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:

Atheroprotective roles of smooth muscle cell phenotypic modulation and the TCF21 disease gene as revealed by single-cell analysis

Abstract

In response to various stimuli, vascular smooth muscle cells (SMCs) can de-differentiate, proliferate and migrate in a process known as phenotypic modulation. However, the phenotype of modulated SMCs in vivo during atherosclerosis and the influence of this process on coronary artery disease (CAD) risk have not been clearly established. Using single-cell RNA sequencing, we comprehensively characterized the transcriptomic phenotype of modulated SMCs in vivo in atherosclerotic lesions of both mouse and human arteries and found that these cells transform into unique fibroblast-like cells, termed ‘fibromyocytes’, rather than into a classical macrophage phenotype. SMC-specific knockout of TCF21—a causal CAD gene—markedly inhibited SMC phenotypic modulation in mice, leading to the presence of fewer fibromyocytes within lesions as well as within the protective fibrous cap of the lesions. Moreover, TCF21 expression was strongly associated with SMC phenotypic modulation in diseased human coronary arteries, and higher levels of TCF21 expression were associated with decreased CAD risk in human CAD-relevant tissues. These results establish a protective role for both TCF21 and SMC phenotypic modulation in this disease.

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: Transcriptomic characterization of mouse aortic root atherosclerotic plaques and Tcf21 expression.
Fig. 2: Characterization of SMC phenotypic modulation in the mouse aortic root.
Fig. 3: SMC-specific Tcf21 knockout markedly inhibits SMC phenotypic modulation in mice.
Fig. 4: Identification of modulated SMCs in diseased human coronary arteries.
Fig. 5: TCF21 is associated with SMC modulation in human coronary arteries.
Fig. 6: Reduced TCF21 expression is associated with increased coronary disease risk.

Similar content being viewed by others

Data availability

High-throughput sequencing data (FASTQ) files for all scRNA-Seq, CITE-Seq and ChIP-Seq, as well as cell–gene count matrices for all scRNA-Seq and CITE-Seq experiments, have been deposited in the Gene Expression Omnibus database with the SuperSeries reference number GSE131780. These data were used to generate the images in Figs. 15 and Extended Data Figs. 25. FASTQ files and processed data are also available from the corresponding author upon request.

References

  1. Davies, M. J., Richardson, P. D., Woolf, N., Katz, D. R. & Mann, J. Risk of thrombosis in human atherosclerotic plaques: role of extracellular lipid, macrophage, and smooth muscle cell content. Br. Heart J. 69, 377–381 (1993).

    Article  CAS  Google Scholar 

  2. Libby, P. & Aikawa, M. Stabilization of atherosclerotic plaques: new mechanisms and clinical targets. Nat. Med. 8, 1257–1262 (2002).

    Article  CAS  Google Scholar 

  3. Ross, R. & Glomset, J. A. Atherosclerosis and the arterial smooth muscle cell: proliferation of smooth muscle is a key event in the genesis of the lesions of atherosclerosis. Science 180, 1332–1339 (1973).

    Article  CAS  Google Scholar 

  4. Bennett, M. R., Sinha, S. & Owens, G. K. Vascular smooth muscle cells in atherosclerosis. Circ. Res. 118, 692–702 (2016).

    Article  CAS  Google Scholar 

  5. Gomez, D. & Owens, G. K. Smooth muscle cell phenotypic switching in atherosclerosis. Cardiovasc. Res. 95, 156–164 (2012).

    Article  CAS  Google Scholar 

  6. Shankman, L. S. et al. KLF4-dependent phenotypic modulation of smooth muscle cells has a key role in atherosclerotic plaque pathogenesis. Nat. Med. 21, 628–637 (2015).

    Article  CAS  Google Scholar 

  7. Gomez, D., Shankman, L. S., Nguyen, A. T. & Owens, G. K. Detection of histone modifications at specific gene loci in single cells in histological sections. Nat. Methods 10, 171–177 (2013).

    Article  CAS  Google Scholar 

  8. Schunkert, H. et al. Large-scale association analysis identifies 13 new susceptibility loci for coronary artery disease. Nat. Genet. 43, 333–338 (2011).

