Skip to main content
Log in

Dual role of the nucleus in cell migration on planar substrates

  • Original Paper
  • Published:
Biomechanics and Modeling in Mechanobiology Aims and scope Submit manuscript

Abstract

Cell migration is essential to sustain life. There have been significant advances in the understanding of the mechanisms that control cell crawling, but the role of the nucleus remains poorly understood. The nucleus exerts a tight control of cell migration in 3D environments, but its influence in 2D migration on planar substrates remains unclear. Here, we study the role of the cell nucleus in 2D cell migration using a computational model of fish keratocytes. Our results indicate that the apparently minor role played by the nucleus emerges from two antagonist effects: While the nucleus modifies the spatial distributions of actin and myosin in a way that reduces cell velocity (e.g., the nucleus displaces myosin to the sides and front of the cell), its mechanical connection with the cytoskeleton alters the intracellular stresses promoting cell migration. Overall, the favorable effect of the nucleus-cytoskeleton connection prevails, which may explain why regular cells usually move faster than enucleated cells.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Arsenovic PT, Ramachandran I, Bathula K, Zhu R, Narang JD, Noll NA, Lemmon CA, Gundersen GG, Conway DE (2016) Nesprin-2g, a component of the nuclear linc complex, is subject to myosin-dependent tension. Biophys J 110(1):34–43

    Google Scholar 

  • Aubry D, Thiam H, Piel M, Allena R (2015) A computational mechanics approach to assess the link between cell morphology and forces during confined migration. Biomech Model Mechanobiol 14(1):143–157

    Google Scholar 

  • Aureille J, Belaadi N, Guilluy C (2017) Mechanotransduction via the nuclear envelope: a distant reflection of the cell surface. Curr Opin Cell Biol 44:59–67

    Google Scholar 

  • Barnhart EL, Allen GM, Jülicher F, Theriot JA (2010) Bipedal locomotion in crawling cells. Biophys J 98(6):933–942

    Google Scholar 

  • Barnhart EL, Lee KC, Keren K, Mogilner A, Theriot JA (2011) An adhesion-dependent switch between mechanisms that determine motile cell shape. PLoS Biol 9(5):e1001059

    Google Scholar 

  • Barnhart E, Lee KC, Allen GM, Theriot JA, Mogilner A (2015) Balance between cell-substrate adhesion and myosin contraction determines the frequency of motility initiation in fish keratocytes. Proc Natl Acad Sci USA 112(16):5045–5050

    Google Scholar 

  • Barnhart EL, Allard J, Lou SS, Theriot JA, Mogilner A (2017) Adhesion-dependent wave generation in crawling cells. Curr Biol 27(1):27–38

    Google Scholar 

  • Barzilai S, Yadav SK, Morrell S, Roncato F, Klein E, Stoler-Barak L, Golani O, Feigelson SW, Zemel A, Nourshargh S, Alon R (2017) Leukocytes breach endothelial barriers by insertion of nuclear lobes and disassembly of endothelial actin filaments. Cell Rep 18(3):685–699

    Google Scholar 

  • Bausch AR, Ziemann F, Boulbitch AA, Jacobson K, Sackmann E (1998) Local measurements of viscoelastic parameters of adherent cell surfaces by magnetic bead microrheometry. Biophys J 75(4):2038–2049

    Google Scholar 

  • Bergert M, Erzberger A, Desai RA, Aspalter IM, Oates AC, Charras G, Salbreux G, Paluch EK (2015) Force transmission during adhesion-independent migration. Nat Cell Biol 17(4):524

    Google Scholar 

  • Biben T, Kassner K, Misbah C (2005) Phase-field approach to three-dimensional vesicle dynamics. Phys Rev E 72(4):041921

    Google Scholar 

  • Borisy GG, Svitkina TM (2000) Actin machinery: pushing the envelope. Curr Opin Cell Biol 12(1):104–112

    Google Scholar 

  • Calero-Cuenca FJ, Janota CS, Gomes ER (2018) Dealing with the nucleus during cell migration. Curr Opin Cell Biol 50:35–41

    Google Scholar 

  • Camley BA, Zhao Y, Li B, Levine H, Rappel WJ (2013) Periodic migration in a physical model of cells on micropatterns. Phys Rev Lett 111(15):158102

    Google Scholar 

  • Camley BA, Zhang Y, Zhao Y, Li B, Ben-Jacob E, Levine H, Rappel WJ (2014) Polarity mechanisms such as contact inhibition of locomotion regulate persistent rotational motion of mammalian cells on micropatterns. Proc Natl Acad Sci USA 111(41):14770–14775

    Google Scholar 

  • Camley BA, Zhao Y, Li B, Levine H, Rappel WJ (2017) Crawling and turning in a minimal reaction-diffusion cell motility model: coupling cell shape and biochemistry. Phys Rev E 95(1):012401

    Google Scholar 

  • Canham PB (1970) The minimum energy of bending as a possible explanation of the biconcave shape of the human red blood cell. J Theor Biol 26(1):61IN777–76IN881

    Google Scholar 

  • Case LB, Waterman CM (2015) Integration of actin dynamics and cell adhesion by a three-dimensional, mechanosensitive molecular clutch. Nat Cell Biol 17(8):955

    Google Scholar 

  • Cho S, Irianto J, Discher DE (2017) Mechanosensing by the nucleus: from pathways to scaling relationships. J Cell Biol 216(2):305–315

