large myocardial infarctions. One month after hESC-CM transplantation, global left ventricular ejection fraction improved 10.6 ± 0.9% vs. 2.5 ± 0.8% in controls, and by 3 months there was an additional 12.4% improvement in treated vs. a 3.5% decline in controls. Grafts averaged 11.6% of infarct size, formed electromechanical junctions with the host heart, and by 3 months contained 99% ventricular myocytes. A subset of animals experienced graft-associated ventricular arrhythmias, shown by electrical mapping to originate from a point-source acting as an ectopic pacemaker. Our data demonstrate that remuscularization of the infarcted macaque heart with human myocardium provides durable improvement in left ventricular function.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Effects of hESC-CM transplantation on cardiac function.
Figure 2: Analysis of arrhythmias.
Figure 3: Structural assessment of infarct, graft size, and graft composition.
Figure 4: Graft maturation, integration, vascularization, and proliferation.
Figure 5: Visualization of grafts by MRI.

Similar content being viewed by others

Zixuan Zhao, Xinyi Chen, … Hanry Yu

Change history

  • 18 July 2018

    In the version of this article initially published, NIH grant P51 OD010425 was omitted. The error has been corrected in the HTML and PDF versions of the article.

  • 01 October 2018

    Nat. Biotechnol. 36, 597–605 (2018); published online 2 July 2018; corrected after print 18 July 2018 In the version of this article initially published, NIH grant P51 OD010425 was omitted. The error has been corrected in the HTML and PDF versions of the article.

References

  1. Forouzanfar, M.H. et al. Assessing the global burden of ischemic heart disease, part 2: analytic methods and estimates of the global epidemiology of ischemic heart disease in 2010. Glob. Heart 7, 331–342 (2012).

    Article  PubMed  Google Scholar 

  2. Laflamme, M.A. & Murry, C.E. Heart regeneration. Nature 473, 326–335 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Eschenhagen, T. et al. Cardiomyocyte regeneration: a consensus statement. Circulation 136, 680–686 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  4. Bergmann, O. et al. Evidence for cardiomyocyte renewal in humans. Science 324, 98–102 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Soonpaa, M.H. & Field, L.J. Survey of studies examining mammalian cardiomyocyte DNA synthesis. Circ. Res. 83, 15–26 (1998).

    Article  CAS  PubMed  Google Scholar 

  6. van Berlo, J.H. et al. c-kit+ cells minimally contribute cardiomyocytes to the heart. Nature 509, 337–341 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Burridge, P.W., Sharma, A. & Wu, J.C. Genetic and epigenetic regulation of human cardiac reprogramming and differentiation in regenerative medicine. Annu. Rev. Genet. 49, 461–484 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Gerbin, K.A. & Murry, C.E. The winding road to regenerating the human heart. Cardiovasc. Pathol. 24, 133–140 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  9. Lee, R.T. & Walsh, K. The future of cardiovascular regenerative medicine. Circulation 133, 2618–2625 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Hassink, R.J. et al. Cardiomyocyte cell cycle activation improves cardiac function after myocardial infarction. Cardiovasc. Res. 78, 18–25 (2008).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. 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 

  14. Shapiro, S.D. et al. Cyclin A2 induces cardiac regeneration after myocardial infarction through cytokinesis of adult cardiomyocytes. Sci. Transl. Med. 6, 224ra27 (2014).

    PubMed  Google Scholar 

  15. Leach, J.P. et al. Hippo pathway deficiency reverses systolic heart failure after infarction. Nature 550, 260–264 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Liu, Z. et al. Single-cell transcriptomics reconstructs fate conversion from fibroblast to cardiomyocyte. Nature 551, 100–104 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  17. Zhou, H. et al. ZNF281 enhances cardiac reprogramming by modulating cardiac and inflammatory gene expression. Genes Dev. 31, 1770–1783 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Cao, N. et al. Conversion of human fibroblasts into functional cardiomyocytes by small molecules. Science 352, 1216–1220 (2016).

