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.

  • Letter
  • Published:

Functional analysis of a common BAG3 allele associated with protection from heart failure

Abstract

Multiple genetic association studies have correlated a common allelic block linked to the BAG3 gene with a decreased incidence of heart failure, but the molecular mechanism remains elusive. In this study, we used induced pluripotent stem cells to test if the only coding variant in this allele block, BAG3C151R, alters protein and cellular function in human cardiomyocytes. Quantitative protein interaction analysis identified changes in BAG3C151R protein partners specific to cardiomyocytes. Knockdown of genes encoding for BAG3-interacting factors in cardiomyocytes followed by myofibrillar analysis revealed that BAG3C151R associates more strongly with proteins involved in the maintenance of myofibrillar integrity. Finally, we demonstrate that cardiomyocytes expressing the BAG3C151R variant have improved response to proteotoxic stress in a dose-dependent manner. This study suggests that BAG3C151R could be responsible for the cardioprotective effect of the haplotype block, by increasing cardiomyocyte protection from stress. Preferential binding partners of BAG3C151R may reveal potential targets for cardioprotective therapies.

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: The putative cardioprotective variant rs2234962-BAG3C151R results in a change in the profile of BAG3 co-precipitation partners in a cardiomyocyte background.
Fig. 2: BAG3C151R favors binding to factors required for maintenance of myofibrillar integrity.
Fig. 3: The rs2234962/BAG3C151R variant provides enhanced resistance to the proteotoxic drug bortezomib in iPS-CMs.

Similar content being viewed by others

Data availability

Mass spectrometry raw data and Skyline files for the targeted proteomics quantification have been deposited to the ProteomeXchange Consortium using the Panorama Public partner repository (https://panoramaweb.org/28ZWVY.url). All other source data used for analyses are included in the manuscript files, except for the source image files, which are available from the authors upon reasonable request.

References

  1. Virani, S. S. et al. Heart disease and stroke statistics—2021 update: a report from the American Heart Association. Circulation 143, e254–e743 (2021).

  2. Burke, M. A., Cook, S. A., Seidman, J. G. & Seidman, C. E. Clinical and mechanistic insights into the genetics of cardiomyopathy. J. Am. Coll. Cardiol. 68, 2871–2886 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Weintraub, R. G., Semsarian, C. & Macdonald, P. Dilated cardiomyopathy. Lancet 390, 400–414 (2017).

    Article  CAS  PubMed  Google Scholar 

  4. Harper, A. R., Nayee, S. & Topol, E. J. Protective alleles and modifier variants in human health and disease. Nat. Rev. Genet. 16, 689–701 (2015).

    Article  CAS  PubMed  Google Scholar 

  5. Villard, E. et al. A genome-wide association study identifies two loci associated with heart failure due to dilated cardiomyopathy. Eur. Heart J. 32, 1065–1076 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Esslinger, U. et al. Exome-wide association study reveals novel susceptibility genes to sporadic dilated cardiomyopathy. PLoS ONE 12, e0172995 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Aragam, K. G. et al. Phenotypic refinement of heart failure in a national biobank facilitates genetic discovery. Circulation https://doi.org/10.1161/CIRCULATIONAHA.118.035774 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Shah, S. et al. Genome-wide association and Mendelian randomisation analysis provide insights into the pathogenesis of heart failure. Nat. Commun. 11, 163 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Choquet, H. et al. Meta-analysis of 26 638 individuals identifies two genetic loci associated with left ventricular ejection fraction. Circ. Genom. Precis. Med. 13, e002804 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Verweij, N. et al. The genetic makeup of the electrocardiogram. Cell Syst. 11, 229–238 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. de Denus, S. et al. A genetic association study of heart failure: more evidence for the role of BAG3 in idiopathic dilated cardiomyopathy. ESC Heart Fail. 7, 4384–4389 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Garnier, S. et al. Genome-wide association analysis in dilated cardiomyopathy reveals two new players in systolic heart failure on chromosomes 3p25.1 and 22q11.23. Eur. Heart J. 42, 2000–2011 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Aung, N. et al. Genome-wide analysis of left ventricular image-derived phenotypes identifies fourteen loci associated with cardiac morphogenesis and heart failure development. Circulation 140, 1318–1330 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Norton, N. et al. Genome-wide studies of copy number variation and exome sequencing identify rare variants in BAG3 as a cause of dilated cardiomyopathy. Am. J. Hum. Genet. 88, 273–282 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Feldman, A. M. et al. Decreased levels of BAG3 in a family with a rare variant and in idiopathic dilated cardiomyopathy. J. Cell. Physiol. 229, 1697–1702 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Homma, S. et al. BAG3 deficiency results in fulminant myopathy and early lethality. Am. J. Pathol. 169, 761–773 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Behl, C. Breaking BAG: the co-chaperone BAG3 in health and disease. Trends Pharmacol. Sci. 37, 672–688 (2016).

