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:

Mechanical force application to the nucleus regulates nucleocytoplasmic transport

Abstract

Mechanical force controls fundamental cellular processes in health and disease, and increasing evidence shows that the nucleus both experiences and senses applied forces. Such forces can lead to the nuclear translocation of proteins, but whether force controls nucleocytoplasmic transport, and how, remains unknown. Here we show that nuclear forces differentially control passive and facilitated nucleocytoplasmic transport, setting the rules for the mechanosensitivity of shuttling proteins. We demonstrate that nuclear force increases permeability across nuclear pore complexes, with a dependence on molecular weight that is stronger for passive than for facilitated diffusion. Owing to this differential effect, force leads to the translocation of cargoes into or out of the nucleus within a given range of molecular weight and affinity for nuclear transport receptors. Further, we show that the mechanosensitivity of several transcriptional regulators can be both explained by this mechanism and engineered exogenously by introducing appropriate nuclear localization signals. Our work unveils a mechanism of mechanically induced signalling, probably operating in parallel with others, with potential applicability across signalling pathways.

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: Nucleocytoplasmic transport is mechanosensitive.
Fig. 2: Passive diffusion through NPCs is mechanosensitive for small MWs.
Fig. 3: Differential mechanosensitivity of facilitated import versus passive diffusion explains force-induced nuclear translocation.
Fig. 4: Balance between affinity to importins and MW defines the mechanosensitivity of nuclear localization.
Fig. 5: Balance between affinity to exportin 1 and MW defines the mechanosensitivity of nuclear localization in constructs containing NES signals.
Fig. 6: The mechanosensitivity of twist1 can be re-engineered with exogenous NLS sequences.

Similar content being viewed by others

Data availability

Source data are provided with this paper. All other data supporting the findings of this study are available from the corresponding author on reasonable request.

Code availability

All codes generated during the current study are available from the corresponding authors on reasonable request. The kinetic mathematical model of transport is available at https://github.com/ravehlab/npctransport_kinetic.

References

  1. Broders-Bondon, F., Ho-Bouldoires, T. H. N., Fernandez-Sanchez, M. E. & Farge, E. Mechanotransduction in tumor progression: the dark side of the force. J. Cell Biol. 217, 1571–1587 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Hamant, O. & Saunders, T. E. Shaping organs: shared structural principles across kingdoms. Annu. Rev. Cell Dev. Biol. 36, 385–410 (2020).

    Article  CAS  PubMed  Google Scholar 

  3. Humphrey, J. D., Schwartz, M. A., Tellides, G. & Milewicz, D. M. Role of mechanotransduction in vascular biology: focus on thoracic aortic aneurysms and dissections. Circ. Res. 116, 1448–1461 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Lombardi, M. L. et al. The interaction between nesprins and sun proteins at the nuclear envelope is critical for force transmission between the nucleus and cytoskeleton. J. Biol. Chem. 286, 26743–26753 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Arsenovic, P. T. et al. Nesprin-2G, a component of the nuclear LINC complex, is subject to myosin-dependent tension. Biophys. J. 110, 34–43 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Elosegui-Artola, A. et al. Force triggers YAP nuclear entry by regulating transport across nuclear pores. Cell 171, 1397–1410 e14 (2017).

    Article  CAS  PubMed  Google Scholar 

  7. Kirby, T. J. & Lammerding, J. Emerging views of the nucleus as a cellular mechanosensor. Nat. Cell Biol. 20, 373–381 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Nava, M. M. et al. Heterochromatin-driven nuclear softening protects the genome against mechanical stress-induced damage. Cell 181, 800–817.e22 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Tajik, A. et al. Transcription upregulation via force-induced direct stretching of chromatin. Nat. Mater. 15, 1287–1296 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Swift, J. et al. Nuclear lamin-A scales with tissue stiffness and enhances matrix-directed differentiation. Science 341, 1240104 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Lomakin, A. J. et al. The nucleus acts as a ruler tailoring cell responses to spatial constraints. Science 370, eaba2894 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Venturini, V. et al. The nucleus measures shape changes for cellular proprioception to control dynamic cell behavior. Science 370, eaba2644 (2020).