    Google Scholar 

  • Chung J, Hulbert GM (1993) A time integration algorithm for structural dynamics with improved numerical dissipation: the generalized-$\alpha $ method. J Appl Mech 60:371–375

    MathSciNet  MATH  Google Scholar 

  • Deguchi S, Maeda K, Ohashi T, Sato M (2005) Flow-induced hardening of endothelial nucleus as an intracellular stress-bearing organelle. J Biomech 38(9):1751–1759

    Google Scholar 

  • Denais C, Lammerding J (2014) Nuclear mechanics in cancer. In: Schirmer EC, de las Heras JI (eds) Cancer biology and the nuclear envelope. Springer, New York, pp 435–470

    Google Scholar 

  • Denais CM, Gilbert RM, Isermann P, McGregor AL, te Lindert M, Weigelin B, Davidson PM, Friedl P, Wolf K, Lammerding J (2016) Nuclear envelope rupture and repair during cancer cell migration. Science 352(6283):353–358

    Google Scholar 

  • Deveraux S, Allena R, Aubry D (2017) A numerical model suggests the interplay between nuclear plasticity and stiffness during a perfusion assay. J Theor Biol 435:62–77

    MathSciNet  MATH  Google Scholar 

  • Euteneuer U, Schliwa M (1984) Persistent, directional motility of cells and cytoplasmic fragments in the absence of microtubules. Nature 310(5972):58

    Google Scholar 

  • Fournier MF, Sauser R, Ambrosi D, Meister JJ, Verkhovsky AB (2010) Force transmission in migrating cells. J Cell Biol 188(2):287–297

    Google Scholar 

  • Friedl P, Wolf K, Lammerding J (2011) Nuclear mechanics during cell migration. Curr Opin Cell Biol 23(1):55–64

    Google Scholar 

  • Fruleux A, Hawkins RJ (2016) Physical role for the nucleus in cell migration. J Phys Condens Matter 28(36):363002

    Google Scholar 

  • Gomes ER, Jani S, Gundersen GG (2005) Nuclear movement regulated by Cdc42, MRCK, myosin, and actin flow establishes MTOC polarization in migrating cells. Cell 121(3):451–463

    Google Scholar 

  • Gomez H, van der Zee KG (2018) Computational phase-field modeling. Encyclopedia of computational mechanics, 2nd edn. Wiley, New York, pp 1–35

    Google Scholar 

  • Graham DM, Andersen T, Sharek L, Uzer G, Rothenberg K, Hoffman BD, Rubin J, Balland M, Bear JE, Burridge K (2018) Enucleated cells reveal differential roles of the nucleus in cell migration, polarity, and mechanotransduction. J Cell Biol 217(3):895–914

    Google Scholar 

  • Guilak F, Tedrow JR, Burgkart R (2000) Viscoelastic properties of the cell nucleus. Biochem Biophys Res Commun 269(3):781–786

    Google Scholar 

  • Guilluy C, Osborne LD, Van Landeghem L, Sharek L, Superfine R, Garcia-Mata R, Burridge K (2014) Isolated nuclei adapt to force and reveal a mechanotransduction pathway in the nucleus. Nat Cell Biol 16(4):376

    Google Scholar 

  • Harada T, Swift J, Irianto J, Shin JW, Spinler KR, Athirasala A, Diegmiller R, Dingal PDP, Ivanovska IL, Discher DE (2014) Nuclear lamin stiffness is a barrier to 3D migration, but softness can limit survival. J Cell Biol 204(5):669–682

    Google Scholar 

  • Helfrich W (1973) Elastic properties of lipid bilayers: theory and possible experiments. Z Naturforschung C 28(11–12):693–703

    Google Scholar 

  • Hochmuth RM (2000) Micropipette aspiration of living cells. J Biomech 33(1):15–22

    MathSciNet  Google Scholar 

  • Houben F, Willems C, Declercq I, Hochstenbach K, Kamps M, Snoeckx L, Ramaekers F, Broers J (2009) Disturbed nuclear orientation and cellular migration in a-type lamin deficient cells. Biochim Biophys Acta (BBA)-Mol Cell Res 1793(2):312–324

    Google Scholar 

  • Hughes TJ, Cottrell JA, Bazilevs Y (2005) Isogeometric analysis: Cad, finite elements, nurbs, exact geometry and mesh refinement. Comput Methods Appl Mech Eng 194(39):4135–4195

    MathSciNet  MATH  Google Scholar 

  • Jahed Z, Shams H, Mofrad MR (2015) A disulfide bond is required for the transmission of forces through sun-kash complexes. Biophys J 109(3):501–509

    Google Scholar 

  • Jansen K, Whiting C, Hulbert G (2000) Generalized-$\alpha $ method for integrating the filtered Navier–Stokes equations with a stabilized finite element method. Comput Methods Appl Mech Eng 190:305–319

    MathSciNet  MATH  Google Scholar 

  • Jurado C, Haserick JR, Lee J (2005) Slipping or gripping? Fluorescent speckle microscopy in fish keratocytes reveals two different mechanisms for generating a retrograde flow of actin. Mol Biol Cell 16(2):507–518

    Google Scholar 

  • Keren K, Pincus Z, Allen GM, Barnhart EL, Marriott G, Mogilner A, Theriot JA (2008) Mechanism of shape determination in motile cells. Nature 453(7194):475

