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:

Epithelia are multiscale active liquid crystals

Abstract

Biological processes such as embryogenesis, wound healing and cancer progression rely on the ability of epithelial cells to coordinate their mechanical activity over length scales that are orders of magnitude larger than the typical cellular size1. Although this process is regulated by various signalling pathways2, it has recently become evident that this coordination can be understood using physics tools, of which liquid crystal order is a prominent example3,4. However, prior studies have given inconsistent results in this respect. Whereas nematic order is often invoked in the interpretation of experimental data3,4,5, computational models have instead suggested that hexatic order could serve as a linchpin for collective migration in confluent cell layers6,7,8. In this Letter, we resolve this dilemma. Using a combination of in vitro experiments, numerical simulations and analytical work, we demonstrate that both nematic and hexatic order are present in epithelial layers, with the former being dominant at large length scales and the latter at small length scales. In cells on uncoated glass, these different types of liquid crystal order cross over at a length scale of the order of ten cell sizes. Our work provides a framework for deciphering epithelial structure and may lead to a predictive mesoscopic theory of tissues9.

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: Orientational order parameter and defect misidentification.
Fig. 2: Symmetry of MDCK cells in confluent monolayers.
Fig. 3: Coarse-graining and multiscale order in confluent epithelial monolayers.

Similar content being viewed by others

Data availability

The data that support the findings of this study and the code for data analysis are openly available in GitHub at https://github.com/hexanematic/orientation_tracker. Source data are provided with this paper.

References

  1. Friedl, P. & Gilmour, D. Collective cell migration in morphogenesis, regeneration and cancer. Nat. Rev. Mol. Cell Biol. 10, 445–457 (2009).

    Article  Google Scholar 

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

    Article  Google Scholar 

  3. Duclos, G., Erlenkämper, C., Joanny, J.-F. & Silberzan, P. Topological defects in confined populations of spindle-shaped cells. Nat. Phys. 13, 58 (2016).

    Article  Google Scholar 

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

    Article  ADS  Google Scholar 

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

    Article  ADS  Google Scholar 

  6. Li, Y.-W. & Ciamarra, M. P. Role of cell deformability in the two-dimensional melting of biological tissues. Phys. Rev. Mater. 2, 045602 (2018).

    Article  Google Scholar 

  7. Durand, M. & Heu, J. Thermally driven order-disorder transition in two-dimensional soft cellular systems. Phys. Rev. Lett. 123, 188001 (2019).

    Article  ADS  Google Scholar 

  8. Pasupalak, A., Yan-Wei, L., Ni, R. & Pica Ciamarra, M. Hexatic phase in a model of active biological tissues. Soft Matter 16, 3914 (2020).

    Article  ADS  Google Scholar 

  9. Armengol Collado, J.-M., Carenza, L. N. & Giomi, L. Hydrodynamics and multiscale order in confluent epithelia. Preprint at https://arxiv.org/abs/2202.00651 (2022).

  10. Balasubramaniam, L. et al. Investigating the nature of active forces in tissues reveals how contractile cells can form extensile monolayers. Nat. Mater. 20, 1156 (2021).

    Article  ADS  Google Scholar 

  11. Maroudas-Sacks, Y. et al. Topological defects in the nematic order of actin fibres as organization centres of hydra morphogenesis. Nat. Phys. 17, 251 (2021).

    Article  Google Scholar 

  12. Hoffmann, L. A., Carenza, L. N., Eckert, J. & Giomi, L. Theory of defect-mediated morphogenesis. Sci. Adv. 8, eabk2712 (2022).

    Article  Google Scholar 

  13. Streichan, S. J., Lefebvre, M. F., Noll, N., Wieschaus, E. F. & Shraiman, B. I. Global morphogenetic flow is accurately predicted by the spatial distribution of myosin motors. eLife 7, e27454 (2018).