    Article  CAS  PubMed  Google Scholar 

  18. Knezevic, T. et al. BAG3: a new player in the heart failure paradigm. Heart Fail. Rev. 20, 423–434 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Gamerdinger, M., Carra, S. & Behl, C. Emerging roles of molecular chaperones and co-chaperones in selective autophagy: focus on BAG proteins. J. Mol. Med. 89, 1175–1182 (2011).

    Article  CAS  PubMed  Google Scholar 

  20. Martin, T. G. & Kirk, J. A. Under construction: the dynamic assembly, maintenance, and degradation of the cardiac sarcomere. J. Mol. Cell. Cardiol. 148, 89–102 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Judge, L. M. et al. A BAG3 chaperone complex maintains cardiomyocyte function during proteotoxic stress. JCI Insight 2, e94623 (2017).

  22. Ulbricht, A., Arndt, V. & Höhfeld, J. Chaperone-assisted proteostasis is essential for mechanotransduction in mammalian cells. Commun. Integr. Biol. 6, e24925 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  23. Fang, X. et al. Loss-of-function mutations in co-chaperone BAG3 destabilize small HSPs and cause cardiomyopathy. J. Clin. Invest. 127, 3189–3200 (2017).

  24. Martin, T. G. et al. Cardiomyocyte contractile impairment in heart failure results from reduced BAG3-mediated sarcomeric protein turnover. Nat. Commun. 12, 2942 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Franaszczyk, M. et al. The BAG3 gene variants in Polish patients with dilated cardiomyopathy: four novel mutations and a genotype–phenotype correlation. J. Transl. Med. 12, 192 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Chami, N. et al. Nonsense mutations in BAG3 are associated with early-onset dilated cardiomyopathy in French Canadians. Can. J. Cardiol. 30, 1655–1661 (2014).

    Article  PubMed  Google Scholar 

  27. Citro, R. et al. Polymorphisms of the antiapoptotic protein bag3 may play a role in the pathogenesis of tako-tsubo cardiomyopathy. Int. J. Cardiol. 168, 1663–1665 (2013).

    Article  PubMed  Google Scholar 

  28. Meister-Broekema, M. et al. Myopathy associated BAG3 mutations lead to protein aggregation by stalling Hsp70 networks. Nat. Commun. 9, 5342 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Lo Sardo, V. et al. Unveiling the role of the most impactful cardiovascular risk locus through haplotype editing. Cell 175, 1796–1810 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Kishore, S. et al. A non-coding disease modifier of pancreatic agenesis identified by genetic correction in a patient-derived iPSC line. Cell Stem Cell 27, 137–146 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Wang, G. et al. Modeling the mitochondrial cardiomyopathy of Barth syndrome with induced pluripotent stem cell and heart-on-chip technologies. Nat. Med. 20, 616–623 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Warren, C. R. et al. Induced pluripotent stem cell differentiation enables functional validation of GWAS variants in metabolic disease. Cell Stem Cell 20, 547–557 (2017).

    Article  CAS  PubMed  Google Scholar 

  33. Domínguez, F. et al. Dilated cardiomyopathy due to BLC2-associated athanogene 3 (BAG3) mutations. J. Am. Coll. Cardiol. 72, 2471–2481 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  34. McDermott-Roe, C. et al. Investigation of a dilated cardiomyopathy-associated variant in BAG3 using genome-edited iPSC-derived cardiomyocytes. JCI Insight 4, e128799 (2019).

  35. Orciuolo, E. et al. Unexpected cardiotoxicity in haematological bortezomib treated patients. Br. J. Haematol. 138, 396–397 (2007).

    Article  CAS  Google Scholar 

  36. Grandin, E. W., Ky, B., Cornell, R. F., Carver, J. & Lenihan, D. J. Patterns of cardiac toxicity associated with irreversible proteasome inhibition in the treatment of multiple myeloma. J. Card. Fail. 21, 138–144 (2015).

    Article  CAS  PubMed  Google Scholar 

  37. Kieserman, J. M., Myers, V. D., Dubey, P., Cheung, J. Y. & Feldman, A. M. Current landscape of heart failure gene therapy. J. Am. Heart Assoc. 8, e012239 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Knezevic, T. et al. Adeno-associated virus serotype 9-driven expression of BAG3 improves left ventricular function in murine hearts with left ventricular dysfunction secondary to a myocardial infarction. JACC Basic Transl. Sci. 1, 647–656 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Grafton, F. et al. Deep learning detects cardiotoxicity in a high-content screen with induced pluripotent stem cell-derived cardiomyocytes. eLife 10, e68714 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Maddah, M. et al. Quantifying drug-induced structural toxicity in hepatocytes and cardiomyocytes derived from hiPSCs using a deep learning method. J. Pharmacol. Toxicol. Methods 105, 106895 (2020).

    Article  CAS  PubMed  Google Scholar 

  41. Rosales, W. & Lizcano, F. The histone demethylase JMJD2A modulates the induction of hypertrophy markers in iPSC-derived cardiomyocytes. Front. Genet. 9, 14 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  42. Fine, M. et al. Human-induced pluripotent stem cell-derived cardiomyocytes for studies of cardiac ion transporters. Am. J. Physiol. Cell Physiol. 305, C481–C491 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Yang, J. et al. Phenotypic screening with deep learning identifies HDAC6 inhibitors as cardioprotective in a BAG3 mouse model of dilated cardiomyopathy. Sci. Transl. Med. 14, eabl5654 (2022).

    Article  CAS  PubMed  Google Scholar 

  44. Schroer, A., Pardon, G., Castillo, E., Blair, C. & Pruitt, B. Engineering hiPSC cardiomyocyte in vitro model systems for functional and structural assessment. Prog. Biophys. Mol. Biol. 144, 3–15 (2019).

    Article  CAS  PubMed  Google Scholar 

  45. van Mil, A. et al. Modelling inherited cardiac disease using human induced pluripotent stem cell-derived cardiomyocytes: progress, pitfalls, and potential. Cardiovasc. Res. 114, 1828–1842 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Musunuru, K. et al. Induced pluripotent stem cells for cardiovascular disease modeling and precision medicine: a scientific statement from the American Heart Association. Circ. Genom. Precis. Med. 11, e000043 (2018).

    PubMed  PubMed Central  Google Scholar 

  47. Kane, R. C., Farrell, A. T., Sridhara, R. & Pazdur, R. United States Food and Drug Administration approval summary: bortezomib for the treatment of progressive multiple myeloma after one prior therapy. Clin. Cancer Res. 12, 2955–2960 (2006).

    Article  CAS  PubMed  Google Scholar 

  48. Métais, A. et al. Asb2α–filamin A axis is essential for actin cytoskeleton remodeling during heart development. Circ. Res. 122, e34–e48 (2018).

    Article  PubMed  Google Scholar 

  49. Mao, Z. & Nakamura, F. Structure and function of filamin C in the muscle Z-disc. Int. J. Mol. Sci. 21, 2696 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Collier, M. P. et al. HspB1 phosphorylation regulates its intramolecular dynamics and mechanosensitive molecular chaperone interaction with filamin C. Sci. Adv. 5, eaav8421 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Waldt, N. et al. Filamin A phosphorylation at serine 2152 by the serine/threonine kinase Ndr2 controls TCR-induced LFA-1 activation in T cells. Front. Immunol. 9, 2852 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Klimek, C. et al. The Hippo network kinase STK38 contributes to protein homeostasis by inhibiting BAG3-mediated autophagy. Biochim. Biophys. Acta Mol. Cell Res. 1866, 1556–1566 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Fang, X., Bogomolovas, J., Trexler, C. & Chen, J. The BAG3-dependent and -independent roles of cardiac small heat shock proteins. JCI Insight 4, e126464 (2019).