    Article  CAS  Google Scholar 

  9. Miller, C. L. et al. Disease-related growth factor and embryonic signaling pathways modulate an enhancer of TCF21 expression at the 6q23.2 coronary heart disease locus. PLoS Genet. 9, e1003652 (2013).

    Article  CAS  Google Scholar 

  10. Miller, C. L. et al. Coronary heart disease-associated variation in TCF21 disrupts a miR-224 binding site and miRNA-mediated regulation. PLoS Genet. 10, e1004263 (2014).

    Article  Google Scholar 

  11. Dettman, R. W., Denetclaw, W. Jr, Ordahl, C. P. & Bristow, J. Common epicardial origin of coronary vascular smooth muscle, perivascular fibroblasts, and intermyocardial fibroblasts in the avian heart. Dev. Biol. 193, 169–181 (1998).

    Article  CAS  Google Scholar 

  12. Winter, E. M. & Gittenberger-de Groot, A. C. Epicardium-derived cells in cardiogenesis and cardiac regeneration. Cell. Mol. Life Sci. 64, 692–703 (2007).

    Article  CAS  Google Scholar 

  13. Acharya, A. et al. The bHLH transcription factor Tcf21 is required for lineage-specific EMT of cardiac fibroblast progenitors. Development 139, 2139–2149 (2012).

    Article  CAS  Google Scholar 

  14. Nurnberg, S. T. et al. Coronary artery disease associated transcription factor TCF21 regulates smooth muscle precursor cells that contribute to the fibrous cap. PLoS Genet. 11, e1005155 (2015).

    Article  Google Scholar 

  15. Herring, B. P., Hoggatt, A. M., Burlak, C. & Offermanns, S. Previously differentiated medial vascular smooth muscle cells contribute to neointima formation following vascular injury. Vasc. Cell 6, 21 (2014).

    Article  Google Scholar 

  16. Wirth, A. et al. G12-G13-LARG-mediated signaling in vascular smooth muscle is required for salt-induced hypertension. Nat. Med. 14, 64–68 (2008).

    Article  CAS  Google Scholar 

  17. Madisen, L. et al. A robust and high-throughput Cre reporting and characterization system for the whole mouse brain. Nat. Neurosci. 13, 133–140 (2010).

    Article  CAS  Google Scholar 

  18. Zheng, G. X. et al. Massively parallel digital transcriptional profiling of single cells. Nat. Commun. 8, 14049 (2017).

    Article  CAS  Google Scholar 

  19. Dobnikar, L. et al. Disease-relevant transcriptional signatures identified in individual smooth muscle cells from healthy mouse vessels. Nat. Commun. 9, 4567 (2018).

    Article  Google Scholar 

  20. Kitchen, C. M., Cowan, S. L., Long, X. & Miano, J. M. Expression and promoter analysis of a highly restricted integrin alpha gene in vascular smooth muscle. Gene 513, 82–89 (2013).

    Article  CAS  Google Scholar 

  21. Stoeckius, M. et al. Simultaneous epitope and transcriptome measurement in single cells. Nat. Methods 14, 865–868 (2017).

    Article  CAS  Google Scholar 

  22. Jacobsen, K. et al. Diverse cellular architecture of atherosclerotic plaque derives from clonal expansion of a few medial SMCs. JCI Insight 2, 95890 (2017).

    Article  Google Scholar 

  23. Vengrenyuk, Y. et al. Cholesterol loading reprograms the microRNA-143/145–myocardin axis to convert aortic smooth muscle cells to a dysfunctional macrophage-like phenotype. Arterioscler. Thromb. Vasc. Biol. 35, 535–546 (2015).

    Article  CAS  Google Scholar 

  24. Sazonova, O. et al. Characterization of TCF21 downstream target regions identifies a transcriptional network linking multiple independent coronary artery disease loci. PLoS Genet. 11, e1005202 (2015).

    Article  Google Scholar 

  25. Franzen, O. et al. Cardiometabolic risk loci share downstream cis- and trans-gene regulation across tissues and diseases. Science 353, 827–830 (2016).

    Article  CAS  Google Scholar 

  26. Clement, N. et al. Notch3 and IL-1β exert opposing effects on a vascular smooth muscle cell inflammatory pathway in which NF-κB drives crosstalk. J. Cell Sci. 120, 3352–3361 (2007).

