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:

Mechanosensation of cyclical force by PIEZO1 is essential for innate immunity

An Author Correction to this article was published on 12 November 2019

This article has been updated

Abstract

Direct recognition of invading pathogens by innate immune cells is a critical driver of the inflammatory response. However, cells of the innate immune system can also sense their local microenvironment and respond to physiological fluctuations in temperature, pH, oxygen and nutrient availability, which are altered during inflammation. Although cells of the immune system experience force and pressure throughout their life cycle, little is known about how these mechanical processes regulate the immune response. Here we show that cyclical hydrostatic pressure, similar to that experienced by immune cells in the lung, initiates an inflammatory response via the mechanically activated ion channel PIEZO1. Mice lacking PIEZO1 in innate immune cells showed ablated pulmonary inflammation in the context of bacterial infection or fibrotic autoinflammation. Our results reveal an environmental sensory axis that stimulates innate immune cells to mount an inflammatory response, and demonstrate a physiological role for PIEZO1 and mechanosensation in immunity.

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: PIEZO1 signalling in macrophages induces transcriptional reprogramming via HIF1α.
Fig. 2: PIEZO1 signalling stabilizes HIF1α via EDN1.
Fig. 3: PIEZO1 recognition of the cyclical hydrostatic pressure microenvironment in the lung drives inflammation via EDN1.

Similar content being viewed by others

Data availability

The authors declare that all data supporting the findings of this study are available within this article and its Supplementary Information files. RNA-sequencing data have been deposited in the Gene Expression Omnibus database under the accession code GSE133069.

Change history

  • 12 November 2019

    An Amendment to this paper has been published and can be accessed via a link at the top of the paper.

References

  1. Pritchard, M. T., Li, Z. & Repasky, E. A. Nitric oxide production is regulated by fever-range thermal stimulation of murine macrophages. J. Leukoc. Biol. 78, 630–638 (2005).

    Article  CAS  Google Scholar 

  2. Anand, R. J. et al. Hypoxia causes an increase in phagocytosis by macrophages in a HIF-1α-dependent manner. J. Leukoc. Biol. 82, 1257–1265 (2007).

    Article  CAS  Google Scholar 

  3. Ip, W. K. E. & Medzhitov, R. Macrophages monitor tissue osmolarity and induce inflammatory response through NLRP3 and NLRC4 inflammasome activation. Nat. Commun. 6, 6931 (2015).

    Article  ADS  CAS  Google Scholar 

  4. Littlewood-Evans, A. et al. GPR91 senses extracellular succinate released from inflammatory macrophages and exacerbates rheumatoid arthritis. J. Exp. Med. 213, 1655–1662 (2016).

    Article  CAS  Google Scholar 

  5. Palm, N. W. & Medzhitov, R. Pattern recognition receptors and control of adaptive immunity. Immunol. Rev. 227, 221–233 (2009).

    Article  CAS  Google Scholar 

  6. Huse, M. Mechanical forces in the immune system. Nat. Rev. Immunol. 17, 679–690 (2017).

    Article  CAS  Google Scholar 

  7. Hsiai, T. K. et al. Monocyte recruitment to endothelial cells in response to oscillatory shear stress. FASEB J. 17, 1648–1657 (2003).

    Article  CAS  Google Scholar 

  8. McWhorter, F. Y., Davis, C. T. & Liu, W. F. Physical and mechanical regulation of macrophage phenotype and function. Cell. Mol. Life Sci. 72, 1303–1316 (2015).

    Article  CAS  Google Scholar 

  9. Coste, B. et al. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science 330, 55–60 (2010).

    Article  ADS  CAS  Google Scholar 

  10. Murthy, S. E. et al. The mechanosensitive ion channel Piezo2 mediates sensitivity to mechanical pain in mice. Sci. Transl. Med. 10, eaat9897 (2018).

    Article  Google Scholar 

  11. Gudipaty, S. A. et al. Mechanical stretch triggers rapid epithelial cell division through Piezo1. Nature 543, 118–121 (2017).

    Article  ADS  CAS  Google Scholar 

  12. Wang, S. et al. Endothelial cation channel PIEZO1 controls blood pressure by mediating flow-induced ATP release. J. Clin. Invest. 126, 4527–4536 (2016).

    Article  Google Scholar 

  13. Delmas, P., Hao, J. & Rodat-Despoix, L. Molecular mechanisms of mechanotransduction in mammalian sensory neurons. Nat. Rev. Neurosci. 12, 139–153 (2011).

