Skip to main content
Log in

Abstract

In this paper a minimal, one-dimensional, two-phase, viscoelastic, reactive, flow model for a crawling cell is presented. Two-phase models are used with a variety of constitutive assumptions in the literature to model cell motility. We use an upper-convected Maxwell model and demonstrate that even the simplest of two-phase, viscoelastic models displays features relevant to cell motility. We also show care must be exercised in choosing parameters for such models as a poor choice can lead to an ill–posed problem. A stability analysis reveals that the initially stationary, spatially uniform strip of cytoplasm starts to crawl in response to a perturbation which breaks the symmetry of the network volume fraction or network stress. We also demonstrate numerically that there is a steady travelling-wave solution in which the crawling velocity has a bell-shaped dependence on adhesion strength, in agreement with biological observation.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Alt W, Dembo M (1999) Cytoplasm dynamics and cell motion: two-phase flow models. Math Biosci 156(1–2):207–228

    Article  MATH  Google Scholar 

  • Alt W, Tranquillo R (1995) Basic morphogenetic system modeling shape changes of migrating cells: how to explain fluctuating lamellipodial dynamics. J Biol Syst 3(4):905–916

    Article  Google Scholar 

  • Alt W, Bock M, Möhl C (2010) Cell mechanics: from single-scale based models to multiscale modeling. In: Chauvière A, Preziosi L, Verdier C (eds) Coupling of cytoplasm and adhesion dynamics determines cell polarization and locomotion. Chapman and Hall/CRC, Boca Raton, London, New York, p 89–132

  • Balay S, Gropp WD, Curfman McInnes L, Smith BF (1997) Modern software tools for scientific computing. In: Arge E, Bruaset AM, Langtangen HP (eds) Efficient management of parallelism in object oriented numerical software libraries. Birkhäuser Press, Boston, p 163–202

  • Balay S, Brown J, Buschelman K, Eijkhout V, Gropp WD, Kaushik D, Knepley MG, Curfman McInnes L, Smith BF, Zhang H (2010) PETSc users manual. anl-95/11—revision 3.1. Argonne National Laboratory

  • Balay S, Brown J, Buschelman K, Gropp WD, Kaushik D, Knepley MG, Curfman McInnes L, Smith BF, Zhang H (2011) PETSc Web page. http://www.mcs.anl.gov/petsc.

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

    Article  Google Scholar 

  • Borm B, Requardt R, Herzog V, Kirfel G (2005) Membrane ruffles in cell migration: indicators of inefficient lamellipodia adhesion and compartments of actin filament reorganization. Exp Cell Res 302(1):83–95

    Article  Google Scholar 

  • Bray D (2001) Cell movements: from molecules to motility. Garland Science, New York

    Google Scholar 

  • Burnette D, Manley S, Sengupta P, Sougrat R, Davidson M, Kachar B, Lippincott-Schwartz J (2011) A role for actin arcs in the leading-edge advance of migrating cells. Nat Cell Biol 13(4):371–382

    Article  Google Scholar 

  • Burton K, Park J, Taylor D (1999) Keratocytes generate traction forces in two phases. Mol Biol Cell 10(11):3745–3769

    Article  Google Scholar 

  • Callan-Jones A, Jülicher F (2011) Hydrodynamics of active permeating gels. New J Phys 13:093027

    Article  Google Scholar 

  • Callan-Jones A, Joanny J, Prost J (2008) Viscous-fingering-like instability of cell fragments. Phys Rev Lett 100(25):258106

    Article  Google Scholar 

  • Cogan N, Guy R (2010) Multiphase flow models of biogels from crawling cells to bacterial biofilms. HFSP J 4(1):11–25

    Article  Google Scholar 

  • Cogan NG, Keener JP (2004) The role of the biofilm matrix in structural development. Math Med Biol 21(2):147–166

    Google Scholar 

  • Dembo M, Harlow F (1986) Cell motion, contractile networks, and the physics of interpenetrating reactive flow. Biophys J 50(1):109–121

    Article  Google Scholar 

  • Dembo M, Harlow F, Alt W (1984) Cell surface dynamics: concepts and models. In: Perelson AS, DeLisi C, Wiegel FW (eds) The biophysics of cell surface motility. Marcel Dekker, New York, p 495–542

  • Doubrovinski K, Kruse K (2007) Self-organization of treadmilling filaments. Phys Rev Lett 99(22):228104

    Article  Google Scholar 

  • Doubrovinski K, Kruse K (2008) Cytoskeletal waves in the absence of molecular motors. Eur Phys Lett 83:18003

    Article  Google Scholar 

  • Doubrovinski K, Kruse K (2010) Self-organization in systems of treadmilling filaments. Eur Phys J E 31(1):95–104

    Article  Google Scholar 

  • Doubrovinski K, Kruse K (2011) Cell motility resulting from spontaneous polymerization waves. Phys Rev Lett 107(25):258103

    Article  Google Scholar 

  • Du X, Doubrovinski K, Osterfield M (2012) Self-organized cell motility from motor-filament interactions. Biophys J 102:1738–1745

    Article  Google Scholar 

  • Eriksson K (1996) Computational differential equations. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Giannone G, Dubin-Thaler B, Döbereiner H, Kieffer N, Bresnick A, Sheetz M (2004) Periodic lamellipodial contractions correlate with rearward actin waves. Cell 116(3):431–443

    Article  Google Scholar 

  • Gracheva M, Othmer H (2004) A continuum model of motility in ameboid cells. Bull Math Biol 66(1):167–193