    Article  Google Scholar 

  14. Bigun, J. Optimal Orientation Detection of Linear Symmetry (Linköping Univ. Electronic Press, 1987).

  15. Aubouy, M., Jiang, Y., Glazier, J. A. & Graner, F. A texture tensor to quantify deformations. Granul. Matter 5, 67 (2003).

    Article  MATH  Google Scholar 

  16. Asipauskas, M., Aubouy, M., Glazier, J. A., Graner, F. & Jiang, Y. A texture tensor to quantify deformations: the example of two-dimensional flowing foams. Granul. Matter 5, 71 (2003).

    Article  MATH  Google Scholar 

  17. Chaikin, P. M. & Lubensky, T. C. Principles of Condensed Matter Physics (Cambridge Univ. Press, 1995).

  18. Giomi, L. Geometry and topology of turbulence in active nematics. Phys. Rev. X 5, 031003 (2015).

    Google Scholar 

  19. Friedel, G. Les états mésomorphes de la matière. Ann. Phys. 9, 273 (1922).

    Article  Google Scholar 

  20. Giomi, L., Toner, J. & Sarkar, N. Long-ranged order and flow alignment in sheared p-atic liquid crystals. Phys. Rev. Lett. 129, 067801 (2022).

    Article  ADS  MathSciNet  Google Scholar 

  21. Giomi, L., Toner, J. & Sarkar, N. Hydrodynamic theory of p-atic liquid crystals. Phys. Rev. E 106, 024701 (2022).

    Article  ADS  MathSciNet  Google Scholar 

  22. Virga, E. G. Octupolar order in two dimensions. Eur. Phys. J. E 38, 63 (2015).

    Article  Google Scholar 

  23. Halperin, B. I. & Nelson, D. R. Theory of two-dimensional melting. Phys. Rev. Lett. 41, 121 (1978).

    Article  ADS  MathSciNet  Google Scholar 

  24. Nelson, D. R. & Halperin, B. I. Dislocation-mediated melting in two dimensions. Phys. Rev. B 19, 2457 (1979).

    Article  ADS  Google Scholar 

  25. Bernard, E. P. & Krauth, W. Two-step melting in two dimensions: first-order liquid-hexatic transition. Phys. Rev. Lett. 107, 155704 (2011).

    Article  ADS  Google Scholar 

  26. Digregorio, P. et al. Full phase diagram of active brownian disks: from melting to motility-induced phase separation. Phys. Rev. Lett. 121, 098003 (2018).

    Article  ADS  Google Scholar 

  27. Yamada, S., Pokutta, S., Drees, F., Weis, W. I. & Nelson, W. J. Deconstructing the cadherin-catenin-actin complex. Cell 123, 889 (2005).

    Article  Google Scholar 

  28. Loewe, B., Chiang, M., Marenduzzo, D. & Marchetti, M. C. Solid-liquid transition of deformable and overlapping active particles. Phys. Rev. Lett. 125, 038003 (2020).

    Article  ADS  Google Scholar 

  29. Monfared, S., Ravichandran, G., Andrade, J. & Doostmohammadi, A. Mechanical basis and topological routes to cell elimination. eLife 12, e82435 (2023).

    Article  Google Scholar 

  30. Carenza, L. N., Gonnella, G., Lamura, A., Negro, G. & Tiribocchi, A. Lattice boltzmann methods and active fluids. Eur. Phys. J. E 42, 81 (2019).

    Article  Google Scholar 

  31. Bi, D., Yang, X., Marchetti, M. C. & Manning, M. L. Motility-driven glass and jamming transitions in biological tissues. Phys. Rev. X 6, 021011 (2016).

    Google Scholar 

  32. Frenkel, D. & Eppenga, R. Evidence for algebraic orientational order in a two-dimensional hard-core nematic. Phys. Rev. A 31, 1776 (1985).