    Article  CAS  Google Scholar 

  14. Schipke, K. J., To, S. D. F. & Warnock, J. N. Design of a cyclic pressure bioreactor for the ex vivo study of aortic heart valves. J. Vis. Exp. (54):3316 (2011).

  15. Li, J. et al. Piezo1 integration of vascular architecture with physiological force. Nature 515, 279–282 (2014).

    Article  ADS  CAS  Google Scholar 

  16. He, Y., Hara, H. & Núñez, G. Mechanism and regulation of NLRP3 inflammasome activation. Trends Biochem. Sci. 41, 1012–1021 (2016).

    Article  CAS  Google Scholar 

  17. Tannahill, G. M. et al. Succinate is an inflammatory signal that induces IL-1β through HIF-1α. Nature 496, 238–242 (2013).

    Article  ADS  CAS  Google Scholar 

  18. Yamashita, K., Discher, D. J., Hu, J., Bishopric, N. H. & Webster, K. A. Molecular regulation of the endothelin-1 gene by hypoxia. Contributions of hypoxia-inducible factor-1, activator protein-1, GATA-2, and p300/CBP. J. Biol. Chem. 276, 12645–12653 (2001).

    Article  CAS  Google Scholar 

  19. Lee, J. J. et al. Hypoxia activates the cyclooxygenase-2-prostaglandin E synthase axis. Carcinogenesis 31, 427–434 (2010).

    Article  CAS  Google Scholar 

  20. Palazon, A., Goldrath, A. W., Nizet, V. & Johnson, R. S. HIF transcription factors, inflammation, and immunity. Immunity 41, 518–528 (2014).

    Article  CAS  Google Scholar 

  21. Varia, M. A. et al. Pimonidazole: a novel hypoxia marker for complementary study of tumor hypoxia and cell proliferation in cervical carcinoma. Gynecol. Oncol. 71, 270–277 (1998).

    Article  CAS  Google Scholar 

  22. Mekhail, K., Gunaratnam, L., Bonicalzi, M.-E. & Lee, S. HIF activation by pH-dependent nucleolar sequestration of VHL. Nat. Cell Biol. 6, 642–647 (2004).

    Article  CAS  Google Scholar 

  23. Glogowska, E. et al. Novel mechanisms of PIEZO1 dysfunction in hereditary xerocytosis. Blood 130, 1845–1856 (2017).

    Article  CAS  Google Scholar 

  24. Miyamoto, T. et al. Functional role for Piezo1 in stretch-evoked Ca2+ influx and ATP release in urothelial cell cultures. J. Biol. Chem. 289, 16565–16575 (2014).

    Article  CAS  Google Scholar 

  25. Stow, L. R., Jacobs, M. E., Wingo, C. S. & Cain, B. D. Endothelin-1 gene regulation. FASEB J. 25, 16–28 (2011).

    Article  CAS  Google Scholar 

  26. Li, M. et al. Endothelin-1 induces hypoxia inducible factor 1α expression in pulmonary artery smooth muscle cells. FEBS Lett. 586, 3888–3893 (2012).

    Article  CAS  Google Scholar 

  27. Liu, Y. V. et al. Calcineurin promotes hypoxia-inducible factor 1α expression by dephosphorylating RACK1 and blocking RACK1 dimerization. J. Biol. Chem. 282, 37064–37073 (2007).

    Article  CAS  Google Scholar 

  28. Cheng, T.-H. et al. Reactive oxygen species mediate cyclic strain-induced endothelin-1 gene expression via Ras/Raf/extracellular signal-regulated kinase pathway in endothelial cells. J. Mol. Cell. Cardiol. 33, 1805–1814 (2001).

    Article  CAS  Google Scholar 

  29. Bailis, W. et al. Distinct modes of mitochondrial metabolism uncouple T cell differentiation and function. Nature 571, 403–407 (2019).

    Article  ADS  CAS  Google Scholar 

  30. Lindsey, A. S. et al. Analysis of pulmonary vascular injury and repair during Pseudomonas aeruginosa infection-induced pneumonia and acute respiratory distress syndrome. Pulm. Circ. 9, 1–13 (2019).

    Article  Google Scholar 

  31. Novak, J., Georgakoudi, I., Wei, X., Prossin, A. & Lin, C. P. In vivo flow cytometer for real-time detection and quantification of circulating cells. Opt. Lett. 29, 77–79 (2004).

    Article  ADS  CAS  Google Scholar 

  32. Abram, C. L., Roberge, G. L., Hu, Y. & Lowell, C. A. Comparative analysis of the efficiency and specificity of myeloid-Cre deleting strains using ROSA-EYFP reporter mice. J. Immunol. Methods 408, 89–100 (2014).