  54. Muranova, L. K., Shatov, V. M., Slushchev, A. V. & Gusev, N. B. Quaternary structure and hetero-oligomerization of recombinant human small heat shock protein HspB7 (cvHsp). Int. J. Mol. Sci. 22, 7777 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Rauch, J. N. & Gestwicki, J. E. Binding of human nucleotide exchange factors to heat shock protein 70 (Hsp70) generates functionally distinct complexes in vitro. J. Biol. Chem. 289, 1402–1414 (2014).

    Article  CAS  PubMed  Google Scholar 

  56. Clarke, L. et al. The International Genome Sample Resource (IGSR): a worldwide collection of genome variation incorporating the 1000 Genomes Project data. Nucleic Acids Res. 45, D854–D859 (2017).

    Article  CAS  PubMed  Google Scholar 

  57. Marcus, J. H. & Novembre, J. Visualizing the geography of genetic variants. Bioinformatics 33, 594–595 (2017).

    Article  CAS  PubMed  Google Scholar 

  58. Chang, K. T., Guo, J., di Ronza, A. & Sardiello, M. Aminode: identification of evolutionary constraints in the human proteome. Sci. Rep. 8, 1357 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Mandegar, M. A. et al. CRISPR interference efficiently induces specific and reversible gene silencing in human iPSCs. Cell Stem Cell 18, 541–553 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Miyaoka, Y. et al. Isolation of single-base genome-edited human iPS cells without antibiotic selection. Nat. Methods 11, 291–293 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Hockemeyer, D. et al. Genetic engineering of human pluripotent cells using TALE nucleases. Nat. Biotechnol. 29, 731–734 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Regan, J. F. et al. A rapid molecular approach for chromosomal phasing. PLoS ONE 10, e0118270 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  63. Guan, S., Price, J. C., Prusiner, S. B., Ghaemmaghami, S. & Burlingame, A. L. A data processing pipeline for mammalian proteome dynamics studies using stable isotope metabolic labeling. Mol. Cell Proteomics 10, M111.010728 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  64. Verschueren, E. et al. Scoring large-scale affinity purification mass spectrometry datasets with MiST. Curr. Protoc. Bioinformatics 49, 8.19.1–8.19.16 (2015).

    PubMed  Google Scholar 

  65. Teo, G. et al. SAINTexpress: improvements and additional features in significance analysis of INTeractome software. J. Proteomics 100, 37–43 (2014).

    Article  CAS  PubMed  Google Scholar 

  66. Knight, J. D. R. et al. ProHits-viz: a suite of web tools for visualizing interaction proteomics data. Nat. Methods 14, 645–646 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Shannon, P. et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 13, 2498–2504 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Giurgiu, M. et al. CORUM: the comprehensive resource of mammalian protein complexes—2019. Nucleic Acids Res. 47, D559–D563 (2019).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  70. Teo, G. et al. SAINTq: scoring protein–protein interactions in affinity purification—mass spectrometry experiments with fragment or peptide intensity data. Proteomics 16, 2238–2245 (2016).

    Article  CAS  PubMed  Google Scholar 

  71. Hulsen, T., de Vlieg, J. & Alkema, W. BioVenn—a web application for the comparison and visualization of biological lists using area-proportional Venn diagrams. BMC Genomics 9, 488 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  72. Larsson, J. eulerr: Area-Proportional Euler and Venn Diagrams with Ellipses. R package eulerr version 6.1.0. https://cran.r-project.org/web/packages/eulerr/index.html (2020).

  73. Kusebauch, U. et al. Human SRMAtlas: a resource of targeted assays to quantify the complete human proteome. Cell 166, 766–778 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Pino, L. K. et al. The Skyline ecosystem: informatics for quantitative mass spectrometry proteomics. Mass Spectrom. Rev. 39, 229–244 (2020).

    Article  CAS  PubMed  Google Scholar 

  75. Peterson, A. C., Russell, J. D., Bailey, D. J., Westphall, M. S. & Coon, J. J. Parallel reaction monitoring for high resolution and high mass accuracy quantitative, targeted proteomics. Mol. Cell Proteomics 11, 1475–1488 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Choi, M. et al. MSstats: an R package for statistical analysis of quantitative mass spectrometry-based proteomic experiments. Bioinformatics 30, 2524–2526 (2014).

    Article  CAS  PubMed  Google Scholar 

  77. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  78. Phipson, B., Lee, S., Majewski, I. J., Alexander, W. S. & Smyth, G. K. Robust hyperparameter estimation protects against hypervariable genes and improves power to detect differential expression. Ann. Appl. Stat. 10, 946–963 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Sharma, V. et al. Panorama Public: a public repository for quantitative data sets processed in Skyline. Mol. Cell Proteomics 17, 1239–1244 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Pau, G., Fuchs, F., Sklyar, O., Boutros, M. & Huber, W. EBImage—an R package for image processing with applications to cellular phenotypes. Bioinformatics 26, 979–981 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    Article  CAS  PubMed  Google Scholar 

  82. Maddah, M. et al. A non-invasive platform for functional characterization of stem-cell-derived cardiomyocytes with applications in cardiotoxicity testing. Stem Cell Rep. 4, 621–631 (2015).

    Article  CAS  Google Scholar 

  83. Grafton, F. et al. Deep learning detects cardiotoxicity in a high-content screen with induced pluripotent stem cell-derived cardiomyocytes. eLife 10, e68714 (2021).

  84. Ritz, C., Baty, F., Streibig, J. C. & Gerhard, D. Dose–response analysis using R. PLoS ONE 10, e0146021 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  85. Hothorn, T., Bretz, F. & Westfall, P. Simultaneous inference in general parametric models. Biom. J. 50, 346–363 (2008).

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Gladstone Stem Cell Core and the Gladstone Assay Development and Drug Discovery Core for providing their technical support and experimental expertise. We also would like to thank R. Thomas from the Gladstone Institutes Bioinformatics Core for advice on data analysis. We are also very grateful to J. Johnson, R. Huttenhain and G. Jang from the University of California, San Francisco (UCSF) and the Gladstone Institutes for their advice on affinity purification and mass spectrometry experiments. We thank J. Gestwicki and team (UCSF) for their scientific and technical advice. We thank A. Chan, A. Birk, E. Shin, C. Marley and S. Kang (Gladstone Institutes) for their technical support. We also thank F. Chanut from the Gladstone Institutes Editorial Services for feedback on manuscript preparation. We would also like to thank the scientific reviewers at Nature Cardiovascular Research for their invaluable feedback, which contributed greatly to improving this manuscript. J.A.P.-B. was supported by a Graduate Fellowship from Fundación ‘La Caixa’ (ID 100010434, LCF/BQ/US10/10230024), a Bristol Myers Squibb PCO Graduate Fellowship for Assessing Early Drug Liabilities (ID 63376) and a Predoctoral Fellowship from the American Heart Association (15PRE2570008507 and 13PRE1612001307). B.R.C. was supported by the National Institutes of Health (NIH) (R01-HL130533, R01-HL13535801 and P01-HL146366) and by funding from Tenaya Therapeutics. B.R.C. acknowledges support through a gift from the Roddenberry Foundation and Pauline and Thomas Tusher. N.J.K. was supported by P01 HL146366. R.M.K. was supported by NIH fellowship F32AI127291. L.M.J. was supported by a postdoctoral fellowship from the California Institute of Regenerative Medicine (TG2-01160) and a Career Development Award from the National Institute of Child Health and Development (1K12HD072222). The remaining authors received no specific funding for this work.

Author information

Authors and Affiliations

Authors

Contributions

J.A.P.-B., L.M.J., P.-L.S., N.J.K. and B.R.C. designed and supervised the study. J.A.P.-B., C.L.J., A.T., J.J.H., W.V.R. and K.W. performed cell line generation, cell culture and differentiation. J.A.P.-B., R.M.K., E.H.P. and D.L.S. performed affinity purification mass spectrometry experiments and analyses. J.A.P.-B., K.W. and M.A.M. performed siRNA knockdown panel and image analysis. J.A.P.-B. and L.M.J. performed bortezomib toxicity assay and analyses. All authors contributed to writing the manuscript and preparing the figures.

Corresponding author

Correspondence to Bruce R. Conklin.