    Article  CAS  PubMed  Google Scholar 

  13. Kassianidou, E., Kalita, J. & Lim, R. Y. H. The role of nucleocytoplasmic transport in mechanotransduction. Exp. Cell. Res. 377, 86–93 (2019).

    Article  CAS  PubMed  Google Scholar 

  14. Zhao, X. H. et al. Force activates smooth muscle α-actin promoter activity through the Rho signaling pathway. J. Cell Sci. 120, 1801–1809 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Ho, C. Y., Jaalouk, D. E., Vartiainen, M. K. & Lammerding, J. Lamin A/C and emerin regulate MKL1–SRF activity by modulating actin dynamics. Nature 497, 507–511 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Fernandez-Sanchez, M. E. et al. Mechanical induction of the tumorigenic β-catenin pathway by tumour growth pressure. Nature 523, 92–95 (2015).

    Article  CAS  PubMed  Google Scholar 

  17. Gayrard, C., Bernaudin, C., Dejardin, T., Seiler, C. & Borghi, N. Src- and confinement-dependent FAK activation causes E-cadherin relaxation and β-catenin activity. J. Cell Biol. 217, 1063–1077 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Aragona, M. et al. Mechanisms of stretch-mediated skin expansion at single-cell resolution. Nature 584, 268–273 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Chang, L. et al. The SWI/SNF complex is a mechanoregulated inhibitor of YAP and TAZ. Nature 563, 265–269 (2018).

    Article  CAS  PubMed  Google Scholar 

  20. Ege, N. et al. Quantitative analysis reveals that actin and src-family kinases regulate nuclear YAP1 and its export. Cell Syst. 6, 692–708.e13 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Jacchetti, E. et al. The nuclear import of the transcription factor MyoD is reduced in mesenchymal stem cells grown in a 3D micro-engineered niche. Sci. Rep. 11, 1–19 (2021).

    Article  CAS  Google Scholar 

  22. SR, W. & MP, R. The nuclear pore complex and nuclear transport. Cold Spring Harb. Perspect. Biol. 2, a000562 (2010).

    Google Scholar 

  23. M, B. & E, H. The nuclear pore complex: understanding its function through structural insight. Nat. Rev. Mol. Cell Biol. 18, 73–89 (2017).

    Article  Google Scholar 

  24. Paine, P. L. & Feldherr, C. M. Nucleocytoplasmic exchange of macromolecules. Exp. Cell. Res. 74, 81–98 (1972).

    Article  CAS  PubMed  Google Scholar 

  25. Timney, B. L. et al. Simple rules for passive diffusion through the nuclear pore complex. J. Cell Biol. 215, 57–76 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Mohr, D., Frey, S., Fischer, T., Güttler, T. & Görlich, D. Characterisation of the passive permeability barrier of nuclear pore complexes. EMBO J. 28, 2541–2553 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Denning, D. P., Patel, S. S., Uversky, V., Fink, A. L. & Rexach, M. Disorder in the nuclear pore complex: the FG repeat regions of nucleoporins are natively unfolded. Proc. Natl Acad. Sci. USA 100, 2450–2455 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Yuh, M. C. & Blobel, G. Karyopherins and nuclear import. Curr. Opin. Struct. Biol. 11, 703–715 (2001).

    Article  Google Scholar 

  29. Nachury, M. V. & Weis, K. The direction of transport through the nuclear pore can be inverted. Proc. Natl Acad. Sci. USA 96, 9622–9627 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Cautain, B., Hill, R., De Pedro, N. & Link, W. Components and regulation of nuclear transport processes. FEBS J. 282, 445–462 (2015).

    Article  CAS  PubMed  Google Scholar 

  31. Kalderon, D., Roberts, B. L., Richardson, W. D. & Smith, A. E. A short amino acid sequence able to specify nuclear location. Cell 39, 499–509 (1984).

    Article  CAS  PubMed  Google Scholar 

  32. Görlich, D. Transport into and out of the cell nucleus. EMBO J. 17, 2721–2727 (1998).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Wente, S. R. & Rout, M. P. The nuclear pore complex and nuclear transport. Cold Spring Harb. Perspect. Biol. 2, a000562 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Niopek, D., Wehler, P., Roensch, J., Eils, R. & Di Ventura, B. Optogenetic control of nuclear protein export. Nat. Commun. 7, 1–9 (2016).

    Article  CAS  Google Scholar 

  35. Elosegui-Artola, A. et al. Mechanical regulation of a molecular clutch defines force transmission and transduction in response to matrix rigidity. Nat. Cell Biol. 18, 540–548 (2016).

    Article  CAS  PubMed  Google Scholar 

  36. Schiller, H. B. et al. β1- and αv-class integrins cooperate to regulate myosin II during rigidity sensing of fibronectin-based microenvironments. Nat. Cell Biol. 15, 625–636 (2013).

    Article  CAS  PubMed  Google Scholar 

  37. Hodel, M. R., Corbett, A. H. & Hodel, A. E. Dissection of a nuclear localization signal. J. Biol. Chem. 276, 1317–1325 (2001).

    Article  CAS  PubMed  Google Scholar 

  38. Lyman, S. K., Guan, T., Bednenko, J., Wodrich, H. & Gerace, L. Influence of cargo size on Ran and energy requirements for nuclear protein import. J. Cell Biol. 159, 55–67 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Ribbeck, K. & Görlich, D. The permeability barrier of nuclear pore complexes appears to operate via hydrophobic exclusion. EMBO J. 21, 2664–2671 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Lowe, A. R. et al. Selectivity mechanism of the nuclear pore complex characterized by single cargo tracking. Nature 467, 600–603 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Kopito, R. B. & Elbaum, M. Nucleocytoplasmic transport: a thermodynamic mechanism. HFSP J. 3, 130–141 (2010).

    Article  CAS  Google Scholar 

  42. Kim, S. & Elbaum, M. A simple kinetic model with explicit predictions for nuclear transport. Biophys. J. 105, 565–569 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Jovanovic-Talisman, T. & Zilman, A. Protein transport by the nuclear pore complex: Simple biophysics of a complex biomachine. Biophys. J. 113, 6–14 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Görlich, D., Seewald, M. J. & Ribbeck, K. Characterization of Ran-driven cargo transport and the RanGTPase system by kinetic measurements and computer simulation. EMBO J. 22, 1088–1100 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Kanwal, C., Mu, S., Kern, S. E. & Lim, C. S. Bidirectional on/off switch for controlled targeting of proteins to subcellular compartments. J. Control. Release 98, 379–393 (2004).

    Article  CAS  PubMed  Google Scholar 

  46. Kim, S. & Elbaum, M. Enzymatically driven transport: a kinetic theory for nuclear export. Biophys. J. 105, 1997–2005 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  48. Wei, S. C. et al. Matrix stiffness drives epithelial–mesenchymal transition and tumour metastasis through a TWIST1–G3BP2 mechanotransduction pathway. Nat. Cell Biol. 17, 678–688 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Zhang, K. et al. Mechanical signals regulate and activate SNAIL1 protein to control the fibrogenic response of cancer-associated fibroblasts. J. Cell Sci. 129, 1989–2002 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Furumatsu, T. et al. Mechanical stretch increases Smad3-dependent CCN2 expression in inner meniscus cells. J. Orthop. Res. 30, 1738–1745 (2012).

    Article  CAS  PubMed  Google Scholar 

  51. Mammoto, A. et al. A mechanosensitive transcriptional mechanism that controls angiogenesis. Nature 457, 1103–1108 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Ishihara, S. et al. Substrate stiffness regulates temporary NF-κB activation via actomyosin contractions. Exp. Cell. Res. 319, 2916–2927 (2013).

    Article  CAS  PubMed  Google Scholar 

  53. Bischoff, F. R., Klebe, C., Kretschmer, J., Wittinghofer, A. & Ponstingl, H. RanGAP1 induces GTPase activity of nuclear Ras-related Ran. Proc. Natl Acad. Sci. USA 91, 2587–2591 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Soderholm, J. F. et al. Importazole, a small molecule inhibitor of the transport receptor importin-β. ACS Chem. Biol. 6, 700–708 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. F, H. et al. HER2/EGFR-AKT signaling switches TGFβ from inhibiting cell proliferation to promoting cell migration in breast cancer. Cancer Res. 78, 6073–6085 (2018).