    Article  CAS  Google Scholar 

  33. Mack, M. et al. Expression and characterization of the chemokine receptors CCR2 and CCR5 in mice. J. Immunol. 166, 4697–4704 (2001).

    Article  CAS  Google Scholar 

  34. Phillips, J. E. et al. Bleomycin induced lung fibrosis increases work of breathing in the mouse. Pulm. Pharmacol. Ther. 25, 281–285 (2012).

    Article  CAS  Google Scholar 

  35. Ashcroft, T., Simpson, J. M. & Timbrell, V. Simple method of estimating severity of pulmonary fibrosis on a numerical scale. J. Clin. Pathol. 41, 467–470 (1988).

    Article  CAS  Google Scholar 

  36. Wu, J. et al. Inactivation of mechanically activated Piezo1 ion channels is determined by the C-terminal extracellular domain and the inner pore helix. Cell Rep. 21, 2357–2366 (2017).

    Article  CAS  Google Scholar 

  37. Abshire, M. Y., Thomas, K. S., Owen, K. A. & Bouton, A. H. Macrophage motility requires distinct α5β1/FAK and α4β1/paxillin signaling events. J. Leukoc. Biol. 89, 251–257 (2011).

    Article  CAS  Google Scholar 

  38. Young, S. R. L., Gerard-O’Riley, R., Kim, J.-B. & Pavalko, F. M. Focal adhesion kinase is important for fluid shear stress-induced mechanotransduction in osteoblasts. J. Bone Miner. Res. 24, 411–424 (2009).

    Article  CAS  Google Scholar 

  39. Bassotti, G. et al. Gastrointestinal motility disorders in inflammatory bowel diseases. World J. Gastroenterol. 20, 37–44 (2014).

    Article  Google Scholar 

  40. Rajamäki, K. et al. Extracellular acidosis is a novel danger signal alerting innate immunity via the NLRP3 inflammasome. J. Biol. Chem. 288, 13410–13419 (2013).

    Article  Google Scholar 

Download references

Acknowledgements

We thank J. Alderman, C. Lieber, C. Hughes, E. H.-P. and P. Rainey for help in facilitating this work; B. Kazmierczak for providing the P. aeruginosa; M. Roulis for insightful comments and reagents; L. Orr for helpful insight regarding statistical analysis; I. Odell for help with computational software; C. Rothlin and Dr. C. Abraham for continued support and feedback as this work progressed; P.-M. Chen for help and reagents related to hypoxia studies. This work was supported by the Howard Hughes Medical Institute and the Blavatnik Family Foundation (R.A.F.). This work was supported in part by the Searle Scholars Program, the Leukemia Research Foundation, the Gruber Foundation and the NIH (R01GM122984). R.J. was supported in part by the Crohn’s and Colitis Foundation. M.A.S. was supported by the NIH (HL RO1 75092). A.G.S. was supported in part by an NIH training grant (T32 GM007499) and the American Society for Microbiology.

Author information

Authors and Affiliations

Authors

Contributions

A.G.S. designed and performed experiments, collected and analysed data, and wrote the manuscript. P.B. performed microbiological experiments and offered vital conceptual insight. H.R.S. developed reagents and performed Cas9 experiments. L.S. performed in vivo fibrosis experiments. C.C.D.H. performed all bioinformatic analysis. S.Y. performed in vitro shear stress experiments. M.R.d.Z. offered conceptual insight. J.N.W. and S.D.F.T. designed and built the bioreactor and software necessary to complete mechanistic experiments. A.G.Y. helped collect samples. M.M. provided critical reagents and advice on experimental design. M.A.S. and C.S.D.C. provided intellectual support and resources. N.W.P. originally proposed the study. R.J. and R.A.F supervised the project, helped interpret the work and supervised writing of the manuscript.

Corresponding authors

Correspondence to Ruaidhrí Jackson or Richard A. Flavell.

Ethics declarations

Competing interests

: R.A.F. is a scientific advisor to GlaxoSmithKline, a consultant for Hatteras Venture Partners and a shareholder and consultant for Zai Lab Ltd. All other authors declare no competing interests.