    Article  MathSciNet  Google Scholar 

  • Hanein D, Horwitz A (2012) The structure of cell-matrix adhesions: the new frontier. Curr Opin Cell Biol 24:134–140

    Article  Google Scholar 

  • Herant M, Marganski W, Dembo M (2003) The mechanics of neutrophils: synthetic modeling of three experiments. Biophys J 84(5):3389–3413

    Article  Google Scholar 

  • Herant M, Heinrich V, Dembo M (2006) Mechanics of neutrophil phagocytosis: experiments and quantitative models. J Cell Sci 119:1903–1913

    Google Scholar 

  • Hinz B, Alt W, Johnen C, Herzog V, Kaiser H (1999) Quantifying lamella dynamics of cultured cells by SACED, a new computer-assisted motion analysis. Exp Cell Res 251(1):234–243

    Article  Google Scholar 

  • Hodge N, Papadopoulos P (2012) Continuum modeling and numerical simulation of cell motility. J Math Biol 64:1253–1279

    Article  MathSciNet  MATH  Google Scholar 

  • Joanny J, Jülicher F, Kruse K, Prost J (2007) Hydrodynamic theory for multi-component active polar gels. New J Phys 9:422

    Article  Google Scholar 

  • Jülicher F, Kruse K, Prost J, Joanny J (2007) Active behavior of the cytoskeleton. Phys Rep 449:3–28

    Article  MathSciNet  Google Scholar 

  • Keren K, Yam P, Kinkhabwala A, Mogilner A, Theriot J (2009) Intracellular fluid flow in rapidly moving cells. Nat Cell Biol 11(10):1219–1224

    Article  Google Scholar 

  • Kimpton L, Whiteley J, Waters S, King J, Oliver J (2012) Multiple travelling-wave solutions in a minimal model for cell motility. Math Med Biol Adv Access. doi:10.1093/imammb/dqs023 (published July 11, 2012)

  • King J, Oliver J (2005) Thin-film modelling of poroviscous free surface flows. Eur J Appl Math 16(04):519–553

    Article  MathSciNet  MATH  Google Scholar 

  • Knapp D, Barocas V, Moon A, Yoo K, Petzold L, Tranquillo R (1997) Rheology of reconstituted type I collagen gel in confined compression. J Rheol 41(5):971–993

    Article  Google Scholar 

  • Kole T, Tseng Y, Jiang I, Katz J, Wirtz D (2005) Intracellular mechanics of migrating fibroblasts. Mol Biol Cell 16:328–338

    Article  Google Scholar 

  • Kruse K, Joanny J, Jülicher F, Prost J, Sekimoto K (2005) Generic theory of active polar gels: a paradigm for cytoskeletal dynamics. Eur Phys J E 16:5–16

    Article  Google Scholar 

  • Kuusela E, Alt W (2009) Continuum model of cell adhesion and migration. J Math Biol 58(1):135–161

    Article  MathSciNet  MATH  Google Scholar 

  • Larripa K, Mogilner A (2006) Transport of a 1D viscoelastic actin-myosin strip of gel as a model of a crawling cell. Phys A 372(1):113–123

    Article  Google Scholar 

  • Lee J, Leonard M, Oliver T, Ishihara A, Jacobson K (1994) Traction forces generated by locomoting keratocytes. J Cell Biol 127(6):1957–1964

    Article  Google Scholar 

  • Levayer R, Lecuit T (2011) Biomechanical regulation of contractility: spatial control and dynamics. Trends Cell Biol 22(2):61–81

    Article  Google Scholar 

  • Moeendarbary E, Valon L, Fritzsche M, Harris A, Moulding D, Thrasher A, Stride E, Mahadevan L, Charras G (2013) The cytoplasm of living cells behaves as a poroelastic material. Nat Mater 12(3):253–261

    Article  Google Scholar 

  • Mofrad M (2009) Rheology of the cytoskeleton. Annu Rev Fluid Mech 41:433–453

    Article  Google Scholar 

  • Mogilner A (2009) Mathematics of cell motility: have we got its number? J Math Biol 58(1):105–134

    Article  MathSciNet  MATH  Google Scholar 

  • Mogilner A, Marland E, Bottino D (2001) A minimal model of locomotion applied to the steady gliding movement of fish keratocyte cells. In: Maini P, Othmer H (eds) Mathematical models for biological pattern formation, the IMA volumes in mathematics and its applications, vol 121. Springer, New York, pp 269–293

    Chapter  Google Scholar 

  • Ohsumi T, Flaherty J, Evans M, Barocas V (2008) Three-dimensional simulation of anisotropic cell-driven collagen gel compaction. Biomech Model Mechanobiol 7(1):53–62

    Article  Google Scholar 

  • Oliver J, King J, McKinlay K, Brown P, Grant D, Scotchford C, Wood J (2005) Thin-film theories for two-phase reactive flow models of active cell motion. Math Med Biol 22(1):53

    Article  MATH  Google Scholar 

  • Palecek SP, Loftus JC, Ginsberg MH, Lauffenburger DA, Horwitz AF (1997) Integrin-ligand binding properties govern cell migration speed through cell-substratum adhesiveness. Nature 385:537–540

    Article  Google Scholar 

  • Pollard T, Borisy G (2003) Cellular motility driven by assembly and disassembly of actin filaments. Cell 112(4):453–465

    Article  Google Scholar 

  • Ridley A (2011) Life at the leading edge. Cell 145(7):1012–1022

    Article  Google Scholar 

  • Rottner K, