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:

Integer topological defects organize stresses driving tissue morphogenesis

This article has been updated

Abstract

Tissues acquire function and shape via differentiation and morphogenesis. Both processes are driven by coordinating cellular forces and shapes at the tissue scale, but general principles governing this interplay remain to be discovered. Here we report that self-organization of myoblasts around integer topological defects, namely spirals and asters, suffices to establish complex multicellular architectures. In particular, these arrangements can trigger localized cell differentiation or, alternatively, when differentiation is inhibited, they can drive the growth of swirling protrusions. Both localized differentiation and growth of cellular vortices require specific stress patterns. By analysing the experimental velocity and orientational fields through active gel theory, we show that integer topological defects can generate force gradients that concentrate compressive stresses. We reveal these gradients by assessing spatial changes in nuclear volume and deformations of elastic pillars. We propose integer topological defects as mechanical organizing centres controlling differentiation 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: Myoblasts arrange into integer topological defects on circular micropatterns.
Fig. 2: Cellular spiral and aster configurations.
Fig. 3: Minimal nematic cellular mounds.
Fig. 4: Integer topological defects concentrate active stress.
Fig. 5: Soft pillars in compressive cellular mounds.
Fig. 6: Topological defects organize differentiation and morphogenesis in 3D nematic tissues.

Similar content being viewed by others

Data availability

The data that support the findings of this study are available in the Zenodo repository https://doi.org/10.5281/zenodo.5549501. High-resolution images and videos are available from the corresponding authors upon request.

Code availability

The MATLAB code used for the 3D-orientation analysis is available in the Zenodo repository https://doi.org/10.5281/zenodo.5680045.

Change history

  • 16 March 2022

    In the version of this article initially published, Supplementary Videos 1,2 were earlier versions of the final, accepted videos, and have been now replaced in the HTML version of the article.

References

  1. Stooke-Vaughan, G. A. & Campàs, O. Physical control of tissue morphogenesis across scales. Curr. Opin. Genet. Dev. 51, 111–119 (2018).

    Article  CAS  Google Scholar 

  2. de Gennes, P. G. & Prost, J. The Physics of Liquid Crystals (Clarendon Press, 1995).

  3. Kemkemer, R., Kling, D., Kaufmann, D. & Gruler, H. Elastic properties of nematoid arrangements formed by amoeboid cells. Eur. Phys. J. E 1, 215 (2000).

    Article  CAS  Google Scholar 

  4. Junkin, M., Leung, S. L., Whitman, S., Gregorio, C. C. & Wong, P. K. Cellular self-organization by autocatalytic alignment feedback. J. Cell Sci. 124, 4213–4220 (2011).

    Article  CAS  Google Scholar 

  5. Duclos, G., Garcia, S., Yevick, H. G. & Silberzan, P. Perfect nematic order in confined monolayers of spindle-shaped cells. Soft Matter 10, 2346–2353 (2014).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  7. Sanchez, T., Chen, D. T. N., DeCamp, S. J., Heymann, M. & Dogic, Z. Spontaneous motion in hierarchically assembled active matter. Nature 491, 431–434 (2012).

    Article  CAS  Google Scholar 

  8. Keber, F. C. et al. Topology and dynamics of active nematic vesicles. Science 345, 1135–1139 (2014).

    Article  CAS  Google Scholar 

  9. Kumar, N., Zhang, R., de Pablo, J. J. & Gardel, M. L. Tunable structure and dynamics of active liquid crystals. Sci. Adv. 4, eaat7779 (2018).

    Article  CAS  Google Scholar 

  10. Thampi, S. P., Golestanian, R. & Yeomans, J. M. Instabilities and topological defects in active nematics. Europhys. Lett. 105, 18001 (2014).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  13. 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 

  14. Duclos, G., Erlenkamper, 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 

  15. Thompson, D. W. On Growth and Form (Cambridge Univ. Press, 1942).

  16. Mandal, S., Mahajan, D., Roy, S., Singh, M. & Khurana, N. Fibroma with minor sex cord elements: an incidental finding in a normal sized ovary—a case report with literature review. Diagn. Pathol. 2, 46 (2007).