    Article  ADS  Google Scholar 

  33. Bagchi, K., Andersen, H. C. & Swope, W. Computer simulation study of the melting transition in two dimensions. Phys. Rev. Lett. 76, 255 (1996).

    Article  ADS  Google Scholar 

  34. Kittel, C. & McEuen, P. Introduction to Solid State Physics, Vol. 8 (Wiley, 1996).

  35. Eckert, J., Ladoux, B., Giomi, L., & Schmidt, T. Hexanematic crossover in epithelial monolayers depends on cell adhesion and cell density. Preprint at bioRxiv https://doi.org/10.1101/2022.10.07.511294 (2022).

  36. Majmudar, T. S. & Behringer, R. P. Contact force measurements and stress-induced anisotropy in granular materials. Nature 435, 1079 (2005).

    Article  ADS  Google Scholar 

  37. Gao, G.-J. J., Holcomb, M. C., Thomas, J. H. & Blawzdziewicz, J. Embryo as an active granular fluid: stress-coordinated cellular constriction chains. J. Phys. Condens. Matter 28, 414021 (2016).

    Article  Google Scholar 

  38. Keber, F. C. et al. Topology and dynamics of active nematic vesicles. Science 345, 1135 (2014).

    Article  ADS  Google Scholar 

  39. Guillamat, P., Blanch-Mercader, C., Kruse, K. & Roux, A. Integer topological defects organize stresses driving tissue morphogenesis. Nat. Mater. 21, 588–597 (2022).

    Article  ADS  Google Scholar 

  40. Blankenship, J. T., Backovic, S. T., Sanny, J., Weitz, O. & Zallen, J. A. Multicellular rosette formation links planar cell polarity to tissue morphogenesis. Dev. Cell 11, 459 (2006).

    Article  Google Scholar 

  41. Rauzi, M. Cell intercalation in a simple epithelium. Phil. Trans. R. Soc. B 375, 20190552 (2020).

    Article  Google Scholar 

  42. Haeger, A. et al. Collective cancer invasion forms an integrin-dependent radioresistant niche. J. Exp. Med. 217, e20181184 (2020).

    Article  Google Scholar 

  43. Serra-Picamal, X. et al. Mechanical waves during tissue expansion. Nat. Phys. 8, 628 (2012).

    Article  Google Scholar 

Download references

Acknowledgements

This work is supported by the European Union via the ERC-CoGgrant HexaTissue (L.N.C., D.K. and L.G.) and by the Netherlands Organization for Scientific Research (NWO/OCW) as part of the research programme ‘The active matter physics of collective metastasis’ with project number Science-XL 2019.022 (J.-M.A.-C. and L.G.). Part of this work was carried out on the Dutch national e-infrastructure with the support of SURF through Grant no. 2021.028 for computational time. J.E. and L.G. thank M. Gloerich, UMC Utrecht, for providing us with the MDCK cells. L.N.C. thanks G. Negro, University of Edinburgh, for providing us with the data for passive Brownian particles. All authors thank L. Hoffmann for fruitful discussions.

Author information

Authors and Affiliations

Authors

Contributions

J.-M.A.-C. performed analytic work and Voronoi model simulations and analysed data. L.N.C. coordinated the research, performed the multiphase-field simulations and analysed data. J.E. performed analytic work and experiments and analysed data. D.K. performed analytic work and analysed data. L.G. devised and coordinated the research. All authors wrote the paper. J.-M.A.-C., L.N.C., J.E. and D.K. contributed equally to this work.

Corresponding author

Correspondence to Luca Giomi.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Physics thanks Joshua Shaevitz 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 Misidentification of nematic defects in a honeycomb lattice.

Because of the 6-fold symmetry of regular hexagons, there is no well defined longitudinal direction, thus it is possible to construct a defective configuration, featuring a pair of ± 1/2 disclinations, even though the lattice is defect free.

Extended Data Fig. 2 Probability distribution of cell coordination number for experiments and simulations.

