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:

Primitive macrophages induce sarcomeric maturation and functional enhancement of developing human cardiac microtissues via efferocytic pathways

Abstract

Yolk sac macrophages are the first to seed the developing heart; however, owing to a lack of accessible tissue, there is no understanding of their roles in human heart development and function. In this study, we bridge this gap by differentiating human embryonic stem (hES) cells into primitive LYVE1+ macrophages (hESC-macrophages) that stably engraft within contractile cardiac microtissues composed of hESC-cardiomyocytes and fibroblasts. Engraftment induces a human fetal cardiac macrophage gene program enriched in efferocytic pathways. Functionally, hESC-macrophages trigger cardiomyocyte sarcomeric protein maturation, enhance contractile force and improve relaxation kinetics. Mechanistically, hESC-macrophages engage in phosphatidylserine-dependent ingestion of apoptotic cardiomyocyte cargo, which reduces microtissue stress, leading hESC-cardiomyocytes to more closely resemble early human fetal ventricular cardiomyocytes, both transcriptionally and metabolically. Inhibiting hESC-macrophage efferocytosis impairs sarcomeric protein maturation and reduces cardiac microtissue function. Together, macrophage-engineered human cardiac microtissues represent a considerably improved model for human heart development and reveal a major beneficial role for human primitive macrophages in enhancing early cardiac tissue function.

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: hESC-macrophages undergo stepwise in vivo yolk sac tissue macrophage programming in bioengineered human cardiac microtissues.
Fig. 2: hESC-macrophages enhance electromechanical properties of human cardiac microtissues across cell sources and platforms.
Fig. 3: hESC-macrophages induce maturation of cardiomyocyte sarcomeric apparatus and alter ECM composition in human cardiac microtissues.
Fig. 4: Multidimensional CyTOF analyses reveal that hESC-macrophages reduce mitochondrial debris, increase ATP production and realign cardiomyocyte metabolism to approximate early fetal ventricular cardiomyocytes.
Fig. 5: hESC-macrophage incorporation into cardiac microtissues results in reduced cytotoxicity and increased differentiation of cardiomyocytes toward in vivo human fetal cardiomyocytes.
Fig. 6: Efferocytic uptake of cardiomyocytes by hESC-macrophages is driven by phosphatidylserine recognition and is required for full macrophage maturation.
Fig. 7: Inhibiting hESC-macrophage efferocytosis of cardiomyocytes leads to blunted cardiomyocyte maturation, increased cytotoxicity and impaired microtissue function.
Fig. 8: Graphical summary of findings.

Similar content being viewed by others

Data availability

RNA sequencing data generated in this study were deposited to the Gene Expression Omnibus under accession number GSE261628. Proteomics data generated in this study were deposited to the ProteomeXchange with the identifiers PXD050990 (control versus macrophage dataset) and PXD050996 (efferocytosis inhibition with annexin dataset). Transcriptomics and proteomics data are available to explore through an interactive web browser (Shiny app) at https://www.epelmanlab.com/resources. Accession codes and sample information for publicly available single-cell RNA sequencing data analyzed in this study are summarized in Supplementary Table 9. All data generated in this study are included in the main article and associated source files.

Code availability

The custom-made MATLAB code for quantitative image analysis of cardiac structural properties are described in ref. 95. Scripts for the analysis of bulk RNA sequencing and single-cell RNA sequencing data are available at https://github.com/HomairaH/EpelmanLab_Human-cardiac-microtissues.

References

  1. Dick, S. A. et al. Self-renewing resident cardiac macrophages limit adverse remodeling following myocardial infarction. Nat. Immunol. 20, 29–39 (2019).

    Article  CAS  PubMed  Google Scholar 

  2. Dick, S. A. et al. Three tissue resident macrophage subsets coexist across organs with conserved origins and life cycles. Sci. Immunol. 7, eabf7777 (2022).

    Article  CAS  PubMed  Google Scholar 

  3. Epelman, S. et al. Embryonic and adult-derived resident cardiac macrophages are maintained through distinct mechanisms at steady state and during inflammation. Immunity 40, 91–104 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Epelman, S., Lavine, K. J. & Randolph, G. J. Origin and functions of tissue macrophages. Immunity 41, 21–35 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Bajpai, G. et al. The human heart contains distinct macrophage subsets with divergent origins and functions. Nat. Med. 24, 1234–1245 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Yona, S. et al. Fate mapping reveals origins and dynamics of monocytes and tissue macrophages under homeostasis. Immunity 38, 79–91 (2013).

    Article  CAS  PubMed  Google Scholar 

  7. Ginhoux, F. & Guilliams, M. Tissue-resident macrophage ontogeny and homeostasis. Immunity 44, 439–449 (2016).

    Article  CAS  PubMed  Google Scholar 

  8. Bian, Z. et al. Deciphering human macrophage development at single-cell resolution. Nature 582, 571–576 (2020).

    Article  CAS  PubMed  Google Scholar 

  9. Popescu, D.-M. et al. Decoding human fetal liver haematopoiesis. Nature 574, 365–371 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Hulsmans, M. et al. Macrophages facilitate electrical conduction in the heart. Cell 169, 510–522 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Lavine, K. et al. Distinct macrophage lineages contribute to disparate patterns of cardiac recovery and remodeling in the neonatal and adult heart. Proc. Natl Acad. Sci. USA 111, 16029–16034 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Aurora, A. B. et al. Macrophages are required for neonatal heart regeneration. J. Clin. Invest. 124, 1382–1392 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Bajpai, G. et al. Tissue resident CCR2 and CCR2+ cardiac macrophages differentially orchestrate monocyte recruitment and fate specification following myocardial injury. Circ. Res. 124, 263–278 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Epelman, S., Liu, P. P. & Mann, D. L. Role of innate and adaptive immune mechanisms in cardiac injury and repair. Nat. Rev. Immunol. 15, 117–129 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Epelman, S. & Mann, D. L. Communication in the heart: the role of the innate immune system in coordinating cellular responses to ischemic injury. J. Cardiovasc. Transl. Res. 5, 827–836 (2012).

    Article  PubMed  Google Scholar 

  16. Wong, A., Hamidzada, H. & Epelman, S. A cardioimmunologist’s toolkit: genetic tools to dissect immune cells in cardiac disease. Nat. Rev. Cardiol. 19, 395–413 (2022).

    Article  PubMed  Google Scholar 

  17. Wong, N. R. et al. Resident cardiac macrophages mediate adaptive myocardial remodeling. Immunity 54, 2072–2088 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zaman, R., Hamidzada, H. & Epelman, S. Exploring cardiac macrophage heterogeneity in the healthy and diseased myocardium. Curr. Opin. Immunol. 68, 54–63 (2020).

    Article  PubMed  Google Scholar 

  19. Zaman, R. et al. Selective loss of resident macrophage-derived insulin-like growth factor-1 abolishes adaptive cardiac growth to stress. Immunity 54, 2057–2071 (2021).

    Article  CAS  PubMed  Google Scholar 

  20. Wan, E. et al. Enhanced efferocytosis of apoptotic cardiomyocytes through myeloid-epithelial-reproductive tyrosine kinase links acute inflammation resolution to cardiac repair after infarction. Circ. Res. 113, 1004–1012 (2013).

    Article  CAS  PubMed  Google Scholar 

  21. Glinton, K. E. et al. Macrophage-produced VEGFC is induced by efferocytosis to ameliorate cardiac injury and inflammation. J. Clin. Invest. 132, e140685 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Morioka, S., Maueröder, C. & Ravichandran, K. S. Living on the edge: efferocytosis at the interface of homeostasis and pathology. Immunity 50, 1149–1162 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Valentin, J. E., Stewart-Akers, A. M., Gilbert, T. W. & Badylak, S. F. Macrophage participation in the degradation and remodeling of extracellular matrix scaffolds. Tissue Eng. Part A 15, 1687–1694 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Doran, A. C., Yurdagul, A. & Tabas, I. Efferocytosis in health and disease. Nat. Rev. Immunol. 20, 254–267 (2020).

    Article  CAS  PubMed  Google Scholar 

  25. Leid, J. M. et al. Primitive embryonic macrophages are required for coronary development and maturation. Circ. Res. 118, 1498–1511 (2016).

  26. Cahill, T. J. et al. Tissue-resident macrophages regulate lymphatic vessel growth and patterning in the developing heart. Development 148, dev194563 (2021).

  27. Lewis-Israeli, Y. R. et al. Self-assembling human heart organoids for the modeling of cardiac development and congenital heart disease. Nat. Commun. 12, 5142 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Wang, E. Y. et al. Biowire model of interstitial and focal cardiac fibrosis. ACS Cent. Sci. 5, 1146–1158 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Thomas, D., Choi, S., Alamana, C., Parker, K. K. & Wu, J. C. Cellular and engineered organoids for cardiovascular models. Circ. Res. 130, 1780–1802 (2022).

    Article  CAS  PubMed  Google Scholar 

  30. Drakhlis, L. et al. Human heart-forming organoids recapitulate early heart and foregut development. Nat. Biotechnol. 39, 737–746 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Hofbauer, P. et al. Cardioids reveal self-organizing principles of human cardiogenesis. Cell 184, 3299–3317 (2021).

    Article  CAS  PubMed  Google Scholar 

  32. Filippo Buono, M. et al. Human cardiac organoids for modeling genetic cardiomyopathy. Cells 9, 1733 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Marini, V. et al. Long-term culture of patient-derived cardiac organoids recapitulated Duchenne muscular dystrophy cardiomyopathy and disease progression. Front. Cell Dev. Biol. 10, 878311 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Micheu, M. M. & Rosca, A. M. Patient-specific induced pluripotent stem cells as ‘disease-in-a-dish’ models for inherited cardiomyopathies and channelopathies—15 years of research. World J. Stem Cells 13, 281–303 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  35. Silva, A. C. et al. Co-emergence of cardiac and gut tissues promotes cardiomyocyte maturation within human iPSC-derived organoids. Cell Stem Cell 28, 2137–2152 (2021).

    Article  CAS  PubMed  Google Scholar 

  36. Ergir, E. et al. Generation and maturation of human iPSC-derived cardiac organoids in long term culture. Sci. Rep. 12, 17409 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Feric, N. T. & Radisic, M. Maturing human pluripotent stem cell-derived cardiomyocytes in human engineered cardiac tissues. Adv. Drug Deliv. Rev. 96, 110–134 (2016).

    Article  CAS  PubMed  Google Scholar 

  38. Liang, P.-Y., Chang, Y., Jin, G., Lian, X. & Bao, X. Wnt signaling directs human pluripotent stem cells into vascularized cardiac organoids with chamber-like structures. Front. Bioeng. Biotechnol. 10, 1059243 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Helms, H. R., Jarrell, D. K. & Jacot, J. G. Generation of cardiac organoids using cardiomyocytes, endothelial cells, epicardial cells, and cardiac fibroblasts derived from human induced pluripotent stem cells. FASEB J. 33, lb170 (2019).

    Article  Google Scholar 

  40. Kahn-Krell, A. et al. A three-dimensional culture system for generating cardiac spheroids composed of cardiomyocytes, endothelial cells, smooth-muscle cells, and cardiac fibroblasts derived from human induced-pluripotent stem cells. Front. Bioeng. Biotechnol. 10, 908848 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Cui, Y. et al. Single-cell transcriptome analysis maps the developmental track of the human heart. Cell Rep. 26, 1934–1950 (2019).

    Article  CAS  PubMed  Google Scholar 

  42. Nicolás-Ávila, J. A. et al. A network of macrophages supports mitochondrial homeostasis in the heart. Cell 183, 94–109 (2020).

    Article  PubMed  Google Scholar 

  43. Atkins, M. H. et al. Modeling human yolk sac hematopoiesis with pluripotent stem cells. J. Exp. Med. 219, e20211924 (2022).

    Article  CAS  PubMed  Google Scholar 

  44. Sturgeon, C. M., Ditadi, A., Awong, G., Kennedy, M. & Keller, G. Wnt signaling controls the specification of definitive and primitive hematopoiesis from human pluripotent stem cells. Nat. Biotechnol. 32, 554–561 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Nunes, S. S. et al. Biowire: a platform for maturation of human pluripotent stem cell-derived cardiomyocytes. Nat. Methods 10, 781–787 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Zhao, Y. et al. A platform for generation of chamber-specific cardiac tissues and disease modeling. Cell 176, 913–927 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Wu, Q. et al. Automated fabrication of a scalable heart-on-a-chip device by 3D printing of thermoplastic elastomer nanocomposite and hot embossing. Bioact. Mater. 33, 46–60 (2024).

    CAS  PubMed  Google Scholar 

  48. Suryawanshi, H. et al. Cell atlas of the foetal human heart and implications for autoimmune-mediated congenital heart block. Cardiovasc. Res. 116, 1446–1457 (2020).

    Article  CAS  PubMed  Google Scholar 

  49. Dhahri, W. et al. In vitro matured human pluripotent stem cell-derived cardiomyocytes form grafts with enhanced structure and function in injured hearts. Circulation 145, 1412–1426 (2022).

    Article  CAS  PubMed  Google Scholar 

  50. Kattman, S. J. et al. Stage-specific optimization of activin/nodal and BMP signaling promotes cardiac differentiation of mouse and human pluripotent stem cell lines. Cell Stem Cell 8, 228–240 (2011).

    Article  CAS  PubMed  Google Scholar 

  51. Euler, D. E. Cardiac alternans: mechanisms and pathophysiological significance. Cardiovasc. Res. 42, 583–590 (1999).

    Article  CAS  PubMed  Google Scholar 

  52. Kim, R. et al. Mechanical alternans is associated with mortality in acute hospitalized heart failure: prospective mechanical alternans study (MAS). Circ. Arrhythm. Electrophysiol. 7, 259–266 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Chen, B. et al. Critical roles of junctophilin-2 in T-tubule and excitation–contraction coupling maturation during postnatal development. Cardiovasc. Res. 100, 54–62 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Rienks, M., Papageorgiou, A.-P., Frangogiannis, N. G. & Heymans, S. Myocardial extracellular matrix. Circ. Res. 114, 872–888 (2014).

    Article  CAS  PubMed  Google Scholar 

  55. Singh, J. P. & Young, J. L. The cardiac nanoenvironment: form and function at the nanoscale. Biophys. Rev. 13, 625–636 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Chen, Q. M. & Maltagliati, A. J. Nrf2 at the heart of oxidative stress and cardiac protection. Physiol. Genomics 50, 77–97 (2018).

    Article  CAS  PubMed  Google Scholar 

  57. Segawa, K. & Nagata, S. An apoptotic ‘eat me’ signal: phosphatidylserine exposure. Trends Cell Biol. 25, 639–650 (2015).

    Article  CAS  PubMed  Google Scholar 

  58. Naeini, M. B., Bianconi, V., Pirro, M. & Sahebkar, A. The role of phosphatidylserine recognition receptors in multiple biological functions. Cell. Mol. Biol. Lett. 25, 23 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Gerlach, B. D. et al. Efferocytosis induces macrophage proliferation to help resolve tissue injury. Cell Metab. 33, 2445–2463 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Gomes, M. T. et al. Phosphatidylserine externalization by apoptotic cells is dispensable for specific recognition leading to innate apoptotic immune responses. J. Biol. Chem. 298, 102034 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Krahling, S., Callahan, M. K., Williamson, P. & Schlegel, R. A. Exposure of phosphatidylserine is a general feature in the phagocytosis of apoptotic lymphocytes by macrophages. Cell Death Differ. 6, 183–189 (1999).

    Article  CAS  PubMed  Google Scholar 

  62. Liao, Y. H. et al. Interleukin-17A contributes to myocardial ischemia/reperfusion injury by regulating cardiomyocyte apoptosis and neutrophil infiltration. J. Am. Coll. Cardiol. 59, 420–429 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chang, S.-L. et al. Interleukin-17 enhances cardiac ventricular remodeling via activating MAPK pathway in ischemic heart failure. J. Mol. Cell. Cardiol. 122, 69–79 (2018).

    Article  CAS  PubMed  Google Scholar 

  64. Baldeviano, G. C. et al. Interleukin-17A is dispensable for myocarditis but essential for the progression to dilated cardiomyopathy. Circ. Res. 106, 1646–1655 (2010).