    Google Scholar 

  • Khatau SB, Hale CM, Stewart-Hutchinson P, Patel MS, Stewart CL, Searson PC, Hodzic D, Wirtz D (2009) A perinuclear actin cap regulates nuclear shape. Proc Natl Acad Sci USA 106(45):19017–19022

    Google Scholar 

  • Khatau SB, Kim DH, Hale CM, Bloom RJ, Wirtz D (2010) The perinuclear actin cap in health and disease. Nucleus 1(4):337–342

    Google Scholar 

  • Khatau SB, Bloom RJ, Bajpai S, Razafsky D, Zang S, Giri A, Wu PH, Marchand J, Celedon A, Hale CM et al (2012) The distinct roles of the nucleus and nucleus-cytoskeleton connections in three-dimensional cell migration. Sci Rep 2:488

    Google Scholar 

  • Kim DH, Khatau SB, Feng Y, Walcott S, Sun SX, Longmore GD, Wirtz D (2012) Actin cap associated focal adhesions and their distinct role in cellular mechanosensing. Sci Rep 2:555

    Google Scholar 

  • Kim DH, Chambliss AB, Wirtz D (2013) The multi-faceted role of the actin cap in cellular mechanosensation and mechanotransduction. Soft Matter 9(23):5516–5523

    Google Scholar 

  • Kim DH, Cho S, Wirtz D (2014) Tight coupling between nucleus and cell migration through the perinuclear actin cap. J Cell Sci 127:2528–2541

    Google Scholar 

  • Lautscham LA, Kämmerer C, Lange JR, Kolb T, Mark C, Schilling A, Strissel PL, Strick R, Gluth C, Rowat AC et al (2015) Migration in confined 3D environments is determined by a combination of adhesiveness, nuclear volume, contractility, and cell stiffness. Biophys J 109(5):900–913

    Google Scholar 

  • Lee JS, Hale CM, Panorchan P, Khatau SB, George JP, Tseng Y, Stewart CL, Hodzic D, Wirtz D (2007) Nuclear lamin a/c deficiency induces defects in cell mechanics, polarization, and migration. Biophys J 93(7):2542–2552

    Google Scholar 

  • Lele TP, Dickinson RB, Gundersen GG (2018) Mechanical principles of nuclear shaping and positioning. J Cell Biol 217(10):3330–3342

    Google Scholar 

  • Li X, Lowengrub J, Rätz A, Voigt A (2009) Solving pdes in complex geometries: a diffuse domain approach. Commun Math Sci 7(1):81

    MathSciNet  MATH  Google Scholar 

  • Li Y, Lovett D, Zhang Q, Neelam S, Kuchibhotla RA, Zhu R, Gundersen GG, Lele TP, Dickinson RB (2015) Moving cell boundaries drive nuclear shaping during cell spreading. Biophys J 109(4):670–686

    Google Scholar 

  • Liu L, Luo Q, Sun J, Song G (2016) Nucleus and nucleus-cytoskeleton connections in 3D cell migration. Exp Cell Res 348(1):56–65

    Google Scholar 

  • Löber J, Ziebert F, Aranson IS (2014) Modeling crawling cell movement on soft engineered substrates. Soft Matter 10(9):1365–1373

    Google Scholar 

  • Lowengrub JS, Rätz A, Voigt A (2009) Phase-field modeling of the dynamics of multicomponent vesicles: spinodal decomposition, coarsening, budding, and fission. Phys Rev E 79(3):031926

    MathSciNet  Google Scholar 

  • Luxton GG, Gomes ER, Folker ES, Vintinner E, Gundersen GG (2010) Linear arrays of nuclear envelope proteins harness retrograde actin flow for nuclear movement. Science 329(5994):956–959

    Google Scholar 

  • Marth W, Voigt A (2014) Signaling networks and cell motility: a computational approach using a phase field description. J Math Biol 69(1):91–112

    MathSciNet  MATH  Google Scholar 

  • Mogilner A, Manhart A (2018) Intracellular fluid mechanics: coupling cytoplasmic flow with active cytoskeletal gel. Annu Rev Fluid Mech 50:347–370

    MathSciNet  MATH  Google Scholar 

  • Mori Y, Jilkine A, Edelstein-Keshet L (2008) Wave-pinning and cell polarity from a bistable reaction-diffusion system. Biophys J 94(9):3684–3697

    Google Scholar 

  • Moure A, Gomez H (2016) Computational model for amoeboid motion: coupling membrane and cytosol dynamics. Phys Rev E 94(4):042423

    MathSciNet  Google Scholar 

  • Moure A, Gomez H (2017) Phase-field model of cellular migration: three-dimensional simulations in fibrous networks. Comput Methods Appl Mech Eng 320:162–197

    MathSciNet  MATH  Google Scholar 

  • Moure A, Gomez H (2018) Three-dimensional simulation of obstacle-mediated chemotaxis. Biomech Model Mechanobiol 17(5):1243–1268

    Google Scholar 

  • Nickaeen M, Novak IL, Pulford S, Rumack A, Brandon J, Slepchenko BM, Mogilner A (2017) A free-boundary model of a motile cell explains turning behavior. PLoS Comput Biol 13(11):e1005862

    Google Scholar 

  • Novak IL, Slepchenko BM, Mogilner A (2008) Quantitative analysis of g-actin transport in motile cells. Biophys J 95(4):1627–1638