Ethics declarations

Competing interests

B.R.C. is a founder of Tenaya Therapeutics (www.tenayatherapeutics.com), a company focused on finding treatments for heart failure, including genetic cardiomyopathies. B.R.C. and J.J.H. hold equity in Tenaya Therapeutics. The Krogan laboratory has received research support from Vir Biotechnology and F. Hoffmann-La Roche. N.J.K. has a consulting agreements with the Icahn School of Medicine at Mount Sinai. He is a shareholder of Tenaya Therapeutics, Maze Therapeutics and Interline Therapeutics. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature Cardiovascular Research thanks Jonathan Kirk 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 Expanded genetic data graphs for rs2234962, BAG3C151R, along with conservation of residues affected by BAG3C151R and BAG3E455K variants.

(a) Zoomed out version of Fig. 1a, showing a window of 200KB. Dot color indicates type of nucleotide change. (b) Allele frequency map for rs2234962 depicting all 1000 Genomes populations. (c) Top: Diagram of BAG3 domain structure. Bottom: Amino acid conservation plot for matching BAG3 regions. Decreasing Relative Substitution Score regions (valleys) indicate sequences with high conservation across species and are annotated as Evolutionary Constrained Regions (ECRs). (D-E) Zoomed in regions for the ECRs around BAG3C151(d) and BAG3E455(e).

Extended Data Fig. 2 Co-precipitation profiles of different BAG3 variants overexpressed in a HEK293 cell background.

Dot size represents the amount of co-precipitated protein normalized across variants. Dot color represents absolute protein abundance (spectral counts). Dot rim represents statistical significance. Yellow colored variants are known pathogenic variants associated with DCM. Green variant is putative cardioprotective variant BAG3C151R. Red variant (BAG3P209L) is associated with skeletal myofibrillar myopathy. Black variants are not associated to DCM or any other pathology. Rightmost two columns depict data for BAG3 truncated variant without the BAG domain (BAG3_ ΔBAG) and for the BAG3 protein BAG domain only (BAG3_ΔBAG). For the truncated variants, no normalization by bait levels was performed. N = 5 separate pulldown experiments. Bayesian False Discovery Rate (FDR) obtained using SAINTexpress analysis software (see Methods section).

Source data

Extended Data Fig. 3 Generation of the isogenic cell lines carrying BAG3 variants and a 3xFLAG epitope tag fusion in the endogenous copy of the BAG3 gene.

(a) Workflow for the cell line generation. (b) Strategy for the insertion of a 3xFLAG epitope fusion at the C-terminal of the BAG3 gene. The BAG3C151R-FLAG variant was generated using the same process on a preexisting cell line bearing the C151R mutation. To generate the BAG3E455K-FLAG cell line, the homology arms were engineered to contain the SNP and insert it during recombination. (C-D) Genotypes of the single-cell clones picked for 3xFLAG insertion (c) and the co-segregation of the BAG3E455K variant (d). (e) Genotyping the products of the 3xFLAG insertion by PCR. These genotyping reactions was performed at least twice with identical results. (f) Cells with a heterozygous insertion of the 3xFLAG epitope tag also had a SNP in the other allele that extended the BAG3 protein product by 4 amino acids. (G-H) A droplet digital PCR phasing test was used to select clones that contained the desired SNP variants and the 3xFLAG C-terminal sequence in the same allele. The test used different probes (g) to generate an estimate of linked molecules for each cell line and probe combination (H; See Methods for more details) (i) Insertion of the 3xFLAG fusion in the BAG3 gene did not alter the protein levels. N = 4 and N = 3 cell extracts from separate differentiation batches; analysis performed using a one-way ANOVA. Data is presented as mean values +/− SEM.

Source data

Extended Data Fig. 4 Generation of a cell line with inducible expression of the BAG3WT protein.

(a) Diagram of the editing strategy. On a BAG3−/− cell background, a doxycycline-activated BAG3-3xFLAG expression cassette was inserted in the PPP1R12C (AAVS1) safe-harbor locus. (b) Western blot of the BAG3 expression on BAG3−/−:TetOn-BAG3WT-3xFLAG iPSCs with and without Doxycycline addition. This immunoblotting reaction was performed twice with similar results.