    Article  CAS  PubMed  Google Scholar 

  65. Gosselin, D. et al. Environment drives selection and function of enhancers controlling tissue-specific macrophage identities. Cell 159, 1327–1340 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Gosselin, D. et al. An environment-dependent transcriptional network specifies human microglia identity. Science 356, eaal3222 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  67. Lavin, Y. et al. Tissue-resident macrophage enhancer landscapes are shaped by the local microenvironment. Cell 159, 1312–1326 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Zhou, X. et al. Circuit design features of a stable two-cell system. Cell 172, 744–757 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Bonnardel, J. et al. Stellate cells, hepatocytes, and endothelial cells imprint the Kupffer cell identity on monocytes colonizing the liver macrophage niche. Immunity 51, 638–654 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Israeli-Rosenberg, S., Manso, A. M., Okada, H. & Ross, R. S. Integrins and integrin-associated proteins in the cardiac myocyte. Circ. Res. 114, 572–586 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Song, R. & Zhang, L. Cardiac ECM: its epigenetic regulation and role in heart development and repair. Int. J. Mol. Sci. 21, 8610 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Scott, C. L. et al. The transcription factor ZEB2 is required to maintain the tissue-specific identities of macrophages. Immunity 49, 312–325 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Schulz, C. et al. A lineage of myeloid cells independent of Myb and hematopoietic stem cells. Science 336, 86–90 (2012).

    Article  CAS  PubMed  Google Scholar 

  74. Kelly, L. M., Englmeier, U., Lafon, I., Sieweke, M. H. & Graf, T. MafB is an inducer of monocytic differentiation. EMBO J. 19, 1987–1997 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Aziz, A., Soucie, E., Sarrazin, S. & Sieweke, M. H. MafB/c-Maf deficiency enables self-renewal of differentiated functional macrophages. Science 326, 867–871 (2009).

    Article  CAS  PubMed  Google Scholar 

  76. T’Jonck, W., Guilliams, M. & Bonnardel, J. Niche signals and transcription factors involved in tissue-resident macrophage development. Cell. Immunol. 330, 43–53 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  77. Ueno, M. et al. Layer V cortical neurons require microglial support for survival during postnatal development. Nat. Neurosci. 16, 543–551 (2013).

    Article  CAS  PubMed  Google Scholar 

  78. Wakselman, S. et al. Developmental neuronal death in hippocampus requires the microglial CD11b integrin and DAP12 immunoreceptor. J. Neurosci. 28, 8138–8143 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Marín-Teva, J. L. et al. Microglia promote the death of developing Purkinje cells. Neuron 41, 535–547 (2004).

    Article  PubMed  Google Scholar 

  80. Grune, J. et al. Neutrophils incite and macrophages avert electrical storm after myocardial infarction. Nat. Cardiovasc. Res. 1, 649–664 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  81. Jia, D. et al. Cardiac resident macrophage-derived legumain improves cardiac repair by promoting clearance and degradation of apoptotic cardiomyocytes after myocardial infarction. Circulation 145, 1542–1556 (2022).

    Article  CAS  PubMed  Google Scholar 

  82. Koivumäki, J. T. et al. Structural immaturity of human iPSC-derived cardiomyocytes: in silico investigation of effects on function and disease modeling. Front. Physiol. 9, 80 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  83. Karbassi, E. et al. Cardiomyocyte maturation: advances in knowledge and implications for regenerative medicine. Nat. Rev. Cardiol. 17, 341–359 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  84. Lundy, S. D., Zhu, W. Z., Regnier, M. & Laflamme, M. A. Structural and functional maturation of cardiomyocytes derived from human pluripotent stem cells. Stem Cells Dev. 22, 1991–2002 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Ronaldson-Bouchard, K. et al. Advanced maturation of human cardiac tissue grown from pluripotent stem cells. Nature 556, 239–243 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Tiburcy, M. et al. Defined engineered human myocardium with advanced maturation for applications in heart failure modeling and repair. Circulation 135, 1832–1847 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Guo, Y. & Pu, W. T. Cardiomyocyte maturation. Circ. Res. 126, 1086–1106 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Park, D. S. et al. iPS-cell-derived microglia promote brain organoid maturation via cholesterol transfer. Nature 623, 397–405 (2023).

    Article  CAS  PubMed  Google Scholar 

  89. Patterson, A. J. & Zhang, L. Hypoxia and fetal heart development. Curr. Mol. Med. 10, 653–666 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Irion, S. et al. Identification and targeting of the ROSA26 locus in human embryonic stem cells. Nat. Biotechnol. 25, 1477–1482 (2007).

    Article  CAS  PubMed  Google Scholar 

  91. Reubinoff, B. E., Pera, M. F., Fong, C. Y., Trounson, A. & Bongso, A. Embryonic stem cell lines from human blastocysts: somatic differentiation in vitro. Nat. Biotechnol. 18, 399–404 (2000).

    Article  CAS  PubMed  Google Scholar 

  92. Thomson, J. A. et al. Embryonic stem cell lines derived from human blastocysts. Science 282, 1145–1147 (1998).

    Article  CAS  PubMed  Google Scholar 

  93. Funakoshi, S. et al. Generation of mature compact ventricular cardiomyocytes from human pluripotent stem cells. Nat. Commun. 12, 3155 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. Fernandes, I., Funakoshi, S., Hamidzada, H., Epelman, S. & Keller, G. Modeling cardiac fibroblast heterogeneity from human pluripotent stem cell-derived epicardial cells. Nat. Commun. 14, 8183 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Landau, S., Shor, E., Radisic, M. & Levenberg, S. Quantitative image analysis of tissue properties: a MATLAB tool for measuring morphology and co-localization in 2D images. Preprint at bioRxiv https://doi.org/10.1101/2024.04.03.587971 (2024).