    Article  Google Scholar 

  17. Kepes, J. J. Cellular whorls in brain tumors other than meningiomas. Cancer 37, 2232–2237 (1976).

    Article  CAS  Google Scholar 

  18. Dua, H. S. et al. Vortex or whorl formation of cultured human corneal epithelial cells induced by magnetic fields. Eye 10, 447–450 (1996).

    Article  Google Scholar 

  19. Maroudas-Sacks, Y. et al. Topological defects in the nematic order of actin fibres as organization centres of Hydra morphogenesis. Nat. Phys. 17, 251–259 (2021).

    Article  CAS  Google Scholar 

  20. Turiv, T. et al. Topology control of human fibroblast cells monolayer by liquid crystal elastomer. Sci. Adv. 6, eaaz6485 (2020).

    Article  CAS  Google Scholar 

  21. Endresen, K. D., Kim, M., Pittman, M., Chen, Y. & Serra, F. Topological defects of integer charge in cell monolayers. Soft Matter 17, 5878–5887 (2021).

    Article  CAS  Google Scholar 

  22. Yaffe, D. & Feldman, M. The formation of hybrid multinucleated muscle fibers from myoblasts of different genetic origin. Dev. Biol. 11, 300–317 (1965).

    Article  CAS  Google Scholar 

  23. Duclos, G. et al. Spontaneous shear flow in confined cellular nematics. Nat. Phys. 14, 728–732 (2018).

    Article  CAS  Google Scholar 

  24. Wioland, H., Woodhouse, F. G., Dunkel, J., Kessler, J. O. & Goldstein, R. E. Confinement stabilizes a bacterial suspension into a spiral vortex. Phys. Rev. Lett. 110, 268102 (2013).

    Article  CAS  Google Scholar 

  25. Guillamat, P., Ignés-Mullol, J. & Sagués, F. Taming active turbulence with patterned soft interfaces. Nat. Commun. 8, 564 (2017).

    Article  CAS  Google Scholar 

  26. Opathalage, A. et al. Self-organized dynamics and the transition to turbulence of confined active nematics. Proc. Natl Acad. Sci. USA 116, 4788–4797 (2019).

    Article  CAS  Google Scholar 

  27. Jalal, S. et al. Actin cytoskeleton self-organization in single epithelial cells and fibroblasts under isotropic confinement. J. Cell Sci. 132, jcs220780 (2019).

    Article  CAS  Google Scholar 

  28. Kruse, K., Joanny, J. F., Jülicher, F., Prost, J. & Sekimoto, K. Asters, vortices, and rotating spirals in active gels of polar filaments. Phys. Rev. Lett. 92, 078101 (2004).

    Article  CAS  Google Scholar 

  29. Püspöki, Z., Storath, M., Sage, D. & Unser, M. in Focus on Bio-Image Informatics (eds De Vos, W. H., Munck, S. & Timmermans, J.-P.) 69–93 (Springer, 2016).

  30. 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 

  31. Blanch-Mercader, C., Guillamat, P., Roux, A. & Kruse, K. Quantifying material properties of cell monolayers by analyzing integer topological defects. Phys. Rev. Lett. 126, 028101 (2021).

    Article  CAS  Google Scholar 

  32. Blanch-Mercader, C., Guillamat, P., Roux, A. & Kruse, K. Integer topological defects of cell monolayers: mechanics and flows. Phys. Rev. E. 103, 012405 (2021).

    Article  CAS  Google Scholar 

  33. Schaller, V., Weber, C., Semmrich, C., Frey, E. & Bausch, A. R. Polar patterns of driven filaments. Nature 467, 73–77 (2010).

    Article  CAS  Google Scholar 

  34. Lushi, E., Wioland, H. & Goldstein, R. E. Fluid flows created by swimming bacteria drive self-organization in confined suspensions. Proc. Natl Acad. Sci. USA 111, 9733–9738 (2014).

    Article  CAS  Google Scholar 

  35. Doxzen, K. et al. Guidance of collective cell migration by substrate geometry. Integr. Biol. 5, 1026 (2013).

    Article  CAS  Google Scholar 

  36. Fürthauer, S., Neef, M., Grill, S. W., Kruse, K. & Jülicher, F. The Taylor–Couette motor: spontaneous flows of active polar fluids between two coaxial cylinders. New J. Phys. 14, 023001 (2012).

    Article  CAS  Google Scholar 

  37. Hobson, C. M. et al. Correlating nuclear morphology and external force with combined atomic force microscopy and light sheet imaging separates roles of chromatin and lamin A/C in nuclear mechanics. Mol. Biol. Cell 31, 1788–1801 (2020).

    Article  CAS  Google Scholar 

  38. Collinsworth, A. M., Zhang, S., Kraus, W. E. & Truskey, G. A. Apparent elastic modulus and hysteresis of skeletal muscle cells throughout differentiation. Am. J. Physiol. Cell Physiol. 283, C1219–C1227 (2002).

    Article  CAS  Google Scholar 

  39. Da Costa, A. D. S. et al. Fibroblasts close a void in free space by a purse-string mechanism. Preprint at bioRxiv https://doi.org/10.1101/2020.08.15.250811 (2020).

  40. Barbazan, J. et al. Cancer-associated fibroblasts actively compress cancer cells and modulate mechanotransduction. Preprint at bioRxiv https://doi.org/10.1101/2021.04.05.438443 (2021).

  41. Gilbert, S. F. & Barresi, M. J. F. Developmental Biology 11th edn (Sinauer, 2016).

  42. Sharples, A. P., Al-Shanti, N., Lewis, M. P. & Stewart, C. E. Reduction of myoblast differentiation following multiple population doublings in mouse C2C12 cells: a model to investigate ageing? J. Cell. Biochem. 112, 3773–3785 (2011).

    Article  CAS  Google Scholar 

  43. Poujade, M. et al. Collective migration of an epithelial monolayer in response to a model wound. Proc. Natl Acad. Sci. USA 104, 15988–15993 (2007).

    Article  CAS  Google Scholar 

  44. Dekkers, J. F. et al. High-resolution 3D imaging of fixed and cleared organoids. Nat. Protoc. 14, 1756–1771 (2019).

    Article  CAS  Google Scholar 

  45. Strale, P. O. et al. Multiprotein printing by light-induced molecular adsorption. Adv. Mater. 28, 2024–2029 (2016).

    Article  CAS  Google Scholar 

  46. Huterer, D. & Vachaspati, T. Distribution of singularities in the cosmic microwave background polarization. Phys. Rev. D. 72, 043004 (2005).

    Article  CAS  Google Scholar 

  47. Barbier de Reuille, P. et al. MorphoGraphX: a platform for quantifying morphogenesis in 4D. eLife 4, e05864 (2015).

    Article  Google Scholar 

  48. Guo, M. et al. Cell volume change through water efflux impacts cell stiffness and stem cell fate. Proc. Natl Acad. Sci. USA 114, E8618–E8627 (2017).

    CAS  Google Scholar 

  49. Kloxin, A. M., Kloxin, C. J., Bowman, C. N. & Anseth, K. S. Mechanical properties of cellularly responsive hydrogels and their experimental determination. Adv. Mater. 22, 3484–3494 (2010).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank C. Roffay for suggesting the representation of 3D orientational fields and all the members from the Roux and Kruse groups for fruitful discussions and support. We thank S. Gabriele for useful insights on the elastic properties of C2C12 cell nuclei. P.G. acknowledges support from the Human Frontiers of Science Program (grant number LT-000793/2018-C). A.R. acknowledges funding from SystemsX RTD program EpiPhysX, the Swiss National Fund for Research grant numbers 31003A_130520, 31003A_149975 and 31003A_173087, and the European Research Council Consolidator grant number 311536.

