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:

Investigating the nature of active forces in tissues reveals how contractile cells can form extensile monolayers

An Author Correction to this article was published on 09 March 2021

This article has been updated

Abstract

Actomyosin machinery endows cells with contractility at a single-cell level. However, within a monolayer, cells can be contractile or extensile based on the direction of pushing or pulling forces exerted by their neighbours or on the substrate. It has been shown that a monolayer of fibroblasts behaves as a contractile system while epithelial or neural progentior monolayers behave as an extensile system. Through a combination of cell culture experiments and in silico modelling, we reveal the mechanism behind this switch in extensile to contractile as the weakening of intercellular contacts. This switch promotes the build-up of tension at the cell–substrate interface through an increase in actin stress fibres and traction forces. This is accompanied by mechanotransductive changes in vinculin and YAP activation. We further show that contractile and extensile differences in cell activity sort cells in mixtures, uncovering a generic mechanism for pattern formation during cell competition, and morphogenesis.

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: Changes from extensile to contractile behaviours in the absence of E-cadherin.
Fig. 2: Inter- and intracellular stresses dictate extensile and contractile behaviours.
Fig. 3: Knocking out E-cadherin increases cell–substrate interactions.
Fig. 4: E-cadherin removal triggers mechanotransductive changes within the monolayer.
Fig. 5: Cell sorting triggered by change in nematic behaviour of monolayers.
Fig. 6: Cell sorting is governed by activity of the system.

Similar content being viewed by others

Data availability

Data for cell orientation detection, and Bayesian inference of tissue stress are available upon request from the corresponding author. Source data are provided with this paper.

Code availability

Codes for cell orientation detection, and Bayesian inference of tissue stress are available upon request from the corresponding author. Numerical analyses were performed using a custom-made code: CELADRO (Cells as Active Droplets), which is open source and has been deposited on GitHub (https://github.com/rhomu/celadro).

Change history

References

  1. Barriga, E. H., Franze, K., Charras, G. & Mayor, R. Tissue stiffening coordinates morphogenesis by triggering collective cell migration in vivo. Nature 554, 523–527 (2018).

    Article  CAS  Google Scholar 

  2. Onodera, T. et al. Btbd7 regulates epithelial cell dynamics and branching morphogenesis. Science 329, 562–565 (2010).

    Article  CAS  Google Scholar 

  3. Chen, X. L. et al. VEGF-induced vascular permeability is mediated by FAK. Dev. Cell 22, 146–157 (2012).

    Article  CAS  Google Scholar 

  4. Malinverno, C. et al. Endocytic reawakening of motility in jammed epithelia. Nat. Mater. 16, 587–596 (2017).

    Article  CAS  Google Scholar 

  5. Cetera, M. et al. Epithelial rotation promotes the global alignment of contractile actin bundles during Drosophila egg chamber elongation. Nat. Commun. 5, 5511 (2014).

    Article  CAS  Google Scholar 

  6. Wang, S., Matsumoto, K. & Yamada, K. M. Reconstituting stratified epithelial branching morphogenesis by engineering cell adhesion. Preprint at bioRxiv https://doi.org/10.1101/2020.06.24.165795v1 (2020).

  7. De Pascalis, C. & Etienne-Manneville, S. Single and collective cell migration: the mechanics of adhesions. Mol. Biol. Cell 28, 1833–1846 (2017).

    Article  CAS  Google Scholar 

  8. Schwarz, U. S. & Safran, S. A. Elastic interactions of cells. Phys. Rev. Lett. 88, 048102 (2002).

    Article  CAS  Google Scholar 

  9. Duclos, G., Erlenkämper, C., Joanny, J.-F. & Silberzan, P. Topological defects in confined populations of spindle-shaped cells. Nat. Phys. 13, 58–62 (2017).