    Article  CAS  PubMed  Google Scholar 

  19. Kadota, S., Pabon, L., Reinecke, H. & Murry, C.E. In vivo maturation of human induced pluripotent stem cell-derived cardiomyocytes in neonatal and adult rat hearts. Stem Cell Reports 8, 278–289 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Cho, G.S. et al. Neonatal transplantation confers maturation of PSC-derived cardiomyocytes conducive to modeling cardiomyopathy. Cell. Rep. 18, 571–582 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. van Laake, L.W. et al. Human embryonic stem cell-derived cardiomyocytes survive and mature in the mouse heart and transiently improve function after myocardial infarction. Stem Cell Res. (Amst.) 1, 9–24 (2007).

    Article  Google Scholar 

  22. Laflamme, M.A. et al. Cardiomyocytes derived from human embryonic stem cells in pro-survival factors enhance function of infarcted rat hearts. Nat. Biotechnol. 25, 1015–1024 (2007).

    Article  CAS  PubMed  Google Scholar 

  23. Caspi, O. et al. Transplantation of human embryonic stem cell-derived cardiomyocytes improves myocardial performance in infarcted rat hearts. J. Am. Coll. Cardiol. 50, 1884–1893 (2007).

    Article  PubMed  Google Scholar 

  24. Shiba, Y. et al. Human ES-cell-derived cardiomyocytes electrically couple and suppress arrhythmias in injured hearts. Nature 489, 322–325 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Chong, J.J. et al. Human embryonic-stem-cell-derived cardiomyocytes regenerate non-human primate hearts. Nature 510, 273–277 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Shiba, Y. et al. Allogeneic transplantation of iPS cell-derived cardiomyocytes regenerates primate hearts. Nature 538, 388–391 (2016).

    Article  CAS  PubMed  Google Scholar 

  27. Naumova, A.V., Modo, M., Moore, A., Murry, C.E. & Frank, J.A. Clinical imaging in regenerative medicine. Nat. Biotechnol. 32, 804–818 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhang, W.Y., Ebert, A.D., Narula, J. & Wu, J.C. Imaging cardiac stem cell therapy: translations to human clinical studies. J. Cardiovasc. Transl. Res. 4, 514–522 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  29. Chen, V.C. et al. Development of a scalable suspension culture for cardiac differentiation from human pluripotent stem cells. Stem Cell Res. (Amst.) 15, 365–375 (2015).

    Article  CAS  Google Scholar 

  30. Hohnloser, S.H., Klingenheben, T. & Singh, B.N. Amiodarone-associated proarrhythmic effects. A review with special reference to torsade de pointes tachycardia. Ann. Intern. Med. 121, 529–535 (1994).

    Article  CAS  PubMed  Google Scholar 

  31. Chelsky, L.B. et al. Caffeine and ventricular arrhythmias. An electrophysiological approach. J. Am. Med. Assoc. 264, 2236–2240 (1990).

    Article  CAS  Google Scholar 

  32. Nussbaum, J. et al. Transplantation of undifferentiated murine embryonic stem cells in the heart: teratoma formation and immune response. FASEB J. 21, 1345–1357 (2007).

    Article  CAS  PubMed  Google Scholar 

  33. Kuroda, N., Tanaka, A., Ohe, C. & Nagashima, Y. Recent advances of immunohistochemistry for diagnosis of renal tumors. Pathol. Int. 63, 381–390 (2013).

    Article  CAS  PubMed  Google Scholar 

  34. Bedada, F.B. et al. Acquisition of a quantitative, stoichiometrically conserved ratiometric marker of maturation status in stem cell-derived cardiac myocytes. Stem Cell Reports 3, 594–605 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Bergmann, O. et al. Identification of cardiomyocyte nuclei and assessment of ploidy for the analysis of cell turnover. Exp. Cell Res. 317, 188–194 (2011).

    Article  CAS  PubMed  Google Scholar 

  36. Swijnenburg, R.J. et al. Embryonic stem cell immunogenicity increases upon differentiation after transplantation into ischemic myocardium. Circulation 112 Suppl, I166–I172 (2005).