    Google Scholar 

  • Oliver T, Dembo M, Jacobson K (1999) Separation of propulsive and adhesive traction stresses in locomoting keratocytes. J Cell Biol 145(3):589–604

    Google Scholar 

  • Palmieri B, Bresler Y, Wirtz D, Grant M (2015) Multiple scale model for cell migration in monolayers: elastic mismatch between cells enhances motility. Sci Rep 5(11):745

    Google Scholar 

  • Petrášek Z, Hoege C, Mashaghi A, Ohrt T, Hyman AA, Schwille P (2008) Characterization of protein dynamics in asymmetric cell division by scanning fluorescence correlation spectroscopy. Biophys J 95(11):5476–5486

    Google Scholar 

  • Petrie RJ, Koo H, Yamada KM (2014) Generation of compartmentalized pressure by a nuclear piston governs cell motility in a 3D matrix. Science 345(6200):1062–1065

    Google Scholar 

  • Petrie RJ, Harlin HM, Korsak LIT, Yamada KM (2017) Activating the nuclear piston mechanism of 3D migration in tumor cells. J Cell Biol 216(1):93–100

    Google Scholar 

  • Pollard TD, Blanchoin L, Mullins RD (2000) Molecular mechanisms controlling actin filament dynamics in nonmuscle cells. Annu Rev Biophys Biomol Struct 29(1):545–576

    Google Scholar 

  • Postma M, Bosgraaf L, Loovers HM, Van Haastert PJ (2004) Chemotaxis: signalling modules join hands at front and tail. EMBO Rep 5(1):35–40

    Google Scholar 

  • Ramdas NM, Shivashankar G (2015) Cytoskeletal control of nuclear morphology and chromatin organization. J Mol Biol 427(3):695–706

    Google Scholar 

  • Raynaud F, Ambühl ME, Gabella C, Bornert A, Sbalzarini IF, Meister JJ, Verkhovsky AB (2016) Minimal model for spontaneous cell polarization and edge activity in oscillating, rotating and migrating cells. Nat Phys 12(4):367–373

    Google Scholar 

  • Riaz M, Versaevel M, Mohammed D, Glinel K, Gabriele S (2016) Persistence of fan-shaped keratocytes is a matrix-rigidity-dependent mechanism that requires $\alpha $ 5 $\beta $ 1 integrin engagement. Sci Rep 6(34):141

    Google Scholar 

  • Rowat A, Lammerding J, Ipsen JH (2006) Mechanical properties of the cell nucleus and the effect of emerin deficiency. Biophys J 91(12):4649–4664

    Google Scholar 

  • Rubinstein B, Fournier MF, Jacobson K, Verkhovsky AB, Mogilner A (2009) Actin-myosin viscoelastic flow in the keratocyte lamellipod. Biophys J 97(7):1853–1863

    Google Scholar 

  • Serrano-Alcalde F, García-Aznar JM, Gómez-Benito MJ (2017) The role of nuclear mechanics in cell deformation under creeping flows. J Theor Biol 432:25–32

    Google Scholar 

  • Shao D, Levine H, Rappel WJ (2012) Coupling actin flow, adhesion, and morphology in a computational cell motility model. Proc Natl Acad Sci USA. 109(18):6851–6856

    Google Scholar 

  • Stroka KM, Jiang H, Chen SH, Tong Z, Wirtz D, Sun SX, Konstantopoulos K (2014) Water permeation drives tumor cell migration in confined microenvironments. Cell 157(3):611–623

    Google Scholar 

  • Swift J, Discher DE (2014) The nuclear lamina is mechano-responsive to ECM elasticity in mature tissue. J Cell Sci 127(14):3005–3015

    Google Scholar 

  • Teigen KE, Li X, Lowengrub J, Wang F, Voigt A (2009) A diffuse-interface approach for modeling transport, diffusion and adsorption/desorption of material quantities on a deformable interface. Commun Math Sci 4(7):1009

    MathSciNet  MATH  Google Scholar 

  • Thiam HR, Vargas P, Carpi N, Crespo CL, Raab M, Terriac E, King MC, Jacobelli J, Alberts AS, Stradal T et al (2016) Perinuclear arp2/3-driven actin polymerization enables nuclear deformation to facilitate cell migration through complex environments. Nat Commun 7(10):997

    Google Scholar 

  • Tinevez JY, Schulze U, Salbreux G, Roensch J, Joanny JF, Paluch E (2009) Role of cortical tension in bleb growth. Proc Natl Acad Sci USA 106(44):18581–18586

    Google Scholar 

  • Tocco VJ, Li Y, Christopher KG, Matthews JH, Aggarwal V, Paschall L, Luesch H, Licht JD, Dickinson RB, Lele TP (2018) The nucleus is irreversibly shaped by motion of cell boundaries in cancer and non-cancer cells. J Cell Physiol 233(2):1446–1454

    Google Scholar 

  • Uhler C, Shivashankar G (2018) Nuclear mechanopathology and cancer diagnosis. Trends Cancer 4(4):320–331

    Google Scholar 

  • Wu J, Kent IA, Shekhar N, Chancellor T, Mendonca A, Dickinson RB, Lele TP (2014) Actomyosin pulls to advance the nucleus in a migrating tissue cell. Biophys J 106(1):7–15

    Google Scholar 

  • Xia Y, Pfeifer CR, Discher DE (2019) Nuclear mechanics during and after constricted migration. Acta Mech Sin 35(2):299–308