    Article  CAS  Google Scholar 

  10. Saw, T. B. et al. Topological defects in epithelia govern cell death and extrusion. Nature 544, 212–216 (2017).

    Article  CAS  Google Scholar 

  11. Blanch-Mercader, C. et al. Turbulent dynamics of epithelial cell cultures. Phys. Rev. Lett. 120, 208101 (2018).

    Article  CAS  Google Scholar 

  12. Kawaguchi, K., Kageyama, R. & Sano, M. Topological defects control collective dynamics in neural progenitor cell cultures. Nature 545, 327–331 (2017).

    Article  CAS  Google Scholar 

  13. Xi, W., Saw, T. B., Delacour, D., Lim, C. T. & Ladoux, B. Material approaches to active tissue mechanics. Nat. Rev. Mater. 4, 23–44 (2019).

    Article  Google Scholar 

  14. Gruler, H., Dewald, U. & Eberhardt, M. Nematic liquid crystals formed by living amoeboid cells. Eur. Phys. J. B 11, 187–192 (1999).

    Article  CAS  Google Scholar 

  15. Ladoux, B. & Mège, R.-M. Mechanobiology of collective cell behaviours. Nat. Rev. Mol. Cell Biol. 18, 743–757 (2017).

    Article  CAS  Google Scholar 

  16. Giomi, L., Bowick, M. J., Mishra, P., Sknepnek, R. & Cristina Marchetti, M. Defect dynamics in active nematics. Philos. Trans. R. Soc. A 372, 20130365 (2014).

    Article  CAS  Google Scholar 

  17. Doostmohammadi, A., Ignés-Mullol, J., Yeomans, J. M. & Sagués, F. Active nematics. Nat. Commun. 9, 3246 (2018).

    Article  CAS  Google Scholar 

  18. Marth, W. & Voigt, A. Signaling networks and cell motility: a computational approach using a phase field description. J. Math. Biol. 69, 91–112 (2014).

    Article  Google Scholar 

  19. Mueller, R., Yeomans, J. M. & Doostmohammadi, A. Emergence of active nematic behavior in monolayers of isotropic cells. Phys. Rev. Lett. 122, 048004 (2019).

    Article  CAS  Google Scholar 

  20. Ng, M. R., Besser, A., Brugge, J. S. & Danuser, G. Mapping the dynamics of force transduction at cell–cell junctions of epithelial clusters. eLife 3, e03282 (2014).

    Article  Google Scholar 

  21. Nier, V. et al. Inference of internal stress in a cell monolayer. Biophys. J. 110, 1625–1635 (2016).

    Article  CAS  Google Scholar 

  22. Petitjean, L. et al. Velocity fields in a collectively migrating epithelium. Biophys. J. 98, 1790–1800 (2010).

    Article  CAS  Google Scholar 

  23. Garcia, S. et al. Physics of active jamming during collective cellular motion in a monolayer. Proc. Natl Acad. Sci. USA 112, 15314–15319 (2015).

    Article  CAS  Google Scholar 

  24. Goodwin, K. et al. Cell–cell and cell–extracellular matrix adhesions cooperate to organize actomyosin networks and maintain force transmission during dorsal closure. Mol. Biol. Cell 28, 1301–1310 (2017).

    Article  CAS  Google Scholar 

  25. Dupont, S. et al. Role of YAP/TAZ in mechanotransduction. Nature 474, 179–183 (2011).

    Article  CAS  Google Scholar 

  26. Giannone, G. et al. Periodic lamellipodial contractions correlate with rearward actin waves. Cell 116, 431–443 (2004).

    Article  CAS  Google Scholar 

  27. Zemel, A., Rehfeldt, F., Brown, A. E. X., Discher, D. E. & Safran, S. A. Optimal matrix rigidity for stress-fibre polarization in stem cells. Nat. Phys. 6, 468–473 (2010).

    Article  CAS  Google Scholar 

  28. Gupta, M. et al. Adaptive rheology and ordering of cell cytoskeleton govern matrix rigidity sensing. Nat. Commun. 6, 7525 (2015).