Author information

Authors and Affiliations

Authors

Contributions

P.G. and A.R. designed the research. P.G. performed the experiments. P.G. and C.B.-M. analysed the data. G.P. analysed cell number density in mounds. C.B.-M. and K.K. developed the theoretical model. All the authors participated in writing the manuscript.

Corresponding authors

Correspondence to Karsten Kruse or Aurélien Roux.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Materials thanks Marino Arroyo, Michael Murrell 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 Formation of a nematic cellular monolayer.

a, Time series of a proliferating monolayer of C2C12 myocytes. Scale bar, 500μm. b, Average orientational order parameter as function of time. c, Two-point spatial autocorrelation function Cnn and nematic autocorrelation length ξnn (inset). Colors correspond to the time points indicated in the inset. Data are presented as mean values + /- s.d.

Extended Data Fig. 2 Cell number density for spirals and asters.

a, Time evolution of confined cell monolayers with different initial cell number. Scale bar, 100μm. b, Time of maximum tangential speed with respect of initial cell number. Colormap indicates speed. Dashed line corresponds to a linear fit. c, Average cell number density and d, radial density profiles for spirals (N = 10) and asters (N = 9). Data are presented as mean values ± s.e.m.

Extended Data Fig. 3 Orientational field from cell-shape and actin fibers in spirals.

a, Phase contrast image of a spiral. Yellow lines correspond to the local orientation. Vector length corresponds to the coherency (see Methods). For clarity, the totality of vectors is only shown in the inset. b, Confocal micrograph of the bottom plane of an F-actin-labelled spiral. Actin was stained with SiR-actin. Scale bars, 25μm. c and d, Distributions of the angle ψ extracted from phase contrast images (N = 11) and confocal micrographs (N = 7), respectively. The angle ψ is between the local director and the radial direction with respect to the geometrical center. In panels C and D, the red dashed line indicates ψ = 90°. e and f, Radial profiles of the orientational order parameter S extracted from phase contrast images (N = 11) and confocal micrographs (N = 7), respectively. Data are presented as mean values ± s.e.m.

Extended Data Fig. 4 Influence of myosin activity on spiral morphology and dynamics.

a, Phase contrast image of a Y-27632-treated spiral. The confinement radius is 100μm. b, Average flow field for Y-27632-treated spirals. Streamlines are shown as black curves. The color code indicates the speed. c, Distributions of the angle ψ for spirals under different conditions. ψ is the angle between the local director and the radial direction with respect to the geometrical center. Dashed red line indicates ψ = 90°. d, Radial profiles of the azimuthal and e, radial velocity components for spirals under the conditions corresponding to panel C. N = 11, 20 and 21, for control spirals, Y-27632-treated spirals, and Blebbistatin-treated spirals, respectively. Data are presented as mean values ± s.e.m.

Extended Data Fig. 5 Actin flows in asters.

a, Confocal micrographs of actin at different aster bases. b, Average orientational field from time-lapses of asters in A (N = 3). Actin was stained with SiR-actin. Vector length corresponds to the coherency, showed as well as a colormap (see Methods). Scale bar, 50μm. c, Distribution of the angle ψ extracted from confocal micrographs (N = 3). The angle ψ is between the local director and the radial direction with respect to the geometrical center. Dashed red line indicates ψ = 90° (N = 3). d, Average flow field. Streamlines and vectors (inset) indicate the direction of actin flows. The colormap indicates the speed. e, Radial profiles of the radial (red) and azimuthal (black) velocity components. Data are presented as mean values ± s.e.m.