    Google Scholar 

  • Zeng X, Li S (2012) A three dimensional soft matter cell model for mechanotransduction. Soft Matter 8(21):5765–5776

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Adrian Moure.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

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

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 3781 KB)

Supplementary material 2 (avi 14042 KB)

Appendix

Appendix

1.1 Appendix 1: Additional model details

The cell volume conservation is weakly imposed by the wave-pinning model used for actin dynamics. Since \(N_A=\int _{\Omega }\psi (\rho _f+\rho _g) \mathrm {\,d}\Omega\) is kept constant, the model results show variations of cell area of at most \(\sim \!15\%\). The nuclear volume conservation is imposed with the force \({{\mathbf {F}}^n_{\mathrm{vol}}}\) (see Eq. (9)), which is equivalent to the force \(\Lambda ^{\star }_n[V_{n0}-V_n(t)]\varepsilon \vert \nabla \nu \vert ^2\mathbf {n}_{\Gamma _n}\). There, \(\varepsilon \vert \nabla \nu \vert ^2\) localizes the force to the nuclear envelope, \(\mathbf {n}_{\Gamma _n}\) is the unit normal vector to the nuclear envelope, and \(\Lambda ^{\star }_n\) is a constant.

Some of the forces in Eqs. (8) and (9) can be derived from an energy functional making use of the phase-field formulation (Gomez and van der Zee 2018; Moure and Gomez 2017). Let us consider a generic functional \({\mathscr {F}}_a[\phi ,\nu ]=\int _{\Omega }\Psi _a(\phi ,\nu ) \mathrm {\,d}\Omega\). The forces applied on the cell (i.e., the cytosol) may be written as \({\mathbf {F}}^c_a=\frac{\delta {\mathscr {F}}_a}{\delta \phi }\nabla \phi\) and the forces applied on the nucleus as \({\mathbf {F}}^n_a=\frac{\delta {\mathscr {F}}_a}{\delta \nu }\nabla \nu\). In particular, \({\mathbf {F}}^c_{\mathrm{def}}\) and \({\mathbf {F}}^n_{\mathrm{def}}\) may be derived from the repulsive potential \({\mathscr {F}}_\mathrm{def}[\phi ,\nu ]=\int _{\Omega }\eta _{\mathrm{def}}(1-\phi )^2\nu ^2 \mathrm {\,d}\Omega\), such that \({\mathbf {F}}^c_{\mathrm{def}}=-\eta _{\mathrm{def}}2(1-\phi )\nu ^2 \nabla \phi\) and \({\mathbf {F}}^n_{\mathrm{def}}=\eta _{\mathrm{def}}2(1-\phi )^2\nu \nabla \nu\).

From Eq. (9), we can compute the nucleus velocity as

$$\begin{aligned} {\mathbf {u}}_n &= \frac{1}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2}\bigg \{\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2{\mathbf {u}}_c \\&+ \bigg [\gamma _n\left( \varepsilon \nabla ^2\nu -\frac{G'(\nu )}{\varepsilon }\right) \\&+\Lambda _n\left( V_n(t)-V_{n0}\right) \vert \nabla \nu \vert \\&+ \eta _{\mathrm{def}} 2(1-\phi )^2\nu + \eta _{\mathrm{poly}}(1-\phi )^2\nu \bigg ] \nabla \nu \bigg \}. \end{aligned}$$
(10)

By substituting \({\mathbf {u}}_n\) in Eq. (2), we obtain

$$\begin{aligned}&\frac{\partial \nu }{\partial t} + \frac{\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2}{\mathbf {u}}_c \cdot \nabla \nu \\&\quad = \Gamma _{n}\left( \varepsilon \nabla ^2\nu -\frac{G'(\nu )}{\varepsilon } + c_{\nu } \varepsilon |\nabla \nu |\right) \\&\qquad -\frac{\vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2} \bigg [\gamma _n\left( \varepsilon \nabla ^2\nu -\frac{G'(\nu )}{\varepsilon }\right) \\&\qquad +\Lambda _n\left( V_n(t)-V_{n0}\right) \vert \nabla \nu \vert \\&\qquad + \eta _{\mathrm{def}} 2(1-\phi )^2\nu + \eta _{\mathrm{poly}} (1-\phi )^2\nu \bigg ]. \end{aligned}$$
(11)

In our numerical simulations, we use Eq. (11) instead of Eqs. (2) and (9), to speed up the calculations. The velocity \({\mathbf {u}}_n\), required to compute \({\mathbf {F}}^c_{\mathrm{AC}}\) in Eq. (8), is taken from Eq. (10).

1.2 Appendix 2: Numerical formulation

We consider a fixed computational domain \(\Omega\). We solve Eqs. (1), (3), (4), (7), (8), and (11) on \(\Omega\) using isogeometric analysis (Hughes et al. 2005), a finite element method that employs B-splines as basis functions. Note that we do not need to discretize Eq. (7). The numerical formulation is analogous to that explained in Moure and Gomez (2017). We derive a weak form of Eqs. (1), (3), (4), (8), and (11) by multiplying them with weighting functions, integrating over \(\Omega\), and integrating by parts under the assumption of periodic boundary conditions. The weak form can be written as:

$$\begin{aligned}&\int _{\Omega }w_1\frac{\partial \phi }{\partial t}\mathrm{d}\Omega + \int _{\Omega }w_1\,{\mathbf {u}}_c\cdot \nabla \phi \mathrm{d}\Omega + \int _{\Omega }\Gamma _{\phi }\varepsilon \nabla w_1\cdot \nabla \phi \mathrm{d}\Omega \\&\quad +\int _{\Omega }w_1\Gamma _{\phi }\frac{G'(\phi )}{\varepsilon }\mathrm{d}\Omega - \int _{\Omega }\Gamma _{\phi }\varepsilon \nabla w_1\cdot \nabla \phi \mathrm{d}\Omega \\&\quad -\int _{\Omega }w_1\frac{\Gamma _{\phi }\varepsilon }{\vert \nabla \phi \vert }\nabla \phi \cdot \nabla (\vert \nabla \phi \vert )\mathrm{d}\Omega = 0, \end{aligned}$$
(12)
$$\begin{aligned}&\int _{\Omega }w_2\frac{\partial (\psi \rho _m)}{\partial t}\mathrm{d}\Omega - \int _{\Omega }\nabla w_2\cdot {\mathbf {u}}_c\psi \rho _m\mathrm{d}\Omega \\&\quad + \int _{\Omega }{D}_{m}\psi \nabla w_2 \cdot \nabla \rho _m\mathrm{d}\Omega = 0, \end{aligned}$$
(13)
$$\begin{aligned}&\int _{\Omega }w_3\frac{\partial (\psi \rho _f)}{\partial t}\mathrm{d}\Omega - \int _{\Omega }\nabla w_3\cdot {\mathbf {u}}_c\psi \rho _f\mathrm{d}\Omega \\&\quad + \int _{\Omega }{D}_{f}\psi \nabla w_3 \cdot \nabla \rho _f\mathrm{d}\Omega -\int _{\Omega }w_3 \psi f_\rho (\rho _f,\rho _g)\mathrm{d}\Omega = 0, \end{aligned}$$
(14)
$$\begin{aligned}&-\int _{\Omega }\mu \psi \nabla \mathbf {w}_4 :(\nabla {\mathbf {u}}_c +\nabla {\mathbf {u}}_c^T)\mathrm{d}\Omega - \int _{\Omega }{\eta }_{m}\psi {\rho }_{m}\nabla \mathbf {w}_4 :{\mathbf {I}}\,\mathrm{d}\Omega \\&\quad + \int _{\Omega }{\eta }_{f}\psi {\rho }_{f}\nabla \mathbf {w}_4 :\nabla \phi \otimes \nabla \phi \mathrm{d}\Omega - \int _{\Omega }\varsigma _c\psi \mathbf {w}_4\cdot {\mathbf {u}}_c\mathrm{d}\Omega \\&\quad -\int _{\Omega }\gamma _\phi \,\mathbf {w}_4\cdot \nabla \phi \left( \varepsilon \nabla ^2\phi -\frac{G'(\phi )}{\varepsilon }\right) \mathrm{d}\Omega \\&\quad -\int _{\Omega } \tau _{\mathrm{AC}}\varepsilon \vert \nabla \nu \vert ^2 \mathbf {w}_4\cdot {\mathbf {u}}_c\mathrm{d}\Omega + \int _{\Omega } \tau _{\mathrm{AC}}\varepsilon \vert \nabla \nu \vert ^2 \mathbf {w}_4\cdot {\mathbf {u}}_n\mathrm{d}\Omega \\&\quad -\int _{\Omega } \eta _{\mathrm{def}} 2(1-\phi )\nu ^2 \mathbf {w}_4\cdot \nabla \phi \mathrm{d}\Omega = 0, \end{aligned}$$
(15)
$$\begin{aligned}&\int _{\Omega }w_5\frac{\partial \nu }{\partial t}\mathrm{d}\Omega + \int _{\Omega }w_5\frac{\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2}{\mathbf {u}}_c\cdot \nabla \nu \mathrm{d}\Omega \\&\quad + \int _{\Omega }\Gamma _{n}\varepsilon \nabla w_5\cdot \nabla \nu \mathrm{d}\Omega \\&\quad +\int _{\Omega }w_5\Gamma _{n}\frac{G'(\nu )}{\varepsilon }\mathrm{d}\Omega - \int _{\Omega }\Gamma _{n}\varepsilon \nabla w_5\cdot \nabla \nu \mathrm{d}\Omega \\&\quad -\int _{\Omega }w_5\frac{\Gamma _{n}\varepsilon }{\vert \nabla \nu \vert }\nabla \nu \cdot \nabla (\vert \nabla \nu \vert )\mathrm{d}\Omega \\&\quad + \int _{\Omega }w_5\frac{\vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2}\gamma _{n}\left( \varepsilon \nabla ^2\nu -\frac{G'(\nu )}{\varepsilon } \right) \mathrm{d}\Omega \\&\quad +\int _{\Omega }w_5\frac{\vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2} \Lambda _n\left( V_n(t)-V_{n0}\right) \vert \nabla \nu \vert \mathrm{d}\Omega \\&\quad +\int _{\Omega }w_5\frac{\vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2} \eta _{\mathrm{def}} 2(1-\phi )^2\nu \mathrm{d}\Omega \\&\quad +\int _{\Omega }w_5\frac{\vert \nabla \nu \vert ^2}{\varsigma _n\nu +\tau _\mathrm{AC}\varepsilon \vert \nabla \nu \vert ^2} \eta _{\mathrm{poly}} (1-\phi )^2\nu \mathrm{d}\Omega = 0, \end{aligned}$$
(16)