    Article  CAS  Google Scholar 

  27. Pidkovka, N. A. et al. Oxidized phospholipids induce phenotypic switching of vascular smooth muscle cells in vivo and in vitro. Circ. Res. 101, 792–801 (2007).

    Article  CAS  Google Scholar 

  28. Dandre, F. & Owens, G. K. Platelet-derived growth factor-BB and Ets-1 transcription factor negatively regulate transcription of multiple smooth muscle cell differentiation marker genes. Am. J. Physiol. Heart Circ. Physiol. 286, H2042–H2051 (2004).

    Article  CAS  Google Scholar 

  29. Rong, J. X., Shapiro, M., Trogan, E. & Fisher, E. A. Transdifferentiation of mouse aortic smooth muscle cells to a macrophage-like state after cholesterol loading. Proc. Natl Acad. Sci. USA 100, 13531–13536 (2003).

    Article  CAS  Google Scholar 

  30. Rong, J. X., Berman, J. W., Taubman, M. B. & Fisher, E. A. Lysophosphatidylcholine stimulates monocyte chemoattractant protein-1 gene expression in rat aortic smooth muscle cells. Arterioscler. Thromb. Vasc. Biol. 22, 1617–1623 (2002).

    Article  CAS  Google Scholar 

  31. Allahverdian, S., Chehroudi, A. C., McManus, B. M., Abraham, T. & Francis, G. A. Contribution of intimal smooth muscle cells to cholesterol accumulation and macrophage-like cells in human atherosclerosis. Circulation 129, 1551–1559 (2014).

    Article  CAS  Google Scholar 

  32. Kanisicak, O. et al. Genetic lineage tracing defines myofibroblast origin and function in the injured heart. Nat. Commun. 7, 12260 (2016).

    Article  CAS  Google Scholar 

  33. Chappell, J. et al. Extensive proliferation of a subset of differentiated, yet plastic, medial vascular smooth muscle cells contributes to neointimal formation in mouse injury and atherosclerosis models. Circ. Res. 119, 1313–1323 (2016).

    Article  CAS  Google Scholar 

  34. 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  Google Scholar 

  35. Satija, R., Farrell, J. A., Gennert, D., Schier, A. F. & Regev, A. Spatial reconstruction of single-cell gene expression data. Nat. Biotechnol. 33, 495–502 (2015).

    Article  CAS  Google Scholar 

  36. Kasowski, M. et al. Variation in transcription factor binding among humans. Science 328, 232–235 (2010).

    Article  CAS  Google Scholar 

  37. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    Article  CAS  Google Scholar 

  38. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008).

    Article  Google Scholar 

Download references

Acknowledgements

The authors are grateful to G. Owens for providing TgMyh11-Cre, ApoE−/− animals and experimental protocols. The authors are also grateful to L. Román for assistance with developing the human coronary digestion protocol. This work was supported by NIH grant F32HL129670 and AHA grant 18CDA34110206 (both to R.C.W.), and NIH grants K08HL133375 (to J.B.K.), R00HL125912 (to C.L.M.), R01HL109512 (to T.Q.), R01HL134817 (to T.Q.), R33HL120757 (to T.Q.), R01DK107437 (to T.Q.), R01HL139478 (to T.Q.), 1R01HL141371 (to J.C.W.) and 17GRNT33660474 AHA (to M.D.T.). The sequencing data were generated on an Illumina HiSeq 4000 that was purchased with funds from NIH under award number S10OD018220. Cell sorting was performed on an instrument in the Shared FACS Facility obtained using NIH S10 Shared Instrument Grant S10RR025518-01.