  96. 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 

  97. Butler, A., Hoffman, P., Smibert, P., Papalexi, E. & Satija, R. Integrating single-cell transcriptomic data across different conditions, technologies, and species. Nat. Biotechnol. 36, 411–420 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Stuart, T. et al. Comprehensive integration of single-cell data. Cell 177, 1888–1902 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Hao, Y. et al. Dictionary learning for integrative, multimodal and scalable single-cell analysis. Nat. Biotechnol. 42, 293–304 (2023).

    Article  PubMed  PubMed Central  Google Scholar 

  100. Korsunsky, I. et al. Fast, sensitive and accurate integration of single-cell data with Harmony. Nat. Methods 16, 1289–1296 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Foroutan, M. et al. Single sample scoring of molecular phenotypes. BMC Bioinformatics 19, 404 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Tyanova, S. et al. The Perseus computational platform for comprehensive analysis of (prote)omics data. Nat. Methods 13, 731–740 (2016).

    Article  CAS  PubMed  Google Scholar 

  103. Azam, M. A. et al. Effects of late sodium current blockade on ventricular refibrillation in a rabbit model. Circ. Arrhythm. Electrophysiol. 10, e004331 (2017).

    Article  CAS  PubMed  Google Scholar 

  104. Si, D. et al. Essential role of ryanodine receptor 2 phosphorylation in the effect of azumolene on ventricular arrhythmia vulnerability in a rabbit heart model. J. Cardiovasc. Electrophysiol. 29, 1707–1715 (2018).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to thank the Immune Profiling Team at the Tumor Immunotherapy Program (Princess Margaret Cancer Centre, Toronto, Canada) for processing of mass cytometry samples and guidance on experimental design and analysis (G. Boukhaled, B. Wang and D. Brooks). We would also like to thank the Princess Margaret Genomics Centre (Toronto, Canada) for processing of RNA sequencing samples and the SickKids-UHN Flow Cytometry Facility (Toronto, Canada) for sample sorting. This work was supported by the Canadian Institutes of Health Research (S.E., H.H. and M.R.); the Ted Rogers Centre for Heart Research (S.E. and H.H.); the Peter Munk Cardiac Centre (S.E.); Medicine by Design (S.E., G.K. and M.A.L.); the Stem Cell Network (M.R. (MP-C4R1-3), S.E. and H.H.); and National Institutes of Health (M.R. (2R01 HL076485)). The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

H.H. designed and performed experiments, with the help of S.P.-G., Q.W., S.M., C.K., U.K., N.R., R.A.G., M.W., W.C., S.L., T.S., Y.Z., E.B., J.N. and S.V. H.H. performed the bioinformatics analyses. Q.W., G.M.K., M.H.A., M.J.G.-G., E.Y.W., I.F. and T.V.S. generated bioengineering platforms and produced and differentiated cells. B.R., T.M., A.C.A., A.G., P.B., K.N., M.A.L., G.K. and M.R. contributed to data interpretation and provided expertise. S.E. and M.R. conceived the study and obtained funding. H.H. and S.E. wrote the manuscript. All authors reviewed and approved the final version of the manuscript.

Corresponding author

Correspondence to Slava Epelman.

Ethics declarations

Competing interests

M.R., Y.Z. and Q.H. are inventors on patents related to the Biowire II cardiac tissue cultivation and maturation protocols that are used as a main experimental system in this manuscript (all figures). These patents and applications cover platform fabrication, cell seeding and tissue cultivation (patent 10,034,738; patent 10,254,274; patent 11,913,940; application 17,520,303 filed on 5 November 2021; application 17798047, publication date 9 March 2023; application 17520303, publication date 12 May 2022; application 17798047, publication date 9 March 2023; and application 17520303, publication date 12 May 2022). Patents 10,254,274 and 11,913,940 are licensed to Valo Health. M.R. and Y.Z. receive royalty payments and annual fees for licensing of these inventions. M.R. and Y.Z. are also eligible for milestone payments from Valo Health related to successful discovery and translation of molecules using the Biowire II platform. All other authors declare no competing interests.

Peer review

Peer review information

Nature Cardiovascular Research thanks Charles Murry and the other, anonymous, reviewer(s) 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 Integration of hESC-macrophages into bioengineered human cardiac microtissues.

(a) Flow cytometry of hESC-cardiomyocytes on day 16 post-differentiation. (b) qPCR of generic or cardiac lineage-specific genes in human primary cardiac fibroblasts compared to human primary dermal fibroblasts at passages 3 or 5. n = 3 replicates per group from one experiment. (c-d) Immunofluorescence confocal imaging of microtissues 14 days post-seeding with or without hESC-macrophages. Images were acquired at the surface or deep within the tissue. Scale bar: 100 µm (C, left), 50 µm (C, right), 100 µm (D). (e) hESC-macrophages were seeded either alone or with a range of abundances of human primary cardiac fibroblasts. The number of hESC-macrophages were counted over three weeks. n = 3 per group, representative experiment shown, repeated two times. Diagram made with BioRender.com. (f) hESC-macrophages were incubated with control or conditioned media from human primary cardiac fibroblasts. The number of hESC-macrophages were counted over three weeks. n = 3 per group, representative experiment shown, repeated two times. cTnT: cardiac troponin T; CM: cardiomyocyte; FB: fibroblast; MF: macrophage. One-way ANOVA with P values adjusted for multiple comparisons using the Tukey-Kramer test: *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 2 Defining contamination and dissociation-induced gene expression in hESC-macrophages.