where \(\{w_i({\mathbf {x}})\}_{i=1,\ldots ,5}\) are the weighting functions. In Eqs. (13)–(15), \(\psi\) and \({\mathbf {u}}_n\) must be replaced by \(\phi -\nu\) and the expression given in Eq. (10), respectively. We have not substituted \(\psi\) and \({\mathbf {u}}_n\) into Eqs. (13)–(15) to keep the notation clear. Note that Eqs. (12) and (16) become singular in regions far from the interface due to the curvature term (Biben et al. 2005). To address this issue, we propose a simple method similar to that used in Biben et al. (2005). We multiply the curvature \(c_{\iota }\) by \(\alpha _{\iota } = 0.5+0.5\mathrm {\,tanh}[\mathrm {\,log}_{10}(\vert \nabla \iota \vert +\Upsilon _{\iota })]\), with \(\iota =\phi ,\nu\); see Eqs. (1) and (11). In our computations, we took \(\Upsilon _{\phi }=7\) and \(\Upsilon _{\nu }=3.5\). This regularizes the equations without altering the results. For similar reasons, in Eqs. (15) and (16) we replace \(\varsigma _c\psi\) and \(\varsigma _n\nu\) by \(\varsigma _c\psi +\varsigma _{\epsilon }\) and \(\varsigma _n\nu +\varsigma _{\epsilon }\), respectively, to avoid singularities and improve the condition number of the linear system. Here, \(\varsigma _{\epsilon }\) is a small constant (we took \(\varsigma _{\epsilon }={0.2}\,\hbox {pN s}\,\upmu \hbox {m}^{-3}\)) such that \(0<\varsigma _{\epsilon }<<\varsigma _c\). The Galerkin form is obtained by replacing the unknowns and weighting functions with discrete approximations. Let us denote the basis functions as \(N^A\). We replace \(\phi ({\mathbf {x}},t)\) and \(w_1({\mathbf {x}})\) with

$$\begin{aligned} \phi ^h({\mathbf {x}},t)=\sum _A \phi ^A(t)N^A({\mathbf {x}}), \quad \quad w^h_1({\mathbf {x}})=\sum _A w_1^A N^A({\mathbf {x}}), \end{aligned}$$
(17)

respectively, where the \(\phi ^A\)s are referred to as control variables. The spatial discretization of the rest of the unknowns and weighting functions is analogous.

Our semi-implicit time-stepping scheme is based on the generalized-\(\alpha\) method (Chung and Hulbert 1993; Jansen et al. 2000). The algorithm presents two particular features: first, the explicit, rather than implicit, treatment of the term \(\rho _g\); see Eq. (7) and second, the redefinition of the tangent matrix and residual vector to improve the condition number of the tangent matrix. The procedure consists of replacing certain rows of the tangent matrix and the associated elements of the residual vector; see Moure and Gomez (2017) for more details. The tangent matrix rows and the residual vector elements to be replaced are associated with the control variables of \(\rho _m\) and \(\rho _f\) and are determined according to the value of \(\psi\) (unlike the procedure explained in Moure and Gomez (2017), which employs the value of \(\phi\)). We use the Newton–Raphson method to solve the resulting nonlinear system. In our simulations, we employ a periodic domain of \(50\times {50}\upmu \hbox {m}^2\), which is meshed with \(200\times 200\)\({\mathscr {C}}^1\)-continuous quadratic B-spline elements. We use a time step of \({0.05}\hbox {s}\).

1.3 Appendix 3: Parameter estimation

The parameters used in our simulations are listed in Table 3. We keep constant all the parameters except those related to the nucleus dynamics. Unless otherwise stated, the parameters \(\tau _{\mathrm{AC}}\), \(\eta _{\mathrm{def}}\), \(\eta _{\mathrm{poly}}\), \(\gamma _n\), and \(\varsigma _n\) take the values indicated in Table 3.

The parameter \(\Lambda _n\) is estimated from the value of the nucleus bulk modulus \(K_n\). In our 2D problem, we assume plane stress, which leads to the relation \(K_n=\frac{E_n}{2(1-2\nu _n)}\), where \(E_n\) is the nucleus Young’s modulus and \(\nu _n\) is the nucleus Poisson’s ratio. The force \({\mathbf {F}}^n_{\mathrm{vol}}\) can be understood as a pressure acting on the nucleus, induced by nuclear volume changes. Therefore, we can compare \({\mathbf {F}}^n_\mathrm{vol}=\Lambda ^{\star }_n[V_{n0}-V_n(t)]\varepsilon \vert \nabla \nu \vert ^2\mathbf {n}_{\Gamma _n}\) (see “Appendix 1”) with the definition of the bulk modulus: \(dP=-K_n\frac{dV}{V}\), where P is pressure and V is volume. If we neglect the term \(\varepsilon \vert \nabla \nu \vert ^2\), which localizes the force \({\mathbf {F}}^n_{\mathrm{vol}}\) to the nuclear envelope, we identify \(dV=-[V_{n0}-V_n(t)]\) and \(\frac{K_n}{V}=\Lambda ^{\star }_n\). If we take \(V=V_{n0}\), we obtain \(\Lambda ^{\star }_n=\frac{E_n}{2(1-2\nu _n)V_{n0}}\). Finally, the parameter \(\Lambda _n\) is defined as \(\Lambda _n=\varepsilon \Lambda ^{\star }_n=\frac{\varepsilon E_n}{2(1-2\nu _n)V_{n0}}\). Note that the latter expression is not dimensionally consistent because we omitted the term \(\varepsilon \vert \nabla \nu \vert ^2\). If we assume that \(\varepsilon \vert \nabla \nu \vert ^2={1}{\upmu \hbox {m}^{-1}}\) and incorporate it into the previous derivation, the expression for \(\Lambda _n\) becomes dimensionally consistent.