Source data

Extended Data Fig. 5 Affinity purification - mass spectrometry characterization of BAG3 binding partners in a cardiomyocyte background.

(a) Venn diagram of the high confidence BAG3WT protein-protein interactions identified in three cellular backgrounds. HEK293T cells had overexpressed baits, while iPSC have much lower levels of endogenous BAG3 expression than iPS-CM, which could have influenced the results. Each cell type specific dataset was scored separately against its own matched control(s). (b) Volcano plots depicting co-precipitation intensity in BAG3WT cardiomyocytes treated with Bortezomib (100 nM) relative to DMSO (1:10.000) for 24 hours. Horizontal dashed line indicates statistical significance threshold (adjusted p-value < 0.01) and vertical dashed lines indicate a fold change of 2. N = 4 biological replicates from separate differentiation batches. Analysis performed using a hypervariate linear model analysis, and p-values were adjusted for multiple comparisons (see Methods section). (c) Network diagram of the iPS-CM co-precipitation partners identified for BAG3 in this study. Nodes in orange indicate partners that significantly changed when pulling down BAG3E455K. Nodes in Green indicate partners that significantly changed when pulling down BAG3C151R. Dashed lines: known interactions in the iRefIndex database. (d) Venn diagram comparing the BAG3 binding partners identified in an iPS-CM background when using endogenous basal levels of expression, an overexpression system, or endogenous expression under proteotoxic stress. The graph highlights the importance of using endogenous expression for accurate characterization of binding partners, and the information gained from using a stress state.

Source data

Extended Data Fig. 6 Additional sample micrographs from BAG3 knockdown iPS-CM.

(a) BAG3 silencing by siRNA was effective at reducing protein levels (~85% reduction). N = 3 separate knockdown samples. Data is presented as mean values +/− SEM. (b) Additional sample micrographs from BAG3- and Scr-siRNA-treated iPS-CM, plus a no-siRNA condition. Orange arrowheads: BAG3 accumulation on myofibrillar breaks; white arrowheads: BAG3 accumulation on polar ends of cells; yellow arrowheads: iPS-CM displaying myofibrillar aggregation and collapse. These are sample micrographs from a set of randomly acquired images, 9 images per well across 3 separate wells. (c) Sample images in the same magnification used for the automated scoring analysis. Lower magnification allowed for faster acquisition and richer features to use directly in the scoring scheme, but higher magnification images were used elsewhere in this manuscript for easier viewing. These are sample micrographs from a set of 9 images per well across >20 separate wells (training wells for myofibrillar scoring).

Source data

Extended Data Fig. 7 Quality control and additional data from the myofibrillar scoring workflow

(a) Western blot demonstrating that titration of BAG3 siRNA results in decreasing BAG3 and HSPB8 (affected by BAG3 downregulation) levels. N = 2 separate knockdown experiments. (b) BAG3 sarcomere score inversely correlates with BAG3 protein levels. N = 2 biological replicates, with 9 images each. P-value and R2 obtained fitting a linear model. (c) BAG3 sarcomere score when MYBPC3 or cardiac troponin (cTnT) staining was used to train the model to score iPS-CM micrographs. Both myofibrillar protein stains resulted in clear discrimination of BAG3 and Scramble siRNA treated cells. N = 63 images per siRNA from 21 separate knockdown wells. (d) MYBPC3 staining intensity for iPS-CM treated with different doses of BAG3 or MYBPC3 siRNA. N = 2 separate knockdown experiments, with 9 images each. Data is presented as mean values +/− SEM. (e) Plot of the image scores that result from training based on BAG3 or DAPI staining. These stains do not display the same dynamic range as the myofibrillar (MYBPC3) staining. N = 9 for scramble; 18 for the rest. Data is presented as mean values +/− 2x standard deviation. (f) BAG3 Sarcomere Score for the knockdown of selected factors that were not identified in our AP-MS studies. Dots represent mean of 3 replicates from separate wells, each being the median score of 9 images from the same well. Data is presented as mean values +/− SEM. P-val cutoff: 0.05 using a one-way ANOVA with post-hoc Dunnett test. (g, h) Plot of the nuclei count(G) and BAG3 staining(H) intensities for the gene knockdowns used in the siRNA-myofibrillar scoring analyses. Dots represent mean of 3 replicates from separate wells, each being the median score of 9 images from the same well. Data is presented as mean values +/− SEM. P-val cutoff: 0.05 using a one-way ANOVA with post-hoc Dunnett test. (i) Plotting of the APMS intensity ratio for BAG3E455K differential interactors and their BAG3 Sarcomere Score. There is no statistically significant correlation. P-value obtained fitting a linear model. Pearson’s product-moment correlation: −0.27.