(a) hESC-macrophages were stimulated with LPS in the presence or absence of the transcription inhibitor Flavopiridol. qPCR was performed on IL6 and TNF with expression normalized to the housekeeping gene b2M. n = 3 replicates per group from one experiment. (b) In control experiments, hESC-macrophages were incubated either (1) alone, (2) with fibroblasts or (3) with fibroblasts and hESC-cardiomyocytes during a 40-minute digestion period at 37 degrees Celsius. hESC-macrophages were sorted for bulk RNA sequencing (n = 3 replicates per group from one experiment). (c) Representative gating strategy for fluorescence activated cell sorting (FACS) isolation of hESC-macrophages from each digestion control group in (B). CD14 + RFP + CD45 + DAPI- live single cells were sorted for bulk RNA sequencing. (d) Principal component analysis. (e) Volcano plots showing differentially expressed genes between CM + FB + MF vs. MF and FB + MF vs. MF. (f-g) Microtissues (HT-Biowires) were seeded in combinations of hESC-cardiomyocytes, human primary cardiac fibroblasts and/or hESC-macrophages. On day 14, hESC-macrophages were sorted for bulk RNA sequencing. (f) Representative gating strategy for fluorescence activated cell sorting (FACS) isolation of hESC-macrophages from microtissues. CD14 + RFP + CD45 + DAPI- live single cells were sorted for bulk RNA sequencing (n = 3 microtissues per group from one experiment). (g) Principal component analysis of hESC-macrophages sorted from each group in microtissues in (F). (h) CM + FB + MF vs. FB + MF DEGs compared in Biowire vs. digestion controls, or FB + MF vs. MF DEGs compared in Biowires vs. in digestion controls. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. One-way ANOVA with P values adjusted for multiple comparisons using the Šídák test: *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 3 hESC-macrophages improve electromechanical function and reduce electrical instability in iCell-cardiomyocyte containing Biowires.

Microtissues were seeded with iPSC-cardiomyocytes (iCells) and primary human cardiac fibroblasts with or without hESC-macrophages in the Biowire II platform. (a) Independent experiments of distinct iPSC-cardiomyocyte and hESC-macrophage batches. Biowires were seeded with or without hESC-macrophages. Force and electrical properties were measured day 11 post-seeding. Batch 2: n = 3 (control) or n = 5 (hESC-macrophage) microtissues per group from one experiment, except for excitation threshold and maximum capture rate where n = 5 control microtissues. Batch 3: n = 4 microtissues per group from one experiment. (b) Microtissues were stimulated at increasing frequencies from 1 Hz to 4 Hz. Graph shows the tracking of pixel movement during contraction and relaxation. n = 4 per group. (c) Schematic depicting the categorization of weak amplitude beats during mechanical alternans. (d) Electrical instability threshold representing the minimum frequency at which an alternating reduced force amplitude pattern was observed. n = 4 microtissues per group from one experiment. We defined a reduced force amplitude based on whether the amplitude was at least 15% less than the prior measured amplitude. (e) Percentage of peaks at 1 Hz or 2 Hz that contain alternating amplitudes were quantified. n = 4 microtissues per group, one experiment. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. Unpaired two-tailed t-test (A, D-E): *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 4 Baseline heart rate correlates with the effect of hESC-macrophages on the heart rate of cardiac microtissues.

(a) Microtissues (HT-Biowires) were seeded with hESC-cardiomyocytes and fibroblasts with or without hESC-macrophages. Active force was measured on day 14 post-seeding. n = 26 (control) or n = 24 (hESC-macrophage) microtissues per group pooled from two independent experiments. Unpaired t-test was performed. (b) Relationship between the baseline heart rate of microtissues (HT-Biowires, except for iCell data point from Biowire II platform) without hESC-macrophages to the change in heart rate upon addition of hESC-macrophages. Each data point represents the average of a distinct experiment, with the batch of cardiomyocytes used indicated. Simple linear regression was performed, reporting R-squared and P value indicating whether the slope is significantly non-zero. (c) Fold-change in active force in microtissues (HT-Biowires, except for iCell data from Biowire II platform) with hESC-macrophages relative to controls in each experiment performed. n = 14, 20, 6, 9, 11, 2, 6, 18, 6 control microtissues (left to right) or n = 8, 30, 5, 11, 14, 9, 16, 18, 6 microtissues with hESC-macrophages (left to right). Each group represents an independent experiment. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. Unpaired two-tailed t-test (A, C): *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 5 Expression of proteins in mass spectrometry-based proteomics of cardiac microtissues.

Liquid chromatography mass spectrometry was performed on total protein isolated from individual microtissues (HT-Biowires) on day 3 and day 14. (a) The number of proteins detected in each sample. (b) Histogram showing the number of proteins that are shared across multiple samples as indicated on the x-axis. (c) LFQ intensity of contractile machinery proteins in microtissues with or without hESC-macrophages on day 3. (d) LFQ intensity of JHP2 in microtissues with or without hESC-macrophages on day 14. n = 9 microtissues per group from one experiment. (e) Immunofluorescence and confocal microscopy was performed on microtissues with or without hESC-macrophages stained with a-actinin and MLC2v (as in Fig. 3G). Myofibril alignment was quantified. n = 8 microtissues per group from one experiment. (f) LFQ intensity of collagen proteins in microtissues with or without hESC-macrophages on day 3. n = 7 (control) or n = 8 (hESC-macrophage) microtissues per group from one experiment. (g) LFQ intensity of extracellular matrix proteins in microtissues with or without hESC-macrophages on day 3. n = 7 (control) or n = 8 (hESC-macrophage) microtissues per group from one experiment. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. Multiple unpaired (two-tailed) t-tests were conducted with P values adjusted for multiple comparisons using the Holm-Šídák method (C, F-G). Unpaired two-tailed t-test was performed for pairwise comparisons of two groups (D-E). *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 6 hESC-macrophages increase calcium amplitude in human cardiac microtissues without changes in calcium transients or in single hESC-cardiomyocyte ion channel function.