    PubMed  Google Scholar 

  37. Lee, A.S., Tang, C., Rao, M.S., Weissman, I.L. & Wu, J.C. Tumorigenicity as a clinical hurdle for pluripotent stem cell therapies. Nat. Med. 19, 998–1004 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Hodgkinson, C.P., Bareja, A., Gomez, J.A. & Dzau, V.J. Emerging concepts in paracrine mechanisms in regenerative cardiovascular medicine and biology. Circ. Res. 118, 95–107 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Gerbin, K.A., Yang, X., Murry, C.E. & Coulombe, K.L. Enhanced electrical integration of engineered human myocardium via intramyocardial versus epicardial delivery in infarcted rat hearts. PLoS One 10, e0131446 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Zhu, K. et al. Lack of remuscularization following transplantation of human embryonic stem cell-derived cardiovascular progenitor cells in infarcted nonhuman primates. Circ. Res. 122, 958–969 (2018).

    Article  CAS  PubMed  Google Scholar 

  41. Tachibana, A. et al. Paracrine effects of the pluripotent stem cell-derived cardiac myocytes salvage the injured myocardium. Circ. Res. 121, e22–e36 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Liao, S.Y. et al. Overexpression of Kir2.1 channel in embryonic stem cell-derived cardiomyocytes attenuates posttransplantation proarrhythmic risk in myocardial infarction. Heart Rhythm 10, 273–282 (2013).

    Article  PubMed  Google Scholar 

  43. Liao, S.Y. et al. Proarrhythmic risk of embryonic stem cell-derived cardiomyocyte transplantation in infarcted myocardium. Heart Rhythm 7, 1852–1859 (2010).

    Article  PubMed  Google Scholar 

  44. Roell, W. et al. Engraftment of connexin 43-expressing cells prevents post-infarct arrhythmia. Nature 450, 819–824 (2007).

    Article  CAS  PubMed  Google Scholar 

  45. Chong, J.J. & Murry, C.E. Cardiac regeneration using pluripotent stem cells—progression to large animal models. Stem Cell Res. (Amst.) 13, 654–665 (2014).

    Article  CAS  Google Scholar 

  46. Fernandes, S. et al. Human embryonic stem cell-derived cardiomyocytes engraft but do not alter cardiac remodeling after chronic infarction in rats. J. Mol. Cell. Cardiol. 49, 941–949 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

These studies were supported in part by NIH Grants R01HL128362, R01 HL084642, P01HL094374, and a grant from the Fondation Leducq Transatlantic Network of Excellence (all to C.E.M.), and grant P51 OD010425 from the NIH Office of Research Infrastructure Programs to the Washington National Primate Research Center. These studies also were supported by the UW Medicine Heart Regeneration Program, the Washington Research Foundation, and a gift from Mike and Lynn Garvey. The manufacture of cells provided by City of Hope was funded in part through the National Heart Lung and Blood Institute's Production Assistance for Cell Therapies (PACT). We also acknowledge the support of the Cell Analysis Facility Flow Cytometry and Imaging Core in the Department of Immunology at the University of Washington. We thank the Garvey Imaging Core for assistance with microscopy. We are indebted to the dedicated staff of the Washington National Primate Research Center for supporting many aspects of this study. We thank M. Laflamme for helpful discussions, W.-Z. Zhu for support with electrophysiological studies, and P. Swanson for consultation on the renal tumor. We thank J. Maki and G. Wilson for help with cardiac MRI protocol.

Author information

Authors and Affiliations

Authors

Contributions

Y.-W.L. designed experiments, conducted animal studies, analyzed data and edited the manuscript. B.C. designed experiments, conducted animal studies, performed histology studies, analyzed data and edited the manuscript. X.Y. performed histology, histomorphometry, analyzed data, prepared figures, and edited the manuscript. J.A.F. prepared cells for transplantation and edited the manuscript. F.A.K. prepared cells for transplantation and edited the manuscript. A.F.-T. prepared cells for transplantation and edited the manuscript. L.C. supervised cell manufacturing. K.W.V. designed experiments, conducted animal studies, and provided veterinary care. C.A.A. designed experiments and conducted animal studies. A.B. performed necropsies and provided pathology consultation. J.O. conducted animal studies and edited the manuscript. C.W.D. designed experiments, conducted animal studies and edited the manuscript. Z.L.S. conducted animal studies and edited the manuscript. S.P.S. designed experiments, conducted electrophysiology studies, analyzed data and edited the manuscript. S.A.T. conducted animal studies, analyzed data and edited the manuscript. H.T. conducted animal studies and edited the manuscript. A.V.N. performed MRI studies, analyzed data and edited the manuscript. S.K.D. performed histology. M.S.L. analyzed MRI scans. J.L. analyzed MRI scans and edited the manuscript. D.W.H. advised on and performed microscopy. H.R. designed experiments, performed histology, conducted molecular analyses, analyzed data, and prepared figures. L.P. designed experiments, analyzed data and edited the manuscript. B.H.F. prepared cells for transplantation. WR.M. designed experiments, analyzed data and edited the manuscript. R.S.T. designed experiments, conducted animal studies, oversaw preparation of cells for transplantation, analyzed data, prepared figures and edited the manuscript. C.E.M. supervised all components of this study, designed experiments, performed animal experiments, analyzed data, obtained funding for the study, prepared figures and wrote the manuscript.