    Article  Google Scholar 

  56. Singh, S. & Gramolini, A. O. Characterization of sequences in human TWIST required for nuclear localization. BMC Cell Biol. 10, 47 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Lin, D. H. et al. Architecture of the symmetric core of the nuclear pore. Science 352, aaf1015 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Kosinski, J. et al. Molecular architecture of the inner ring scaffold of the human nuclear pore complex. Science 352, 363–365 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Petrovic, S. et al. Architecture of the linker-scaffold in the nuclear pore. bioRxiv https://doi.org/10.1101/2021.10.26.465796 (2021).

  60. Hoelz, A., Debler, E. W. & Blobel, G. The structure of the nuclear pore complex. Annu. Rev. Biochem. 80, 613–643 (2011).

    Article  CAS  PubMed  Google Scholar 

  61. Zimmerli, C. E. et al. Nuclear pores dilate and constrict in cellulo. Science 374, eabd9776 (2021).

    Article  CAS  PubMed  Google Scholar 

  62. Schuller, A. P. et al. The cellular environment shapes the nuclear pore complex architecture. Nature 598, 667–671 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. García-García, M. et al. Mechanical control of nuclear import by Importin-7 is regulated by its dominant cargo YAP. Nat. Commun. 13, 1–21 (2022).

    Article  CAS  Google Scholar 

  64. Frey, S. et al. Surface properties determining passage rates of proteins through nuclear pores. Cell 174, 202–217.e9 (2018).

    Article  CAS  PubMed  Google Scholar 

  65. Infante, E. et al. The mechanical stability of proteins regulates their translocation rate into the cell nucleus. Nat. Phys. 15, 973–981 (2019).

    Article  CAS  Google Scholar 

  66. Aramburu, I. V. & Lemke, E. A. Floppy but not sloppy: interaction mechanism of FG-nucleoporins and nuclear transport receptors. Semin. Cell Dev. Biol. 68, 34–41 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Roca-Cusachs, P. et al. Integrin-dependent force transmission to the extracellular matrix by α-actinin triggers adhesion maturation. Proc. Natl Acad. Sci. USA 110, E1361–E1370 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Liu, H. & Naismith, J. H. An efficient one-step site-directed deletion, insertion, single and multiple-site plasmid mutagenesis protocol. BMC Biotechnol. 8, 91 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  69. Zhang, Q. et al. Nesprins: a novel family of spectrin-repeat-containing proteins that localize to the nuclear membrane in multiple tissues. J. Cell Sci. 114, 4485–4498 (2001).

    Article  CAS  PubMed  Google Scholar 

  70. Kazgan, N., Williams, T., Forsberg, L. J. & Brenman, J. E. Identification of a nuclear export signal in the catalytic subunit of AMP-activated protein kinase. Mol. Biol. Cell 21, 3433–3442 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Yang, J. et al. Twist, a master regulator of morphogenesis, plays an essential role in tumor metastasis. Cell 117, 927–939 (2004).

    Article  CAS  PubMed  Google Scholar 

  72. Oria, R. et al. Force loading explains spatial sensing of ligands by cells. Nature 552, 219–224 (2017).

    Article  CAS  PubMed  Google Scholar 

  73. Lakins, J. N., Chin, A. R. & Weaver, V. M. Exploring the link between human embryonic stem cell organization and fate using tension-calibrated extracellular matrix functionalized polyacrylamide gels. Methods Mol. Biol. 916, 317–350 (2012).