Extended Data Fig. 6 Aster arrangements and rings at the pillars’ base.

a, Time-series of a C2C12 cell monolayer initially positioned on a circular ring constricting a hydrogel micro-pillar. Cyan dashed line indicates the initial pillar section. A schematic of the setting is included as an inset. b, Actin-stained cells after constriction show an aster arrangement. Actin was stained with SiR-actin. c, Average flow field around pillars during constriction (N = 9). The color scale corresponds to the speed. d, Radial profiles of the radial (red) and azimuthal (black) velocity components. Data are presented as mean values ± s.e.m. e, Average orientational field. For clarity, only half of the total number of orientation vectors are shown. The color scale represents the angle of the director with respect to the horizontal axis. Dashed circles mark the position of the pillar section and the initial ring. Vector length corresponds to the coherency (see Methods). f, Distribution of the angle ψ between the local director and the radial direction with respect to the geometrical center. Scale bars, 50μm.

Extended Data Fig. 7 Compression dynamics of hydrogel pillars.

a, Schematics of the protocol employed to fabricate cell-adhesive rings enclosing passive fluorescent hydrogel micro-pillars. After activation of the glass substrate, micro-pillars were fabricated by illuminating a photo-polymerizable mPEG solution with disk patterns of UV light with an inverted microscope (step I). Then, the substrate was functionalized with PLL-PEG (step II). PEG chains were locally photo-degraded by illuminating the substrate with ring patterns of UV light (step III). Finally, fibronectin was incubated (step IV). b, Time-series showing the compression of a pillar. Green dashed line on bright-field images depicts the initial size of the pillar section close to the base with a radius of r = 40μm. The radius of the cellular island is 100μm. 3D rendering of the pillar is shown below. c, Temporal evolution of the height up to which pillars are compressed for different radii. d, Compressive stress for pillars with different radii. Data are presented as mean values ± s.e.m. e, Maximum compressive stress vs pillar radii. For panels C, D and E, the radius of pillars (r) is indicated in the legend of panel E. f, Maximum Z projections of a confocal stack of a mound deforming a centred pillar. g, Position of the centroid of the sections of 2 centered pillars (Green and blue). Green curves correspond to the pillar in F. Red curves correspond to a pillar without cells around it. h, Maximum Z projection of a confocal stack of a mound deforming a non-centered pillar. i, Position of the centroid of the sections of 3 non-centered pillars. Green curves correspond to the pillar in H. For G and I, centroids of the pillar base are positioned at the origin. Data are presented as mean values ± s.d. Actin was stained with SiR-actin. Nuclei were stained with Hoechst 33342. Scale bars, 50μm.

Extended Data Fig. 8 Localization of myosin heavy chain expression.

a, Image composites of phase contrast and fluorescence images of confluent monolayers of C2C12 cells competent for differentiation, after 6 days under starvation conditions (see Methods). Right panel corresponds to a zoomed region from the left panel. Scale bars, 500μm. b, Average projection of confocal micrographs of cellular islands (N = 40, base radius of 100μm) grown for 6 days in complete medium (10% FBS). Individual micrographs are shown in the panels on the right. c, Average projection of confocal micrographs of cellular islands (N = 40, base radius of 100 μm) grown for 6 days under starvation conditions (2% HS). Individual micrographs are shown in the panels on the right. For all panels, actin is shown in white, and myosin heavy chain (MyHC) is shown in yellow. Actin was stained with SiR-actin. MyHC was stained with the Myosin-4 Monoclonal Antibody conjugated with Alexa Fluor 488. Scale bars, 50μm. d, Radial profiles of averaged MyHC intensity and e, compressive stress under normal (orange) and starvation (black) conditions. Compressive stresses were measured from nuclear volume deformations. Data are presented as mean values ± s.e.m.

Extended Data Fig. 9 Cellular protrusions.

a, Image composites of phase contrast and fluorescence images of confluent monolayers of C2C12 cells incompetent for differentiation, after 6 days under starvation conditions (see Methods). Right panel corresponds to a zoomed region from the left panel. Scale bars, 500μm. MyHC was stained with the Myosin-4 Monoclonal Antibody conjugated with Alexa Fluor 488. b, Z-projections of actin of 14- (above yellow dashed line) and 16- (below yellow dashed line) days-old cellular protrusions. Red frame corresponds to a zoomed protrusion. Actin was stained with SiR-actin. Scale bars, 50μm.

Extended Data Fig. 10 Growth of cellular mounds and protrusions.

a, Average height of C2C12 islands of radius 100 µm for different time points after the onset of confluence. White symbols () correspond to islands of differentiation-competent cells (N = 17, 10, 16, for mounds at 1, 3, 6 days after the onset of confluence, respectively). Black symbols () correspond to islands of differentiation-incompetent cells (N = 42, 11, 10 and 27, for mounds at 5, 9, 12, and 16 days after the onset of confluence, respectively). Data are presented as mean values ± s.d. Inset: scheme of the time evolution of cellular nematic architectures from 2D to 3D. b, XZ section of mounds under different confinement sizes. The number of mounds in the images is 6, 5, 2 and 1 for diameters of 200, 500, 1000 and 4000 µm, respectively. Mounds were labelled with Sir Actin. c, Z projection of an actin labelled 6 day old mound with a diameter of 500 µm. Colormap indicates height. d, Aspect ratio of the mounds, calculated as height/diameter, for the different confinement sizes (N = 30, 16, 4 and 5 for diameters of 200, 500, 1000 and 4000 µm, respectively). Data are presented as mean values ± s.d. Scale bars, 200 µm.

Supplementary information

Supplementary Information

Captions for Supplementary Video 1–10, Supplementary Notes 1–8, Figs. 1–5 and References.

Reporting Summary

Supplementary Video 1

Spiral-to-aster transition in a C2C12 myoblast disk. Phase contrast time-lapse of myoblast monolayers under circular confinement. In time, cells rearrange from spiral arrangements into aster arrangements.

Supplementary Video 2

Aster-to-spiral transition in an expanding C2C12 myoblast disk. Phase contrast time-lapse of an expanding myoblast disc, from 150 to 300 μm in diameter. Blue shaded area depicts the new fibronectin-coated region. White dashed lines indicate the outer boundary of the new circular domain.

Supplementary Video 3

Formation of cellular mounds. Phase-contrast time-lapse showing the formation of cellular mounds in the centre of an aster of myoblasts.

Supplementary Video 4

Cellular spirals. Phase-contrast time-lapse of low-density circular islands of myoblasts featuring spiral configurations. Division was blocked with Mitomycin-C.

Supplementary Video 5

Actin dynamics in cellular spirals. Fluorescence confocal time-lapse of the bottom plane of a cellular spiral. Actin was stained with SiR-actin.

Supplementary Video 6

Cellular asters. Phase-contrast time-lapse of high-density circular islands of myoblasts. Division was not blocked.

Supplementary Video 7

Actin dynamics in cellular asters. Fluorescence confocal time-lapse of the bottom plane of a cellular aster. Actin was stained with SiR-actin.

Supplementary Video 8

Actin organization and flows in minimal cellular mounds. Z projection of the top of an actin-labelled mound. Actin was stained with SiR-actin.

Supplementary Video 9

Pillar constriction experiment. Differential interference contrast time-lapse showing myoblasts constricting soft hydrogel pillars of different sizes. The consequent pillars’ deformation can be observed in the 3D renderings of the pillars obtained from fluorescence images’ segmentation.

Supplementary Video 10

Dynamics and collapse of 3D cellular protrusions. Phase-contrast time-lapse of a myoblast protrusion, which collapses after confining pattern is degraded.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guillamat, P., Blanch-Mercader, C., Pernollet, G. et al. Integer topological defects organize stresses driving tissue morphogenesis. Nat. Mater. 21, 588–597 (2022). https://doi.org/10.1038/s41563-022-01194-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41563-022-01194-5

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