Author information

Authors and Affiliations

Authors

Contributions

R.C.W. designed and performed all scRNA-Seq experiments, analyzed the scRNA-Seq data, performed the RNAscope in situ hybridization assays, performed and analyzed the CITE-Seq and FACS experiments, analyzed the immunofluorescence data, performed the eQTL analyses, assisted with mouse colony breeding, drafted the manuscript and led the study. D.W. assisted with the design of the scRNA-Seq experiments and performed scRNA-Seq capture and library preparation for all samples. D.T.P. performed scRNA-Seq capture and helped obtain human coronary samples. J.C. assisted with the scRNA-Seq capture, library preparation and sequencing. T.N. performed qPCR experiments, analyzed the qPCR data and performed TCF21 ChIP-Seq. M.P., C.L.M., B.L. and S.B.M. performed the eQTL analyses. R.K. performed the immunohistochemistry experiments and bred the mouse colonies. M.N. performed and analyzed the immunohistochemistry experiments. K.Z., M.A. and R.C. assisted with network analysis. T.K.K., R.F. and Y.J.W. prepared the human tissue samples. M.D.T. and J.C.W. provided critical expert guidance on the manuscript. J.B.K. helped plan the mouse in situ histology studies, managed the mouse colonies, performed the TCF21 overexpression experiment, performed the quantitative immunohistochemistry analysis of lesion characteristics and conceived the study. T.Q. conceived and supervised the study. All authors discussed the results and contributed critical review to the manuscript.

Corresponding author

Correspondence to Thomas Quertermous.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information: Michael Basson was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 Design of mouse experiments.

a, Alleles present in SMClin and SMClin-KO mice. KO (knockout) refers to Tcf21, lin, lineage tracing; Tg, transgene; ΔSMC, SMC cell-specific KO. b, Mice were maintained on a chow diet from birth until 7 weeks of age, then underwent gavage and HFD treatment. For the scRNA-Seq, RNAscope, CITE-Aeq, histology involving BODIPY and FACS staining experiments (upper timeline), mice were gavaged only at 7 weeks of age, before the onset of HFD, as denoted by red arrows. For the scRNA-Seq experiments, mice were sacrificed at baseline (72 h after initial tamoxifen gavage), or after 8 or 16 weeks of HFD. For the RNAscope experiments, mice were sacrificed after either 8 or 16 weeks of HFD. For the CITE-Seq experiment, mice were sacrificed after 16 weeks of HFD. For the BODIPY studies, mice were sacrificed after 16 weeks of HFD. For the FACS staining experiment, two mice (one after 12 weeks HFD and another after 15 weeks HFD) were used. For quantitative histology experiments (lower timeline), mice were gavaged at 7 weeks of age, after 8 weeks of HFD and after 16 weeks of HFD (48 h before sacrifice), as denoted by red arrows. For these quantitative histology experiments, all mice were sacrificed after 16 weeks of HFD. c, FACS workflow for isolating single cells from the mouse aortic root.

Extended Data Fig. 2 SMC phenotypic modulation in the mouse aortic root.

ad, t-SNE visualization of cell types present in the WT mouse aortic root from all time points overlaid with the expression of Col1a1, Acta2, Sca1 and Lgals3 (n = 9 mice). eg, RNAscope staining for Lum (green) and tdT (red) in a plaque after 8 weeks of HFD (e), the non-diseased media of a mouse on 16 weeks HFD (f) and a baseline healthy aorta (g). h, RNAscope negative control. Images in eh are representative of two experiments and scale bars indicate 25 μm. i, t-SNE visualization of cell types present in the WT mouse aortic root from all time points overlaid with osteopontin (Spp1) expression (n = 9 mice). j,k, RNAscope co-localization of Spp1 (green) and tdT (red) in a plaque after 16 weeks of HFD. Yellow arrows indicate co-localization of Spp1 and tdT. l, RNAscope negative control. Images from jl are representative of four experiments, and scale bars indicate 50 μm. m, Heat map representation of the Euclidean distance between cell cluster centroids in 20-dimensional principal component space with the smallest distances in yellow and the largest distances in black. Data are after 16 weeks of HFD. n, Staining of a single-cell suspension from the atherosclerotic mouse aortic root and ascending aorta with antibodies against the macrophage markers Cd16 and Cd32, and analysis of co-expression with the tdT SMC lineage marker. Data are from one experiment and n = 2 mice (after 12 and 15 weeks of HFD). ot, Single cells from the atherosclerotic mouse aortic root and ascending aorta at 16 weeks HFD were incubated with DNA-barcoded antibodies against the macrophage markers Cd16, Cd32, Cd11b, Cd64, Cd86 and F4/80 before undergoing scRNA-Seq (CITE-Seq), yielding simultaneous transcriptomic and antibody binding data within each individual cell. o, Cell type assignments were determined with scRNA-Seq as described previously. pt, Quantitative antibody binding within each cell type. Results are from one experiment and n = 2 mice.