Source data

Extended Data Fig. 8 Sample images from selected siRNA knockdowns.

(a) HSPB8 knockdown was the only knockdown to significantly reduce BAG3 levels. FLNA, DDB1 and STK38 are BAG3C151R differential interactors whose knockdown resulted in sarcomere scores similar to BAG3 knockdown. ACTN2 and TNNT2 are well known sarcomere components that display low sarcomere scores similar to BAG3 knockdown, possibly due to reduced sarcomeric density and increased disarray and their previously described role as BAG3 client proteins. HSPB7 and DNAJB6 knockdowns displayed high sarcomere scores (similar to Scramble control). (b) Additional micrographs from selected conditions, with stains in separate panels. HSPB8 highlighting reduction of BAG3 levels, and STUB1/CHIP resulting in an increase of BAG3 elongated aggregates. For all images, scale bar = 100 µM. Magenta/Red: BAG3; Green: MYBPC3; Cyan: DAPI. These sample micrographs were selected from a set of randomly acquired images, 9 images per well across 3 separate wells.

Extended Data Fig. 9 BAG3C151R variant does not alter Bortezomib EC50 in undifferentiated iPSc, and BAG3 overexpression rescues bortezomib sensitivity phenotype in BAG3−/− cells.

(a) Bortezomib dose-response curves for the data used for EC50 calculations in iPS-CM. (b, c) Independent replication of the dose-response curve for Bortezomib in iPS-CM expressing the BAG3C151R variant in heterozygosity or homozygosity. N = 3 separate sets of cells. Data is presented as EC50 and error bars denote 95% confidence intervals. ****: P-value < 0.0001 and ***: p-value < 0.01 using one-way ANOVA with post-hoc Zidak correction. (d, e) Plot of calculated EC50 and dose response curves for Bortezomib in undifferentiated iPSc from different cell lines used in this study. N = 3 separate sets of cells. ****: P-value < 0.0001 and **: p-value < 0.01 using one-way ANOVA with post-hoc Zidak correction. The higher EC50 value for the BAG3−/− cell line is explained by the faster growth rate of that iPSc line. Data is presented as EC50 and error bars denote 95% confidence intervals. (f, g) Calculated EC50 with 95% confidence intervals and dose-response curves for Bortezomib in control (WT/WT), and BAG3−/− iPS-CM with and without BAG3 overexpression. N = 3 separate sets of cells. Data is presented as EC50 and error bars denote 95% confidence intervals. ****: P-value < 0.0001 using one-way ANOVA with post-hoc Zidak correction. Connecting line and error bands in panels A, C, E and G are based on prediction from fitted dose-response model and 95% confidence intervals, respectively.

Source data

Extended Data Fig. 10 BAG3 variants do not change BAG3 protein levels in iPS-CMs.

Capillary immunoassay (SimpleWestern) quantification of BAG3 protein levels in iPS-CM differentiated from iPSCs heterozygous or homozygous for the indicated BAG3 alleles, untreated or treated with Bortezomib (100 nM, 48 h incubation). N = 3 biological replicates from separate differentiation batches. ***: p-value < 0.01; ****: p-value < 0.0001. Two-way ANOVA with Sidak multiple comparison test. Data is presented as mean values +/− SEM.

Source data

Supplementary information

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 Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 3

Image file (uncropped gel).

Source Data Extended Data Fig. 4

Image file (uncropped gel).

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 6

Image files (uncropped gels).

Source Data Extended Data Fig. 7

Image files (uncropped gels).

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 9

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

Perez-Bermejo, J.A., Judge, L.M., Jensen, C.L. et al. Functional analysis of a common BAG3 allele associated with protection from heart failure. Nat Cardiovasc Res 2, 615–628 (2023). https://doi.org/10.1038/s44161-023-00288-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s44161-023-00288-w

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