(a) Liquid chromatography mass spectrometry was performed on total protein isolated from individual microtissues (HT-Biowires). LFQ intensity of calcium handling proteins in microtissues with or without hESC-macrophages on day 14. n = 8 (control) n = 9 (hESC-macrophage) microtissues per group from one experiment. Multiple unpaired (two-tailed) t-tests were conducted with P values adjusted for multiple comparisons using the Holm-Šídák method. (b-g) Microtissues were seeded with hESC-cardiomyocytes and human primary cardiac fibroblasts with or without hESC-macrophages in a 24-well based HT-Biowire platform. (b) Microtissues were incubated with a calcium indicator dye (Fluo-4). Calcium amplitude relative to baseline intensity was measured 14 days post-seeding either during spontaneous beating or while pacing at 3 Hz. n = 12, 13, 12 13 microtissues per group (left to right) from one experiment. (c) Conduction velocity in microtissues with or without hESC-macrophages paced at 3 Hz. n = 28 (control) or n = 19 (hESC-macrophage) microtissues per group pooled from two independent experiments. Unpaired two-tailed t-test was performed. (d-e) Microtissues were stimulated at increasing frequencies. Calcium transient duration from depolarization to either 50% (CaTD50) or 80% (CaTD80) decay in microtissues with or without hESC-macrophages paced at 2–5 Hz 14 days post-seeding. n = 29, 17, 34, 18, 11, 12, 8, 10 microtissues per group (left to right) pooled from two independent. (f) Relaxation time (peak to baseline) in microtissues with or without hESC-macrophages paced at 2–5 Hz 14 days post-seeding. n = 29, 17, 34, 19, 11, 12, 8, 10 microtissues per group (left to right) pooled from two independent experiments. (g) Time to peak from 90% depolarization to 10% repolarization in microtissues with or without hESC-macrophages paced at 2–5 Hz 14 days post-seeding. n = 29, 16, 36, 17, 12, 11, 8, 9 microtissues per group (left to right) pooled from two independent experiments. (h) Schematic of experimental design made with BioRender.com. Microtissues were seeded with hESC-cardiomyocytes and fibroblasts with or without hESC-macrophages. On day 14, microtissues were dissociated and cells were plated for single cardiomyocyte patch clamp recordings. (i) Representative ICa tracings (left). Peak ICa amplitude at 0 mV (right). n = 11 (control) or n = 8 (hESC-macrophage) cardiomyocytes per group pooled from two independent experiments. (j) Representative INa tracings (left). Peak INa amplitude at −20 mV (right). n = 8 (control) or n = 11 (hESC-macrophage) cardiomyocytes per group from one experiment. (k) INa+ Current density-voltage (I-V) plot and activation curve with the least-square fits to Boltzmann function. n = 8–11 per group, one experiment. (l) Representative action potential (AP) tracings along with peak AP amplitude, AP duration at 50% repolarization (APD50) and resting membrane potential (RMP) (left to right). n = 3 cardiomyocytes per group from one experiment. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. Two-way ANOVA with P values adjusted for multiple comparisons using the Holm-Šídák method (B, D-G). Unpaired two-tailed t-test (C, I, J, L). *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 7 hESC-macrophages reduce accumulation of mitochondrial proteins in cardiac microtissues.

Cytometry by time-of-flight (CyTOF) was performed on microtissues (HT-Biowires) with or without hESC-macrophages on day 3 post-seeding. n = 6 replicates per group, each replicate represented 8 pooled microtissues, experiment performed once. (a) UMAP visualization of 580,776 live and dead cells following standard quality control filtering, split by replicate in each group. (b) Percentage of Dead ATP5hi group relative to all events in each replicate. (c) Number of cardiomyocytes or fibroblasts acquired in each replicate in microtissues with or without hESC-macrophages. (d) Percentage of ATP5Amid cells relative to the number of cardiomyocytes (left) or the number of cardiomyocytes and fibroblasts (right) in microtissues with or without hESC-macrophages. (e) Pathways downregulated in microtissues with hESC-macrophages on day 14 from mass spectrometry-based proteomics data (as in Fig. 3). (f) Normalized expression of DNA 1 or DNA 2 in each cardiomyocyte subcluster (averaged in each replicate) in microtissues with or without hESC-macrophages. (g) Normalized expression live-dead label in each replicate (averaged) in CM-4 from microtissues with or without hESC-macrophages. (h) Microtissues were seeded with hESC-cardiomyocytes and fibroblasts with or without hESC-macrophages. On day 3 post-seeding, microtissues were labelled with the MitoTracker dye and flow cytometry was performed. Geometric mean fluorescence intensity (MFI) of MitoTracker in total cells, cardiomyocytes (CD45CD14), or fibroblasts (CD45dimCD14dim) is shown. n = 6 microtissues per group, two experiments shown (MitoTracker Green and MitoTracker Deep Red). (i) Normalized expression of cardiac troponin T (cTnT) in each replicate (averaged) for all cardiomyocyte subclusters (CyTOF data) in microtissues with or without hESC-macrophages. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. Two-way ANOVA with P values adjusted for multiple comparisons using the Šídák method (B, F, I), or unpaired two-tailed t-test (C, D, E, G, H). *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 8 Global downregulation of metabolic proteins in fibroblasts in cardiac microtissues with hESC-macrophages.