Data are presented as mean values of the total number of analyzed images ± s.e.m. Sample size (number of points per bar): experiment n = 68, Mpf simulations n = 286, Voronoi simulations n = 100. The mean values of the distribution of the coordination number are: 5.8 ± 0.9 (mean ± s.d.) for experiments and 5.9 ± 0.9 (mean ± s.d.) for multiphase field simulations and 6.0 ± 0.6 (mean ± s.d.) for Voronoi simulations.

Source data

Extended Data Fig. 3 Probability distribution of the order parameter’s amplitude γ2 and γ6 for the Voronoi and Mpf model as a function of the main simulation parameters.

(a) Probablity distribution of γp, p = 2, 6 for the self-propelled Voronoi model when varying the shape index p0. At increasing p0, the distribution for the hexatic shape function flattens and it peaks at lower values. The opposite happens for p = 2, which turns from being negligible to exhibit a comparable peak value to the one of the hexatic symmetry. (b) Same effect when varying the self-propulsion v0. In this case, we lower p0 = 3.7 in order to capture the melting transition: there is no substantial variation of the profiles when varying v0 at p0 = 3.9 as the system is deeply in the liquid phase. These observations are coherent with the unjamming transition that has been reported for this particle-based model in Ref. 31. (c) Probablity distribution of γp, p = 2, 6 for the Mpf model when varying the velocity of self-propulsion v0. Analogously to the case of the Voronoi model, γ6 (γ2) is larger (smaller) at low (high) v0 values. The simulations of both the Voronoi and Mpf have been performed under the same values of model parameters provided in Table 1, except for the shape index p0 and self-propulsion velocity which are varied as reported in the legend of panels (a-c). Data are presented as mean values of the total number of analyzed images ± s.d. Sample size (number of points per bar): Voronoi simulations n = 100 (panels (a-b), Mpf simulations n = 286 (panel (c)).

Source data

Extended Data Fig. 4 Orientational order parameter at different length scales.

Nematic (a) and hexatic (b) shape parameter Γ2 and Γ6 at different values of the coarse-graining radius R, expressed in units of the average cell size Rcell = 7.4μm for the same experimental image as in Fig. 3 of the main text. In both top and bottom panels, positive and negative defects are marked in red and blue respectively ( ± 1/2 for nematic and ± 1/6 for hexatic). The spatial resolution of the shape parameter is the same on both panels, but the number of orientation markers in panel (b) is half of that used in panel (a) for sake of clarity.

Extended Data Fig. 5 Misidentification of singular lines in p − atic liquid crystals.

Assigning a 2-fold orientation to individual cells leads to a misdetection of their orientational order, including what concerns the identification of topological defects. (a) The best known example of an analogous misdetection, is obtained when one attempts to track the orientation of the nematic director in the surrounding of ± 1/2 disclinations using a vector field, whose rotational symmetry is 1-fold, rather than 2-fold as required. In this case, the most visible signature of the misdetection consists of a singular line, across which the vector field “jumps” by an angle of 180°. This singular line is marked by oppositely oriented red arrows in the panel. (b) The same effect can be observed when a pair of ± 1/6 hexatic defects is tracked using a 2-fold orientation field. In this case, the 2-fold orientation field jumps by 60°, as indicated by red segments. In general, detecting a ± n/p defect, with n a natural number, using an arbitrary q-fold orientation field, with p ≠ q, yields a singular line where the q-fold orientation jumps by an angle of \(2\pi \min [n/p,| p-nq| /(pq)]\) as demonstrated in the Supplemenatry Sec. S3.

Extended Data Fig. 6 Decay of defect density in simulations.

Decay of the defect density at varying the coarse-graining radius in simulations of the multiphase field model (a) and Voronoi model (b). The scaling of the multiphase field model data is compatible with that shown in panel d of Fig. 3 in the main text for experimental data. In the right panel, curves of oppositely charged defects overlap exactly. Data are presented as mean values of the total number of analyzed configurations ± s.e.m. Sample size (number of configurations analyzed per point): Mpf simulations n = 286, Voronoi simulations n = 100.