    Article  CAS  Google Scholar 

  29. Saraswathibhatla, A. & Notbohm, J. Tractions and stress fibers control cell shape and rearrangements in collective cell migration. Phys. Rev. X 10, 011016 (2020).

    CAS  Google Scholar 

  30. Riveline, D. et al. Focal contacts as mechanosensors. J. Cell Biol. 153, 1175–1186 (2001).

    Article  CAS  Google Scholar 

  31. Yonemura, S., Wada, Y., Watanabe, T., Nagafuchi, A. & Shibata, M. α-Catenin as a tension transducer that induces adherens junction development. Nat. Cell Biol. 12, 533–542 (2010).

    Article  CAS  Google Scholar 

  32. Bays, J. L. & DeMali, K. A. Vinculin in cell–cell and cell–matrix adhesions. Cell. Mol. Life Sci. 74, 2999–3009 (2017).

    Article  CAS  Google Scholar 

  33. Halder, G., Dupont, S. & Piccolo, S. Transduction of mechanical and cytoskeletal cues by YAP and TAZ. Nat. Rev. Mol. Cell Biol. 13, 591–600 (2012).

    Article  CAS  Google Scholar 

  34. Benham-Pyle, B. W., Pruitt, B. L. & Nelson, W. J. Mechanical strain induces E-cadherin-dependent Yap1 and β-catenin activation to drive cell cycle entry. Science 348, 1024–1027 (2015).

    Article  CAS  Google Scholar 

  35. Kim, N.-G., Koh, E., Chen, X. & Gumbiner, B. M. E-cadherin mediates contact inhibition of proliferation through Hippo signaling-pathway components. Proc. Natl Acad. Sci. USA 108, 11930–11935 (2011).

    Article  CAS  Google Scholar 

  36. Discher, D. E., Janmey, P. & Wang, Y.-l Tissue cells feel and respond to the stiffness of their substrate. Science 310, 1139–1143 (2005).

    Article  CAS  Google Scholar 

  37. Maître, J.-L. et al. Adhesion functions in cell sorting by mechanically coupling the cortices of adhering cells. Science 338, 253–256 (2012).

    Article  CAS  Google Scholar 

  38. Steinberg, M. S. Does differential adhesion govern self-assembly processes in histogenesis? Equilibrium configurations and the emergence of a hierarchy among populations of embryonic cells. J. Exp. Zool. 173, 395–433 (1970).

    Article  CAS  Google Scholar 

  39. Manning, M. L., Foty, R. A., Steinberg, M. S. & Schoetz, E.-M. Coaction of intercellular adhesion and cortical tension specifies tissue surface tension. Proc. Natl Acad. Sci. USA 107, 12517–12522 (2010).

    Article  CAS  Google Scholar 

  40. Sahu, P. et al. Small-scale demixing in confluent biological tissues. Soft Matter 16, 3325–3337 (2020).

    Article  CAS  Google Scholar 

  41. Krajnc, M. Solid-fluid transition and cell sorting in epithelia with junctional tension fluctuations. Soft Matter 16, 3209–3215 (2020).

    Article  CAS  Google Scholar 

  42. Thomas, W. A., Thomson, J., Magnani, J. L. & Steinberg, M. S. Two distinct adhesion mechanisms in embryonic neural retina cells: III. Functional specificity. Dev. Biol. 81, 379–385 (1981).

    Article  CAS  Google Scholar 

  43. Foty, R. A. & Steinberg, M. S. Cadherin-mediated cell-cell adhesion and tissue segregation in relation to malignancy. Int. J. Dev. Biol. 48, 397–409 (2004).

    Article  CAS  Google Scholar 

  44. Niessen, C. M. & Gumbiner, B. M. Cadherin-mediated cell sorting not determined by binding or adhesion specificity. J. Cell Biol. 156, 389–400 (2002).

    Article  CAS  Google Scholar 

  45. Maroudas-Sacks, Y. et al. Topological defects in the nematic order of actin fibers as organization centers of Hydra morphogenesis. Nat. Phys. https://doi.org/10.1038/s41567-020-01083-1 (2020).

  46. Comelles, J. et al. Epithelial colonies in vitro elongate through collective effects. eLife 10, e57730 (2020).

    Article  Google Scholar 

  47. Morales-Navarrete, H. et al. Liquid-crystal organization of liver tissue. eLife 8, e44860 (2019).

    Article  CAS  Google Scholar 

  48. Aigouy, B. et al. Cell flow reorients the axis of planar polarity in the wing epithelium of Drosophila. Cell 142, 773–786 (2010).

    Article  CAS  Google Scholar 

  49. Hannezo, E., Prost, J. & Joanny, J.-F. Theory of epithelial sheet morphology in three dimensions. Proc. Natl Acad. Sci. USA 111, 27–32 (2014).

    Article  CAS  Google Scholar 

  50. Doostmohammadi, A. et al. Celebrating soft matter’s 10th anniversary: cell division: a source of active stress in cellular monolayers. Soft Matter 11, 7328–7336 (2015).

    Article  CAS  Google Scholar 

  51. Marthiens, V., Padilla, F., Lambert, M. & Mege, R. M. Complementary expression and regulation of cadherins 6 and 11 during specific steps of motoneuron differentiation. Mol. Cell. Neurosci. 20, 458–475 (2002).

    Article  CAS  Google Scholar 

  52. Glukhova, M. A., Frid, M. G. & Koteliansky, V. E. Developmental changes in expression of contractile and cytoskeletal proteins in human aortic smooth muscle. J. Biol. Chem. 265, 13042–13046 (1990).

    Article  CAS  Google Scholar 

  53. Vedula, S. R. K. et al. Epithelial bridges maintain tissue integrity during collective cell migration. Nat. Mater. 13, 87–96 (2014).

    Article  CAS  Google Scholar 

  54. Hara, Y., Shagirov, M. & Toyama, Y. Cell boundary elongation by non-autonomous contractility in cell oscillation. Curr. Biol. 26, 2388–2396 (2016).

    Article  CAS  Google Scholar 

  55. Meijering, E., Dzyubachyk, O. & Smal, I. Methods for cell and particle tracking. Meth. Enzymol. 504, 183–200 (2012).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  57. Plotnikov, S. V., Sabass, B., Schwarz, U. S. & Waterman, C. M. High-resolution traction force microscopy. Meth. Cell Biol. 123, 367–394 (2014).

    Article  Google Scholar 

  58. Tseng, Q. et al. Spatial organization of the extracellular matrix regulates cell–cell junction positioning. Proc. Natl Acad. Sci. USA 109, 1506–1511 (2012).

    Article  CAS  Google Scholar 

  59. Peyret, G. et al. Sustained oscillations of epithelial cell sheets. Biophys. J. 117, 464–478 (2019).

    Article  CAS  Google Scholar 

  60. Aronson, I. S. (ed.) Physical Models of Cell Motility. Biological and Medical Physics, Biomedical Engineering (Springer, 2016).

  61. Camley, B. A. & Rappel, W.-J. Physical models of collective cell motility: from cell to tissue. J. Phys. D 50, 113002 (2017).

  62. Palmieri, B., Bresler, Y., Wirtz, D. & Grant, M. Multiple scale model for cell migration in monolayers: elastic mismatch between cells enhances motility. Sci. Rep. 5, 11745 (2015).

    Article  Google Scholar 

  63. Winkler, B., Aranson, I. S. & Ziebert, F. Confinement and substrate topography control cell migration in a 3D computational model. Commun. Phys. 2, 82 (2019).

    Article  CAS  Google Scholar 

  64. Alert, R. & Trepat, X. Physical models of collective cell migration. Annu. Rev. Condens. Matter Phys. 11, 77–101 (2020).