(a) CyTOF data of fibroblasts re-clustered showing 5 subclusters. n = 6 replicates per group, each replicate represented 8 pooled microtissues, experiment performed once. (A-F). (b) Frequency of each subcluster of fibroblasts in microtissues with or without hESC-macrophages. (c) Expression of each marker in each subcluster of fibroblasts. (d-f) Expression of each marker in microtissues with or without hESC-macrophages for FB-1 versus FB-2 (d), FB-3 or FB-4 (f). CM: cardiomyocyte; FB: fibroblast; MF: macrophage. Multiple unpaired two-tailed t-tests were conducted with P values adjusted for multiple comparisons using the Bonferroni-Dunn method (D-F), or two-way ANOVA was performed with P values adjusted with Šídák method (B). *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Extended Data Fig. 9 Gating strategy for sorting hESC-cardiomyocytes from microtissues for bulk RNA sequencing.

(a) Microtissues (HT-Biowires) were seeded with hESC-cardiomyocytes and fibroblasts with or without hESC-macrophages for bulk RNA sequencing of hESC-cardiomyocytes on day 3 post-seeding. Representative gating strategy shown. DAPI live single cells were first gated, followed by CD45CD14CD90RFP cells. (b) hESC-cardiomyocytes, fibroblasts and hESC-macrophages were independently stained and acquired for flow cytometry. Data was overlaid into a single plot. CD45CD14 cells were hESC-cardiomyocytes, CD45dimCD14dim cells were fibroblasts, and CD45+CD14+ cells were hESC-macrophages, confirming the gating strategy utilized in (A). (c) CD45 expression of populations shown in (B), showing that hESC-cardiomyocytes are CD45, fibroblasts are CD45dim and hESC-macrophages are CD45+. (d) hESC-cardiomyocytes or fibroblasts were pre-labeled with CFSE prior to seeding microtissues. On day 3, flow cytometry was performed showing that the CFSE+ hESC-cardiomyocytes were CD14RFP, whereas CFSE+ fibroblasts were CD14dimRFPdim, confirming the gating strategy utilized in (A). CM: cardiomyocyte; FB: fibroblast; MF: macrophage.

Extended Data Fig. 10 hESC-macrophage efferocytosis alters the total proteome in cardiac microtissues.

(a) Expression of proliferation genes in bulk RNA sequencing data of hESC-macrophages as in Fig. 1. n = 3 replicates per group from one experiment. (b) Number of hESC-cardiomyocytes or hESC-macrophages in microtissues with hESC-macrophages in PBS-treated versus Annexin-treated microtissues assessed using flow cytometry. n = 4 microtissues per group from one experiment. (c) Pathways enriched in microtissues with hESC-macrophages in PBS treated (with efferocytosis) versus Annexin treated (without efferocytosis) in liquid chromatography mass spectrometry-based proteomics data. Fisher’s one-tailed test was performed with correction for multiple comparisons using the g:SCS method as implemented in gProfiler. (d) Human cytokine array (96-plex Discovery Assay) was performed on culture supernatants from microtissues with or without hESC-macrophages on days 1, 3 and 7. Volcano plots show upregulated versus downregulated cytokines at each timepoint. n = 7 replicates (collected from n = 7 microtissues) per group at each timepoint from one experiment. (e) Concentration of key cytokines in culture supernatants at each timepoint in microtissues with or without hESC-macrophages. n = 7 replicates (collected from n = 7 microtissues) per group at each timepoint from one experiment. (f) Microtissues were seeded with or without hESC-macrophages, containing PBS pre-treated versus Annexin pre-treated hESC-cardiomyocytes. On day 14 post-seeding, active force, contraction slope and relaxation slope were measured. n = 17, 14, 14, 10 microtissues per group (left to right) from one experiment, first experiment shown in Fig. 7. CM: cardiomyocyte; FB: fibroblast; MF: macrophage. One-way ANOVA (F) or two-way ANVOA (A, E) with P values adjusted for multiple comparisons using the Šídák method. Unpaired two-tailed t-test (B, D). *P < 0.05, **P < 0.01, ***P < 0.001. Error bars represent mean ± s.e.m.

Source data

Supplementary information

Supplementary Figs. 1 and 2 and legends

Reporting Summary

44161_2024_471_MOESM3_ESM.xlsx

Supplementary Table 1: Human yolk sac and fetal heart scRNA-seq analysis. Supplementary Table 2: Differential gene expression analysis of bulk RNA-seq of contamination controls. Supplementary Table 3: Differential gene expression analysis of bulk RNA-seq of sorted hESC-macrophages from microtissues. Supplementary Table 4: Proteomics of microtissues with or without hESC-macrophages. Supplementary Table 5: scRNA-seq of human fetal ventricular cardiomyocytes. Supplementary Table 6: Differential gene expression analysis of bulk RNA-seq of sorted hESC-cardiomyocytes from microtissues. Supplementary Table 7: Proteomics of Annexin-treated microtissues with or without hESC-macrophages. Supplementary Table 8: Human cytokine array. Supplementary Table 9: Sample information and accession codes for publicly available scRNA-seq datasets.

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 Fig. 6

Statistical Source Data

Source Data Fig. 7

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. 3

Statistical Source Data

Source Data Extended Data Fig. 4

Statistical Source Data

Source Data Extended Data Fig. 5

Statistical Source Data

Source Data Extended Data Fig. 6

Statistical Source Data

Source Data Extended Data Fig. 7

Statistical Source Data

Source Data Extended Data Fig. 8

Statistical Source Data

Source Data Extended Data Fig. 10

Statistical Source Data

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

Hamidzada, H., Pascual-Gil, S., Wu, Q. et al. Primitive macrophages induce sarcomeric maturation and functional enhancement of developing human cardiac microtissues via efferocytic pathways. Nat Cardiovasc Res 3, 567–593 (2024). https://doi.org/10.1038/s44161-024-00471-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s44161-024-00471-7

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research