Additional information

Peer review information Nature thanks Seth Alper, Luke O’Neill, Sarah Walmsley and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Fig. 1 Pressure chamber schematic and PIEZO1-knockout validation.

a, Schematic of CHP bioreactor showing side view. b, Graph of pressure regimes used within the pressure chamber. c, RT–qPCR analysis of Piezo1 in unstimulated BMDMs from Piezo1fl/fl (WT) and Piezo1ΔLysM mice. Data are presented as five biological replicates from two independent experiments. d, RT–qPCR analysis of Piezo1fl/fl BMDMs treated with static pressure at the indicated magnitude or with CHP for 6 h. Data are presented as four biological replicates from two independent experiments. e, Immunoblot analysis of HIF1α and β-tubulin from Piezo1fl/fl BMDMs treated with static pressure at the indicated magnitude or with CHP for 6 h. Data are representative of two independent experiments. f, ELISA of IL-1β from Piezo1fl/fl and Piezo1ΔLysM BMDMs treated with CHP alone (NT), 10 ng ml−1 LPS with 5 h CHP followed by 1 h CHP, 5 h CHP followed by 1 h nigericin (10 μM) in CHP, or 10 ng ml−1 LPS for 5 h followed by 1 h nigericin (10 μM) with CHP. Data are presented as three biological replicates. Data in c, d, f are mean ± s.e.m.; significance is determined by unpaired two-tailed t-test (*P < 0.05, **P < 0.01, ***P < 0.001 and ****P < 0.0001).

Extended Data Fig. 2 HIF1α stabilization is independent of hypoxia during CHP stimulation.

a, Representative confocal microscopy of BMDMs treated for 6 h in either normoxic conditions or in a hypoxic chamber (2% O2). Cells were treated with pimonidazole (20 μM) for 1 h before completion of treatment. Upon completion, cells were fixed, permeabilized and stained with monoclonal antibody against pimonidazole. Data are representative of two independent experiments. b, Confocal microscopy of BMDMs treated for 6 h in either static, CHP or hypoxic chamber (0.5% O2). Cells were treated with 20 μM pimonidazole 1 h before completion of treatment. Upon completion, cells were fixed, permeabilized and stained with monoclonal antibody against pimonidazole. Data are representative of two independent experiments. c, RT–qPCR analysis of BMDMs treated with cyclical hydrostatic pressure, 2% O2 or cyclical hydrostatic pressure with 2% O2 for 6 h. Data are presented as three biological replicates. d, RT–qPCR analysis of Hif1a in BMDMs following CHP. Data are presented as four biological replicates. Data are from two independent experiments. e, Representative immunoblot analysis of HIF2α and β-tubulin from BMDMs treated with static pressure, CHP or DMOG (200 μM) for 6 h. Data in c, d are mean ± s.e.m.; significance is determined by unpaired two-tailed t-test.

Extended Data Fig. 3 Cyclical hydrostatic pressure-induced HIF1α protein is stabilized independent of acidity.

a, Intracellular pH measurements using pHrodo intracellular pH indicator dye for final 30 min of treatment. Piezo1fl/fl (n = 3) and Piezo1ΔLysM (n = 3) cells were treated with CHP for the indicated periods or with low-pH cell medium for 6 h. b, pH measurements of supernatant from Piezo1fl/fl (n = 3) and Piezo1ΔLysM (n = 3) BMDM culture treated with CHP for the indicated periods. Data in a, b, are mean ± s.e.m.; significance is determined by unpaired two-tailed t-test.

Extended Data Fig. 4 Kinetics and sufficiency of cyclical hydrostatic pressure induced HIF1α stabilization and transcriptional response.

a, Representative immunoblot analysis of HIF1α and β-tubulin from BMDMs in CHP with 5 μM GsMTx4 (Gsm) for 6 h, then washed three times in DMEM before further CHP stimulation. Data are representative of two independent experiments. b, Immunoblot analysis of HIF1α and β-tubulin from BMDMs in CHP for 6 h, pretreated with 10 μM BAPTA-AM for 30 min. Data are representative of two independent experiments. c, Immunoblot analysis of HIF1α and β-tubulin in BMDMs. Piezo1fl/fl (WT) or Piezo1ΔLysM (KO) were cultured in CHP for 6 h along with 3 μg ml BFA. Cells were then washed with DMEM three times and further subjected to additional CHP stimulation. Supernatant (sup) was then clarified and transferred to WT or KO BMDMs cultured in static pressure for two additional hours. Data are representative of two independent experiments. d, Representative immunoblot analysis of HIF1α and β-tubulin from BMDMs treated with 100 μM chloroquine (ChlQ) or 50 μM MG132 for 6 h with CHP. Data are representative of two independent experiments. e, RT–qPCR analysis of Edn1 from Hif1afl/fl and Piezo1ΔLysM BMDMs treated with CHP for 1 h or 6 h. Data are presented as three biological replicates. f, g, RT–qPCR analysis of Piezo1fl/fl (f) or Piezo1ΔLysM (g) BMDMs cultured in 6 h of treatment 200 μM DMOG, 6 h of CHP, 2 h of CHP followed by 4 h of no treatment (NT), 2 h of static pressure followed by 4 h of 200 μM DMOG, or 2 h CHP followed by 4 h of 200 μM DMOG. Data are representative of three biological replicates. Data are mean ± s.e.m.