  65. Smeets, B. et al. Emergent structures and dynamics of cell colonies by contact inhibition of locomotion. Proc. Natl Acad. Sci. USA 113, 14621–14626 (2016).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the European Research Council (grant no. CoG-617233), LABEX Who Am I? (ANR-11-LABX-0071), the Ligue Contre le Cancer (Equipe labellisée), and the Agence Nationale de la Recherche (‘POLCAM’ (ANR-17-CE13-0013 and ‘MechanoAdipo’ ANR-17-CE13-0012). We acknowledge the ImagoSeine core facility of the IJM, member of IBiSA and France-BioImaging (ANR-10-INBS-04) infrastructures. A.D. acknowledges support from the Novo Nordisk Foundation (grant no. NNF18SA0035142), Villum Fonden (grant no. 29476), and funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement no. 847523 (INTERACTIONS). L.B. has received funding from the European Union’s Horizon 2020 research and innovation programme (Marie Sklodowska-Curie grant agreement no. 665850-INSPIRE). T.B.S. acknowledges support from a Lee Kuan Yew (LKY) postdoctoral fellowship and the Singapore Ministry of Education Academic Research Fund Tier 1 (R-397-000-320-114). S.G. and A.Y. were supported by project grants and fellowships from the National Health and Medical Research Council of Australia (1123816 and 1139592) and Australian Research Council (DP190102871). We would like to thank P. Marcq for help with implementation of BISM code, M. A. Glukhova for providing the vinculin antibody and S. Robine for the ZO1 antibody. A.D. and J.M.Y. acknowledge G. Zhang for helpful discussions regarding the model. We also thank the members of cell adhesion and mechanics team at Institut Jacques Monod, M. Piel and F. Gallet, for insightful discussions.

Author information

Authors and Affiliations

Authors

Contributions

L.B., T.B.S., A.D., J.Y., R.M.M. and B.L. designed the research. G.H.N.S.N. and T.D. developed the E-cadherin KO cell line. L.B. performed experiments and analysed the results. S.S. helped in the PA gel experiments. M.T. performed and quantified laser ablation experiments. S.G. performed the MCF7 experiments. A.D. and R.M. performed the in silico simulations. T.B.S. contributed to the analysis tools. A.S.Y., Y.T., R.M.M., A.D., J.Y. and B.L supervised the project.

Corresponding authors

Correspondence to Amin Doostmohammadi, René-Marc Mège, Julia M. Yeomans or Benoît Ladoux.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review informationNature Materials thanks the anonymous reviewers for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 MDCK WT behave as an extensile system.

a, Kymograph of a short junction (<10μm) (top) and long junction (>15μm) (bottom) before and after laser ablation. b, Recoil velocity after laser ablation for short (<10μm) (n=9) (N=4), medium (10-15μm) (n=13)(N=4) and long junctions (> 15μm) (n=12) (N=6). n, is the number of junctions ablated and N is the number of independent experiments from which these results were obtained. Error bars represent the standard deviation. ANOVA test was performed leading to *p<0.05, **p<0.01, ***p<0.001 and ****p<0.0001. Scale bars, 20μm.

Source data

Extended Data Fig. 2 MDCK WT behaves as an extensile system and MDCK E-cadherin KO behave as a contractile system.

a,b, Orientation field (left) and velocity vectors (right) around a single comet-shaped (+1/2) defect (a) and trefoil (-1/2) defect obtained from WT (top) and E-cadherin KO (bottom) monolayers. c,d, Trajectory of several comet (+1/2) (left) and trefoil (-1/2) (right) shaped defects obtained from MDCK WT (c) and MDCK E-cadherin KO (d) monolayers. Scale bars: 40μm.