    Article  CAS  PubMed  Google Scholar 

  74. Elosegui-Artola, A. et al. Rigidity sensing and adaptation through regulation of integrin types. Nat. Mater. 13, 631–637 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Rapsomaniki, M. A. et al. EasyFRAP: an interactive, easy-to-use tool for qualitative and quantitative analysis of FRAP data. Bioinformatics 28, 1800–1801 (2012).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank S. Usieto for technical support, and G. Ceada, S. Conti, M. González, S. Garcia-Manyes, M. Rout and the members of the Roca-Cusachs and Trepat lab for useful feedback and discussions. We acknowledge funding from the Spanish Ministry of Science and Innovation (PGC2018-099645-B-I00 to X.T., BFU2016-79916-P and PID2019-110298GB-I00 to P.R.-C.), the European Commission (H2020-FETPROACT-01-2016-731957 to X.T. and P.R.-C.), the European Research Council (Adv-883739 to X.T.), the Generalitat de Catalunya (2017-SGR-1602 to X.T. and P.R.-C.), The prize ‘ICREA Academia’ for excellence in research to P.R.-C., Fundació la Marató de TV3 (201936-30-31 to P.R.-C.) and ‘la Caixa’ Foundation (grant LCF/PR/HR20/52400004 to X.T. and P.R.-C. and fellowships LCF/BQ/DR19/11740009 to I.G.-M. and LCF/BQ/DR19/11740009 to M.M.J.). A.E.M.B. is recipient of a Sir Henry Wellcome fellowship (210887/Z/18/Z). The Institue for Bioengineering of Catalonia (IBEC) is a recipient of a Severo Ochoa Award of Excellence from MINCIN.

Author information

Authors and Affiliations

Authors

Contributions

I.A. and P.R.-C. conceived the study; I.A. and P.R.-C. designed the experiments; I.A., I.G.-M., M.M.J., A.E.M.B. and A.E.-A. performed the experiments; I.A., I.G.-M., N.R.C., J.F.A., L.R., X.T., B.R. and P.R.-C. analysed the experiments; I.G.-M. designed and cloned all constructs, N.R.C. implemented the FLIP analysis software, K.C. and B.R. implemented the kinetic model of transport, and I.A., I.G.-M. and P.R.-C. wrote the manuscript. All authors commented on the manuscript and contributed to it.

Corresponding authors

Correspondence to Ion Andreu or Pere Roca-Cusachs.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks Dennis Discher, Michael Elbaum, André Hoelz and the other, anonymous reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

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 Fluorescence Loss In Photobleaching (FLIP) technique.

a, b) Examples of curves showing fluorescence intensity as a function of time in the nucleus and cytoplasm in FLIP experiments on two example cells transfected with the diffusive 41 kDa construct and seeded on a) 30 kPa in control condition and b) 30kPa with DN-KASH overexpression. Data represent the mean fluorescence intensity of the compartments (nucleus/cytoplasm), normalized with the mean of the whole cell before the beginning of photobleaching, and corrected for background signal. Each curve depicts a representative experiment of one cell each. c, d) Cartoon and equations describing the model used for fitting curves as in A,B, and calculating influx and efflux rates. The model considers the molecules to freely diffuse inside the nuclear and cytoplasmic compartments (see methods). e) Mobile fraction of the L_NLS 41 kDa construct in the nucleus (Nuc) and cytoplasm (Cyt) of cells seeded on 1.5/30 kPa gels. N = 19 cells from 3 independent experiments, lines show mean ± SEM f) For cells seeded on 1.5 and 30 kPa gels, correlation between nuclear to cytosolic ratios of volume, and of areas as measured in confocal slices used for FLIP measurements; regression equation y = 0,6075 x + 0,05375. N = 20 (1.5kPa) and N = 14 (30kPa) cells from 2 independent experiments. Black line shows the linear regression. Source numerical data are available in source data.

Source data

Extended Data Fig. 2 Blocking nuclear to cytoskeletal force transmission with DN-KASH recapitulates the effects of substrate stiffness on transport rates.

a, b) Influx and efflux rates of diffusive constructs for cells seeded on 30 kPa gels, with or without DN-KASH overexpression. In a, both MW (p < 1e-15) and DN KASH (p = 1e-6) effects tested significant. In b, both MW (p < 1e-15) and DN KASH (p = 0,0002) effects tested significant. c, d) Influx and efflux rates of constructs containing L_NLS for cells seeded on 30 kPa gels, with or without DN-KASH overexpression. In c, both MW (p = 0,0025) and DN KASH (p < 1e-15) effects tested significant. In d, both MW (p < 1e-15) and DN KASH (p = 3.4e-10) effects tested significant. In all panels, N = 30 cells from 3 independent experiments. Two-way ANOVA, Šídák’s multiple comparisons test was used to obtain p-values between conditions. Data are mean ± SEM. Source numerical data are available in source data.