Table 3 List of model parameters

1.4 Appendix 4: Initial conditions

The initial conditions are defined as

$$\begin{aligned}&\phi ({\mathbf {x}},0)=\phi _0({\mathbf {x}})=0.5-0.5\mathrm {\,tanh}\left[ \frac{2\sqrt{2}}{\varepsilon }\left( d_c({\mathbf {x}})-R_c\right) \right] , \end{aligned}$$
(18)
$$\begin{aligned}&\nu ({\mathbf {x}},0)=\nu _0({\mathbf {x}})=0.5-0.5\mathrm {\,tanh}\left[ \frac{2\sqrt{2}}{\varepsilon }\left( d_c({\mathbf {x}})-R_n\right) \right] , \end{aligned}$$
(19)
$$\begin{aligned}&\rho _m({\mathbf {x}},0)=\rho _{m0}\left[ \phi _0({\mathbf {x}})-\nu _0({\mathbf {x}})\right] , \end{aligned}$$
(20)
$$\begin{aligned}&\rho _f({\mathbf {x}},0)= \left\{ \begin{array}{ll} \rho _{f0}^M \left[ \phi _0({\mathbf {x}})-\nu _0({\mathbf {x}})\right] &\;\text {if }y\ge y_c, \\ \rho _{f0}^m \left[ \phi _0({\mathbf {x}})-\nu _0({\mathbf {x}})\right] &\;\text {if }y<y_c, \end{array} \right. \end{aligned}$$
(21)
$$\begin{aligned}&\rho _g({\mathbf {x}},0)=\rho _{g0}\left[ \phi _0({\mathbf {x}})-\nu _0({\mathbf {x}})\right] , \end{aligned}$$
(22)
$$\begin{aligned}&{\mathbf {u}}_c({\mathbf {x}},0)=0, \end{aligned}$$
(23)

where \(d_c({\mathbf {x}})\) is the distance to the center of the cell, denoted here by \(\mathbf {C}=(x_c,y_c)\). \(R_c\) and \(R_n\) are the initial radius of the cell and the nucleus, respectively. The initial myosin and actin densities are defined by the parameters \(\rho _{m0}\), \(\rho _{f0}^M\), \(\rho _{f0}^m\), and \(\rho _{g0}\). Unless otherwise stated, we use the values \(R_c={10.5}\upmu \hbox {m}\), \(R_n={4.3}\,\upmu \hbox {m}\), \(\rho _{m0}={0.3}\,\upmu \hbox {m}^{-2}\), \(\rho _{f0}^M={1.5}\,\upmu \hbox {m}^{-2}\), \(\rho _{f0}^m={0.17}\,\upmu \hbox {m}^{-2}\), and \(\rho _{g0}={1.76}\,\upmu \hbox {m}^{-2}\).

In the simulations corresponding to a cell spreading on a \(75\times {5}\,\upmu \hbox {m}^2\) adhesive micropattern, we take \(R_c={9.5}\,\upmu \hbox {m}\), \(R_n={5.5}\,\upmu \hbox {m}\), \(\rho _{m0}={0.165}\,\upmu \hbox {m}^{-2}\), \(\rho _{f0}^M=\rho _{f0}^m={2.4}\,\upmu \hbox {m}^{-2}\), and \(\rho _{g0}={5}\,\upmu \hbox {m}^{-2}\).

1.5 Appendix 5: Simulations of cells with fixed nucleus geometry

We fixed the nucleus shape by imposing a uniform nucleus velocity in Eq. (2), i.e., \({\mathbf {u}}_n({\mathbf {x}},t)={\mathbf {u}}_n^{\star }(t)\). Thus, we solve Eq. (2) instead of Eq. (11) in our system of equations. The velocity \({\mathbf {u}}^{\star }_n(t)\) is defined as a function of the distance between the nucleus and the cell’s rear, \(d^{\star }\!(t)\), such that

$${\mathbf{u}}_{n}^{{ \star}} (t) = \left\{ {\begin{array}{*{20}l} {1.1{\mathbf{u}}_{c}^{R} (t)\left[ {2\left( {\frac{{d^{{ \star }} (t)}}{{d_{M} }}} \right)^{3} - 3\left( {\frac{{d^{{ \star }} (t)}}{{d_{M} }}} \right)^{2} + 1} \right]} & {{\text{if}}\;0 < d^{{ \star }} (t) < d_{M},} \\ 0 & {{\text{if}}\;d^{{ \star }} (t) \ge d_{M},} \\ \end{array} } \right.$$
(24)

where \(d_M={2}\,\upmu \hbox {m}\) and \({\mathbf {u}}_c^{R}(t)\) is the y-velocity of the membrane at the cell’s rear.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Moure, A., Gomez, H. Dual role of the nucleus in cell migration on planar substrates. Biomech Model Mechanobiol 19, 1491–1508 (2020). https://doi.org/10.1007/s10237-019-01283-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10237-019-01283-6

Keywords

Navigation