Extended Data Fig. 3 Fibroblasts behave as a 2D contractile active nematic.

a, Average yy- and xy-components of strain rate map around +1/2 defect obtained from experiments (left and middle respectively) and corresponding average flow field (right) (n = 1489 defects from 2 independent experiments) for NIH3T3 cells. Colour code is positive for stretching and negative for shrinkage. b, Average yy (left)-, xy (middle)- and isotropic (right) components of stress around a +1/2 defect obtained from experiments for NIH3T3 (n = 1,428 defects from 2 independent experiments).

Extended Data Fig. 4 Characterization of MDCK E-cadherin KO cells and extensile/contractile behaviour of MCF7 cells.

a, Immunofluorescence staining (top) of E-cadherin (left), β-catenin (middle) and ZO1 (right), along with a b) representative western blot and quantification for E-cadherin (left) (n=3) and β-catenin (right) (n=3). Scale bars, 20μm. c, Western blot analysis of total MLC and quantification from 3 independent experiments normalized to α-tubulin. d, Western blot analysis of vinculin and quantification from 3 independent experiments normalized to GAPDH. Error bars represent the standard deviation. e, Western blot showing the reduced level of E-cadherin in siRNA generated E-cadherin KD cell line for MCF7 cells. f,g Average flow field for MCF7 control cells (n = 2047 defects from 3 independent experiments) (f) and siRNA E-cadherin KD MCF7 cells (n = 1256 defects from 3 independent experiments) (f). h, Uncropped blots of all the western blots shown so far.

Source data

Extended Data Fig. 5 E-cadherin rescue changes the behaviour to a 2D extensile active nematic liquid crystal and flow field around trefoil shaped defects.

a, Average yy- and xy-components of strain rate map around comet (+1/2) defect obtained from experiments (left and middle respectively) and corresponding average flow field (right) (n = 1767 defects from 2 independent experiments) for MDCK E-cadherin KO cells rescued with E-cadherin GFP. b, Total number of defects obtained per 0.55mm2 as a function of time on MDCK WT and MDCK E-cadherin KO monolayers. (n=10) from 2 independent experiments. c, Average vorticity and velocity field around trefoil (-1/2) defects in WT (left) (n=1934) and E-cadherin KO (right) (n=2028) monolayers. d, Average vorticity and velocity field around trefoil (-1/2) defects in control (left) (n=3200) and condition without active intercellular forces (right)(n=3200) monolayers obtained from simulations. e, Mean square displacement (MSD) plotted against time lag for comet (+1/2) and trefoil (-1/2) defects obtained from MDCK WT and MDCK E-cadherin KO monolayers (n=11). Error bars represent the standard deviation.

Source data

Extended Data Fig. 6 E-cadherin removal does not affect the contractile behaviour of single cells.

a, Mean traction force for both MDCK WT (n=31) and MDCK E-cadherin KO cells (n=27). b, Average yy- and xy-components of strain rate map around comet (+1/2) defect obtained from experiments (left and middle respectively) and corresponding average velocity flow field (right) (n = 1428 defects from 2 independent experiments) for MDCK E-cadherin KO cells plated on PDMS substrates of stiffness 15kPa from which stress maps were obtained in Fig. 3b’.

Source data

Extended Data Fig. 7 Drug treatment changes 2D active nematic behaviour of MDCK E-cadherin KO cells.

a,b,c, Average yy- and xy-components of strain rate map around +1/2 defect obtained from experiments (left and middle respectively) and corresponding average velocity flow field (right) for MDCK E-cadherin KO cells treated with 5μM blebbistatin (a) (n = 2174 defects from 2 independent experiments), 20μM blebbistatin (b) (n = 1223 defects from 2 independent experiments), and 25μM Y27632 (c) (n = 1965 defects from 2 independent experiments). d,e, Average yy- and xy components of strain rate map around +1/2 defect obtained from experiments (left and middle respectively) and corresponding average velocity flow field (right) for MDCK WT cells treated with 20μM blebbistatin (d) (n = 1287 defects from 2 independent experiments), and 25μM Y27632 (e) (n = 2472 defects from 2 independent experiments).