Extended Data Fig. 5 Bacterial infection has no effect on MSIC expression.

ac, RT–qPCR analysis of known mammalian mechanosensory ion channels from monocytes (a), alveolar macrophages (b) or neutrophils (c) sorted from lungs of mice infected with P. aeruginosa or PBS for 6 h. Data are represented as fold increase over neutrophil expression for each gene. Data are representative of four biological replicates. Data are mean ± s.e.m.

Extended Data Fig. 6 PIEZO1-dependent cyclical hydrostatic pressure stabilizes HIF1α in the lung.

a, Representative flow cytometry of lungs from Piezo1fl/fl (top; n = 9) or Piezo1ΔLysM (bottom; n = 9) mice infected intranasally with P. aeruginosa for 6 h. The lymphoid lineage is TCRβ+NK1.1+B220+CD19+CD90.2+TCRγδ+. Data are representative of two independent experiments. b, c, Infiltrating monocyte quantification in terms of percentage(b) and total numbers (c) following 6 h of intranasal P. aeruginosa infection of Piezo1fl/fl (n = 7) or Piezo1ΔLysM (n = 7) mice. Data are representative of two independent experiments. d, Representative intracellular flow cytometry histograms of HIF1α from total lung interstitial monocytes of Piezo1fl/fl and Piezo1ΔLysM mice infected intranasally with P. aeruginosa for 6 h. Data are from two independent experiments. e, MFI of HIF1α from in vivo-labelled CD45.2+ circulatory monocytes from Piezo1fl/fl (n = 7) or Piezo1ΔLysM (n = 6) mice infected with P. aeruginosa for 6 h. Data are from two independent experiments. f, MFI of HIF1α from tissue-infiltrated and circulatory monocytes from mice infected with P aeruginosa intranasally (n = 7) or intraperitoneally (n = 8) for 6 h. Data are from two independent experiments. g, RT–qPCR analysis from BMDMs treated with oscillatory shear stress for 6 h. Data are representative of three biological replicates. Data in b, c, eg are mean ± s.e.m.; significance is determined by unpaired two-tailed t-test (****P < 0.0001).

Extended Data Fig. 7 EDN1 signalling confers protection to P aeruginosa infection.

a, CFUs from liver of Piezo1fl/fl mice infected intranasally with P. aeruginosa for 24 h, pre-treated with 25 μg EDN1-blocking antibody (n = 5) or isotype control (n = 6) intranasally 6 h before infection. Data are from two independent experiments. b, CFUs from liver of Piezo1ΔLysM mice infected intranasally with P. aeruginosa for 24 h treated with vehicle (n = 10) or 10 μg rEDN1 (n = 9) at the time of infection. Data are from two independent experiments. c, d, CFUs from lung(c) or liver(d) of Piezo1fl/fl mice infected intranasally with P. aeruginosa for 24 h, pre-treated with vehicle (n = 10) or 100 mg per kg (body weight) bosentan (n = 8) intraperitoneally 3 h before infection. Data are from two independent experiments. Data are mean ± s.e.m.; significance is determined by unpaired two-tailed Mann–Whitney U-test (*P < 0.05,**P < 0.01 and ***P < 0.001).

Extended Data Fig. 8 PIEZO1 recognition of cyclical force drives a HIF1α proinflammatory program.

Cyclical force signals via PIEZO1 in myeloid cells, resulting in Ca2+ influx and AP-1-induced EDN1 expression. Ednrb signalling then drives HIF1α stabilization and proinflammatory transcriptional upregulation.

Extended Data Fig. 9 Infiltrating monocytes recognize cyclical hydrostatic pressure in the lung via PIEZO1 to trigger neutrophil-mediated bacterial clearance.

Recruited monocytes recognize cyclical force via PIEZO1 in the lung and secrete EDN1 to drive HIF1α stabilization and CXCL2 expression to induce neutrophilia and bacterial clearance.

Supplementary information

Supplementary Information

This file contains the uncropped gels for Figs. 1 and 2 and extended data 1, 2 and 4.

Reporting Summary

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Solis, A.G., Bielecki, P., Steach, H.R. et al. Mechanosensation of cyclical force by PIEZO1 is essential for innate immunity. Nature 573, 69–74 (2019). https://doi.org/10.1038/s41586-019-1485-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1485-8

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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