Source data

Extended Data Fig. 7 Nematic filamentary structures.

Computer reconstruction of MDCK II cell monolayer (same sample as in Fig. 3b,c) with the cells highlighted based on the orientation of the coarse-grained nematic director (solid lines) for R = 4.5Rcell. The emergence of filamentary structures is especially evident in proximity of the topological defects.

Extended Data Fig. 8 Decay of p − fold orientational order.

Magnitude of Γp, with p = 2, …, 6, versus the coarse-graining radius R measured for experimental (a) and numerical Mpf data (b). Orientational order decays with power law behavior \(| {\Gamma }_{p}| \sim {(R/{R}_{{{{\rm{cell}}}}})}^{-{\eta }_{p}/2}\). We notice that η3, η5 ≈ 2, the critical value corresponding to a disordered state, therefore suggesting 3 − and 5 − fold symmetry to be significantly suppressed in our experimental and Mpf model tissues. For p = 4, on the other hand, η4 = 1.38 ± 0.01 in experiments and η4 = 1.16 ± 0.03 in simulations, thus suggesting the presence of tetratic orientational order. This behavior is, however, the consequence of the degeneracy of the shape funciton illustrated in Section S1 of the Supplement. Hence, the observation of a residual tetratic order follows from the leading nematic order. Finally, we stress that the largest contribution at small and large scale is given by the hexatic and nematic orded parameter, respectively, for both experimental and simulation data. Data are presented as mean values of the total number of analyzed configurations ± s.e.m. (see Methods section). Sample size (number of configurations analyzed per point): experiment n = 68, Mpf simulations n = 286.

Source data

Extended Data Fig. 9 Spatial correlation function of experimental and Mpf data.

Two-point spatial correlation function \(\left\langle {\gamma }_{p}^{* }({{{\boldsymbol{r}}}}){\gamma }_{p}({{{\boldsymbol{0}}}})\right\rangle\), normalized with respect to the magnitude of the shape parameter Γp for p = 2 and p = 6. Data are presented as mean values of the total number of analyzed configurations ± s.d. Sample size (number of configurations analyzed per point): experiment n = 68, Mpf simulations n = 286.

Source data

Extended Data Fig. 10 Orientational order decay and correlation functions of Voronoi model.

(a) Restricting the cell to be Voronoi polygons artificially enhances their hexagonality thereby favoring hexatic order, which then prevails at all scales. (b) Spatial correlation of the shape functions γ2 and γ6. Unlike the correlation functions shown in Fig. E9, corresponding to the experimental results and Mpf simulations, the residual hexatic order causes \(\left\langle {\gamma }_{6}^{* }({{{\boldsymbol{r}}}}){\gamma }_{6}({{{\boldsymbol{0}}}})\right\rangle\) to be greater than \(\left\langle {\gamma }_{2}^{* }({{{\boldsymbol{r}}}}){\gamma }_{2}({{{\boldsymbol{0}}}})\right\rangle\) at large length scales. These simulations have been performed under the same parameter values provided in Table 1 of the main text. Data are presented as mean values of the total number of analyzed configurations ± s.d. Sample size (number of configurations analyzed per point) n = 100.

Source data

Supplementary information

Supplementary Information

Supplementary Discussion: Orientational order in two-dimensional p-atic liquid crystals, the p-fold shape tensor and defect representation in p-atic liquid crystals, and Figs. 1–5.

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. 6

Statistical source data.

Source Data Extended Data Fig. 8

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

Armengol-Collado, JM., Carenza, L.N., Eckert, J. et al. Epithelia are multiscale active liquid crystals. Nat. Phys. 19, 1773–1779 (2023). https://doi.org/10.1038/s41567-023-02179-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41567-023-02179-0

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