Corresponding author

Correspondence to Charles E Murry.

Ethics declarations

Competing interests

C.E.M., R.S.T., and W.R.M. are scientific founders and equity holders in Cytocardia.

Integrated supplementary information

Supplementary Figure 1 Flow cytometry for assessment of cardiomyocyte purity.

(a) Forward and side scatter with main population of cells indicated. (b) Isotype control antibody showing minimal staining in M2 gate. (c) cTnT antibody showing 98.7% of cells expressing this epitope. (d) Overlay of isotype control and cTnT antibodies showing clear spectral separation of the two preparations. This experiment was repeated in 4 biologically independent cell preparation runs, with purities ranging from 86-99% cTnT-positive. PE, phycoerythrin.

Supplementary Figure 2 Flow diagram for animal assignment and inclusion in MRI and electrophysiology (EP) studies.

This figure depicts the experimental flow, mortality at each stage of the study, assignment into various experimental protocols, exclusion for technical limitations and final group sizes for MRI and electrophysiological (EP) analyses.

Supplementary Figure 3 Inter-observer variance in MRI and EKG telemetry.

(a) Regression plot for determination of LVEF by two MRI observers blinded to treatment. There is a high correlation coefficient (R2 = 0.92; p < 0.001) with a slope close to 1.0. (b) Bland and Altman analysis showing that there was no difference of LVEF from observer 1 (Y axis) and observer 2 (X axis) (p = 0.94). (c) Regression plot for quantification of ventricular arrhythmias by two observers. There is a high correlation coefficient (R2=0.98) with a slope close to 1.0. (d) Bland and Altman analysis showing that there was no difference for ventricular arrhythmia duration between the two observers.

Supplementary Figure 4 Regional contractile function (percent LV wall systolic thickening) in the non-infarcted wall at baseline and at 4 weeks post-treatment.

There was no difference between groups at any time or within groups at different times. Each point represents one heart. Data were obtained from 4 biologically independent control animals and 5 biologically independent hESC-CM treated animals. Group data represent means ± SEM.

Supplementary Figure 5 GCaMP3 transgene and amiodarone are not necessary for arrhythmias.

The small infarct protocol (90 min occlusion of distal LAD) was used here. This model demonstrates minimal arrhythmias before transplantation and consistent arrhythmias after transplantation. Control group (not shown) had minimal arrhythmias throughout. (a) Omission of GCaMP transgene in engrafted hESC-CM did not diminish the ventricular arrhythmias. Data are from 4 biologically independent GCaMP transgenic hESC-CM animals and 1 wild-type hESC-CM treated animal. (b) Withholding amiodarone did not diminish the ventricular arrhythmias. Data are from 4 biologically independent animals receiving hESC-CM in the presence of amiodarone and from 1 hESC-CM treated animal that did not receive amiodarone.

Supplementary Figure 6 Analysis of tumor origin identified in one macaque kidney.