Extended Data Fig. 3 Additional characteristics of SMClin versus SMClin-KO mice.

a,b, Tcf21 expression in SMC lineage-labeled cells from SMClin (WT) and SMClin-KO (KO) mice from all time points combined (n = 13 mice). a, Tcf21 expression for all WT cells (left; minimum = 0; maximum = 2.55; mean = 0.071) and all KO cells (right; minimum = 0; maximum = 1.97; mean = 0.004). b, Mean Tcf21 expression visualized for all SMC lineage-labeled WT and KO cells. c, Total Lgals3+ area in the lesion is reduced in SMClin-KO mice. d, Cd68 immunohistochemistry quantification (left) and representative images (right). Scale bars represent 100 μm. e, Lesion area, normalized to the total vessel area. Data from ce are after 16 weeks HFD, and analyzed using a two-sided Student’s t-test. Error bars indicate standard error.

Source data

Extended Data Fig. 4 Joint clustering approach identifies human phenotypically modulated SMCs.

a, Joint clustering of mouse and human datasets using canonical correlation analysis (CCA) as per the Seurat package. b, The shared mouse/human cluster containing bona fide SMC lineage-traced, phenotypically modulated SMCs (fibromyocytes) from the mouse is highlighted in red. c, Mouse cells in the shared mouse/human fibromyocyte cluster in b are highlighted in the independently clustered mouse dataset, confirming their location within the known fibromyocyte cell cluster. d, Human cells in the shared mouse/human fibromyocyte cluster in b are highlighted in the independently clustered human dataset, illustrating their location predominantly in the ‘fibromyocyte’ cluster (also shown in brown in Fig. 4d). e, All joint mouse/human clusters in a were mapped back to the human dataset. Agreement is identified in cell type assignment between the joint clustering approach and the independently clustered human dataset.

Extended Data Fig. 5 Human phenotypically modulated SMCs.

a, t-SNE visualization of cell types in the right coronary artery of four patients, overlaid with LUM expression. Expression levels are indicated by scales in the lower right. b, TNFRSF11B RNAscope staining in a human coronary artery section. Hybridization events are seen as red dots. c, A negative control RNAscope probe shows no staining. The images in b and c are representative of four experiments, and the scale bars represent 50 μm. d, Heat map representation of the Euclidean distance between cell cluster centroids in 20-dimensional principal component space, with the smallest distances in yellow and the largest distances in black. The relationship between ‘fibromyocyte’ and ‘fibroblast 2’ clusters is highlighted with white asterisks. The ‘fibromyocyte’, ‘SMC’ and the main ‘macrophage’ clusters are denoted by black asterisks. e,f, t-SNE visualization of cell types in the right coronary artery of four patients overlaid with CD68 expression (e) and TCF21 expression (f). g, UCSC Genome Browser shots of representative TCF21 ChIP-Seq peaks within the PRELP and MYH11 genes, which are highly correlated and anticorrelated, respectively, with TCF21 and the fibromyocyte phenotype. Images are from one ChIP-Seq experiment.

Source data

Extended Data Fig. 6 Association of genome-wide significant CAD risk SNPs at the 6q23.2 locus with TCF21 expression.

Seven SNPs in the 6q23.2 locus were associated with CAD at genome-wide significance. The association between risk and protective genotypes and TCF21 expression for each of these SNPs was determined using the GTEx database in CAD-relevant tissues and a cohort of 52 HCASMC lines. The number of independent tissue samples included for each SNP is indicated in the GTEx data (‘N’), and n = 52 cell lines for the HCASMC data. In each box plot, the middle line represents the median, the box represents the first to third quartile range, and whiskers represent 1.5× the interquartile range.

Supplementary information

Source data

Source Data Fig. 3

Statistical source data for Fig. 3.

Source Data Fig. 5

Statistical source data for Fig. 5.

Source Data Extended Data Fig. 3

Statistical source data for Extended Data Fig. 3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wirka, R.C., Wagh, D., Paik, D.T. et al. Atheroprotective roles of smooth muscle cell phenotypic modulation and the TCF21 disease gene as revealed by single-cell analysis. Nat Med 25, 1280–1289 (2019). https://doi.org/10.1038/s41591-019-0512-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41591-019-0512-5

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