Extended Data Fig. 8 Colocalization of vinculin and paxillin.

a, Immunostaining of basal plane of vinculin (left), paxillin (middle) and merge (right) in MDCK WT (top) and MDCK E-cadherin KO monolayers. b, Intensity of vinculin plotted against paxillin for n=15 focal adhesions in MDCK WT and n=16 focal adhesions in MDCK E-cadherin KO monolayers. Scale bars: 20μm.

Extended Data Fig. 9 Substrate rigidity alters E-cadherin KO behaviour.

a,b, Average yy- and xy-components of strain rate map around +1/2 defect obtained from experiments (left and middle respectively) and corresponding average flow field (right) for MDCK WT cells (a) (n = 1426 defects from 2 independent experiments) and E-cadherin KO cells (b) (n = 1041 defects from 2 independent experiments).

Source data

Extended Data Fig. 10 Phase diagram on activity change and activity based cell sorting in a mixed culture of MDCK WT and MDCK E-cadherin KO.

a, Phase diagram showing the transition of extensile and contractile behaviour with varying values of intercellular and intracellular stresses obtained from simulations. b, Phase separation (demixing) observed from simulations where the contractile particles (orange) are surrounded by extensile particles (green). c, Cell sorting (demixing) observed for a mixture of MDCK WT and MDCK E-cadherin KO cells, where WT cells are surrounded by E-cadherin KO cells (E-cadherin, green, cadherin 6, red, actin, black). XZ and YZ projection show the height difference between the two cells when mixed. Scale bars, 20μm. d,e, Early stages of cell sorting when MDCK WT (magenta) and MDCK E-cadherin KO (green) monolayers are mixed at 30-70 (d) and 70-30 (e) ratio. Scale bars: 100μm.

Supplementary information

Supplementary Information

Supplementary Table 1 and video legends

Reporting Summary

Supplementary Video 1

Phase contrast images overlaid with orientation vectors obtained from MDCK WT monolayers. Blue arrow shows the location of a +1/2 (comet)-shaped defect and the direction in which they move. Scale bar: 50 μm.

Supplementary Video 2

Phase contrast images overlaid with orientation vectors obtained from MDCK E-cadherin KO monolayers. Blue arrow shows the location of a +1/2 (comet)-shaped defect and the direction in which they move. Scale bar: 50 μm.

Supplementary Video 3

Time-lapse demixing of extensile (magenta) and contractile (green) particles obtained from simulations.

Supplementary Video 4

Time-lapse demixing of MDCK WT (extensile-magenta) and MDCK E-cadherin KO (contractile-green) cells observed from experiments. Scale bar: 100 μm.

Supplementary Video 5

Time-lapse demixing of MDCK WT (extensile-magenta) and MDCK E-cadherin KO (contractile-green) cells before and after the addition of 20 μM blebbistatin. Scale bar: 100 μm.

Source data

Source Data Fig. 2

Data used to plot graphs.

Source Data Fig. 3

Data used to plot graphs.

Source Data Fig. 4

Data used to plot graphs.

Source Data Fig. 5

Data used to plot graphs.

Source Data Fig. 6

Data used to plot graphs.

Source Data Extended Data Fig. 1

Data used to plot graphs.

Source Data Extended Data Fig. 4

Data used to plot graphs and uncropped western blots.

Source Data Extended Data Fig. 5

Data used to plot graphs.

Source Data Extended Data Fig. 6

Data used to plot graphs.

Source Data Extended Data Fig. 9

Data used to plot graphs.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Balasubramaniam, L., Doostmohammadi, A., Saw, T.B. et al. Investigating the nature of active forces in tissues reveals how contractile cells can form extensile monolayers. Nat. Mater. 20, 1156–1166 (2021). https://doi.org/10.1038/s41563-021-00919-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41563-021-00919-2

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