All data in this figure were obtained from one hESC-CM treated animal that developed a renal tumor. (a) Hematoxylin and eosin stained section of kidney containing carcinoma cells within a lymphatic vessel. A glomerulus (Glom) is visible to the right of the tumor. Scale bar, 50 μm. (b) High magnification image of the region boxed in (a), shows that the carcinoma cells have formed tubules and have eosinophilic proteinaceous material in the lumens. Scale bar, 50 μm. (c) PCR genotyping of tumor. DNA extracted from a histological slide containing kidney tissue and tumor was analyzed by PCR using primers specific for human mitochondria (high copy number per cell) and macaque chromosome 3. Human mitochondrial primers generated an amplicon in purified human DNA but not in the kidney tumor sample or monkey DNA sample. Macaque chromosome 3 primers gave amplicons in the monkey and kidney tumor samples, but not in purified human DNA. (d) In situ hybridization positive control sample of hESC-CM graft in macaque heart. The human graft nuclei hybridize intensely with the probe (red), whereas host nuclei are negative. Individual channels are shown below. DAPI counterstain. Scale bar, 50 μm. (e) Negative control for in situ hybridization study on an adjacent section of the tumor where the human-specific probe was omitted. The proteinaceous material in the glandular lumens is intensely autofluorescent (red), but the tumor nuclei are negative. Scale bar, 50 μm. (f) In situ hybridization study of tumor on adjacent section shows autofluorescent proteinacous material in the glandular lumens (red), but the nuclei are unstained. This indicates the tumor is not of human origin. Scale bar, 50 μm.

Supplementary Figure 7 Infarct shrinkage from baseline to 1 month post-treatment.

Infarct size was determined serially in 8 macaques by MRI using delayed Gd enhancement. Both groups show infarct shrinkage, but significantly greater scar contraction occurred in hESC-CM-treated hearts. Data are from 5 biologically independent hESC-CM treated and 3 control animals.

Supplementary Figure 8 Validation of human-specific monoclonal antibodies to cardiac troponin I (cTnI) and slow skeletal troponin I (ssTnI).

(a) Archival section from macaque heart containing an hESC-CM graft that expresses GFP. The graft region is readily identifiable by the brown immunoperoxidase reaction product using anti-GFP immunostaining. (b) Adjacent section from the same tissue block stained by DNA in situ hybridization for human-specific pan-centromeric sequences (red). The human graft is readily identifiable in the background of the primate heart and shows good spatial correspondence with the anti-GFP staining shown in (a). (c) Adjacent section stained with human cTnI antibody (brown) shows graft region that corresponds well with anti-GFP and human pan-centromeric probe in situ hybridization. (d) Adjacent section stained with ssTnI antibody (brown) shows graft region that corresponds well with anti-GFP, human pan-centromeric probe in situ hybridization and human cTnI immunostaining. Note that, although this antibody also labels monkey ssTnI, this isoform is not expressed by the adult monkey heart. These validation studies were performed on a single hESC-CM treated heart.

Supplementary Figure 9 Low magnification whole-slide scans from 4 hearts receiving hESC-CM grafts, stained for human cTnI (brown).

Apical section is on the left and basal section is on the right. Extensive areas of remuscularization are present in all hearts. These experiments were performed on 4 biologically independent hESC-CM treated hearts, and data from each heart are shown here. Scale bar, 5 mm.

Supplementary Figure 10 Maturation of adherens and gap junctions in hESC-CM grafts.

(a) At 1 month post-engraftment, Connexin43+ gap junctions (green) are punctate and often circumferentially distributed around the graft cardiomyocytes (ssTnI, red). This experiment was repeated in 2 biologically independent hESC-CM treated hearts with similar results. (b) At 3 months post-engraftment, many Connexin43+ gap junctions have polarized to intercalated disks, although in other locations (not shown) the staining was still circumferential. This experiment was repeated in 2 biologically independent hESC-CM treated hearts with similar results. (c) At 1 month post-engraftment, pan-cadherin+ adherens junctions (green) are punctate and distributed around the graft cardiomyocytes (ssTnI, red). This experiment was repeated in 2 biologically independent hESC-CM treated hearts with similar results. (d) At 3 months post-engraftment, cadherin+ adherens junctions have polarized to the intercalated disks of graft cardiomyocytes. This experiment was repeated in 2 biologically independent hESC-CM treated hearts with similar results. Scale bar a-d, 25 μm.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–10 (PDF 2321 kb)

Life Sciences Reporting Summary (PDF 310 kb)

Supplementary Tables

Supplementary tables 1–4 (PDF 237 kb)

Supplementary Notes

Supplementary Notes (PDF 175 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, YW., Chen, B., Yang, X. et al. Human embryonic stem cell–derived cardiomyocytes restore function in infarcted hearts of non-human primates. Nat Biotechnol 36, 597–605 (2018). https://doi.org/10.1038/nbt.4162

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nbt.4162

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