Source data

Extended Data Fig. 3 Control experiments on importins and AFM.

a-c) Average fluorescence intensities of nuclear and cytoplasmic areas of cells seeded on substrates of 1.5 or 30 kPa stiffness and immunostained for importin α3 (imp α3) importin α1 (imp α1), and importin β1 (imp β1). N = 90 cells from 3 independent experiments. The effect of substrate stiffness tested significant for importin α3 (p = 7.2e-8) and importin α1 (p = 1.7e-5), but not for importin β1 (p = 0.4971). p-values from Two-way ANOVA d-e) Corresponding example images showing the nucleus (Hoechst) and the distribution of the different importins. f) Corresponding quantification of nuclear to cytoplasmic ratio of importin localization. N = 91,98, 91, 98, 90, 90 cells (from left to right) from 3 independent experiments. p-values from independent two-tailed Mann-Whitney tests. g) N/C ratios of L_NLS-41 kDa or BFP constructs in cells seeded on 1.5 kPa gels before, during, and after nuclear deformation with AFM. h) L_NLS-41 kDa ratios normalized by BFP ratios, from panel g) paired measures. i,j) from g, corresponding paired dot plots of the time points right before and after force application. k) from g, corresponding % change in N/C ratios right after force application for both constructs. In g,h,i,j,k N = 15 cells from 3 independent experiments, p-values were calculated with a two-tailed paired t-test. l) N/C ratios of H_NLS-27 kDa construct in cells seeded on 1.5 kPa gels before, during, and after nuclear deformation with AFM. m) from l, corresponding paired dot plots of the time points right before and after force application. In l, m, N = 15 cells from 3 independent experiments. p-values were calculated with a two-tailed paired t-test. n) Corresponding images of constructs before and during force application, dotted line marks nucleus outline. o) N/C ratios of the L_NLS-41 kDa construct in cells co-transfected with DN-KASH and seeded on 1.5 or 30 kPa gels before, during, and after nuclear deformation with AFM. Data are mean ± SEM. p,q) from o, corresponding paired dot plots of the time points right before and after force application. In o,p,q, N = 15 cells from 3 independent experiments. p-values were calculated with a two-tailed paired t-test, traces of all cells are shown in Extended Data Fig. 8. r) Corresponding images of constructs before and during force application, dotted line marks nucleus outline. Scale bars, 20 µm. Note: in AFM experiments, non-mechanosensitive constructs (BFP and H_NLS) still show a small increase with force, likely due to lensing effects caused by changes in cell shape during indentation. This increase (~6% for BFP, ~2% for H_NLS) is much smaller than that of the mechanosensitive construct (L_NLS 41 kDa, ~14%), see panel k. Panel h in fact shows the response of the L_NLS construct after factoring out the response of BFP. Data are mean ± SEM in all panels. Source numerical data are available in source data.

Source data

Extended Data Fig. 4 Noise levels in N/C ratio measurements.

Relationship between mean N/C ratio as reported in figures, and corresponding coefficient of variation (standard deviation divided by the mean). The different points show all different constructs and conditions reported in the manuscript. Black dots indicate values of overexpressed engineered constructs, red squares indicate values of stained endogenous proteins. Source numerical data are available in source data.

Source data

Extended Data Fig. 5 Effect of the affinity of the NLS signal in influx and efflux rates.

(a-d) Model predictions for N/C ratios (a), mechanosensitivities (b), influx rates (c) and efflux rates (d) for 41 kDa constructs as a function of NLS affinity (modelled by the binding rate kon between the NLS and importin α). e,f) Experimental Influx and efflux rates of 41 kDa constructs containing NLS signals of different affinity for importin β. In both cases (e,f), NLS strength and substrate stiffness effects tested significant (respectively: e) p < 1e-15, p < 1e-15, f) p < 1e-15, p = 2.4e-10). N = 30 cells from 3 independent experiments. p-values from Two-way ANOVA. Data are mean ± SEM. Source numerical data are available in source data.

Source data

Extended Data Fig. 6 Further experiments and modelling results regarding NES constructs.

For M_NES constructs, influx rates (mediated by passive transport) and efflux rates (mediated by facilitated transport) as a function of molecular weight. N = 30 cells from 3 independent experiments. Substrate stiffness effects tested significative in both cases (a) p = 5.1e-13; b) p < 1e-15); MW only tested significative for influx, a) p < 1e-15; b) p = 0.2138). Two-way ANOVA, Šídák’s multiple comparisons test was used to obtain p-values between conditions. Data presented as mean ± SEM. c, d) Model predictions of N/C ratios (c) and mechanosensitivities (d) for an NLS with a binding rate kon of 54 ms−1 as a function of MW. Data are shown for experimentally measured N/C volume ratios (0.29) and for inverted volume ratios (3.5). e, f) Same predictions as in c,d for an NLS with a binding rate kon of 205 ms−1. Note that these predictions simply evaluate the role of N/C volumes on import, they do not explicitly model the export cycle (and hence mechanosensitivities are above and not below 1). Source numerical data are available in source data.

Source data

Extended Data Fig. 7 Mechanosensitivity of transcriptional Regulators.

a-c) For Snail stainings at different conditions, quantifications of N/C ratios on 1.5/30 kPa substrates (a, N = 100 cells from 3 independent repeats), corresponding mechanosensitivities for the 3 different repeats (b), and representative images (c). d-f) For SMAD3 stainings at different conditions, quantifications of N/C ratios on 1.5/30 kPa substrates (d, N = 100 cells from 3 different repeats), corresponding mechanosensitivities for the 3 different repeats (e), and representative images (f). g-i) For GATA2 stainings at different conditions, quantifications of N/C ratios on 1.5/30 kPa substrates (g, N = 90 cells from 3 independent repeats), Corresponding mechanosensitivities for the 3 different repeats (h), and representative images (i). j-l) For NF-κβ stainings at different conditions, quantifications of N/C ratios on 1.5/30 kPa substrates, (j, N = 90 cells from 3 independent repeats), corresponding mechanosensitivities for the 3 different repeats (k), and representative images (l). For a-l, data are presented as mean ± SEM, scale bars correspond to 20 µm, and p-values from corrected multiple two-tailed Mann-Whitney (a,d) and two-tailed Mann-Whitney (g,j) tests. m) Relative gene expression of different genes as assessed with qPCR. Conditions are cells seeded on 1.5 or 30 kPa substrates, overexpressing or not a WT twist1 construct (Ctrl V5-twist1). Gene expression is shown relative to the 1.5 kPa condition without overexpression. n = 2 independent experimental repeats. Source numerical data are available in source data.

Source data

Extended Data Fig. 8 Cell-by-cell fluorescence curves of all AFM experiments.

Plots showing the evolution with time of N/C ratios before, during and after force application to the cell nucleus for all cells measured. a, b) AFM experiments reported in Fig. 3, c) Fig. 5, and d-h) Extended Data Fig. 3. Source numerical data are available in source data.

Source data

Supplementary information

Supplementary Information

Modelling of mechanosensitive nucleocytoplasmic transport: initial conceptual model, kinetic mathematical model of transport (Model Tables 1–3 and Model Fig. 1)

Reporting Summary

Peer Review File

Supplementary Tables

Supplementary Tables 1–3.

Source data

Source Data Fig. 1

Complete source data.

Source Data Fig. 2

Complete source data.

Source Data Fig. 3

Complete source data.

Source Data Fig. 4

Complete source data.

Source Data Fig. 5

Complete source data.

Source Data Fig. 6

Complete source data.

Source Data Extended Data Fig. 1

Complete source data.

Source Data Extended Data Fig. 2

Complete source data.

Source Data Extended Data Fig. 3

Complete source data.

Source Data Extended Data Fig. 4

Complete source data.

Source Data Extended Data Fig. 5

Complete source data.

Source Data Extended Data Fig. 6

Complete source data.

Source Data Extended Data Fig. 7

Complete source data.

Source Data Extended Data Fig. 8

Complete source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Andreu, I., Granero-Moya, I., Chahare, N.R. et al. Mechanical force application to the nucleus regulates nucleocytoplasmic transport. Nat Cell Biol 24, 896–905 (2022). https://doi.org/10.1038/s41556-022-00927-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-022-00927-7

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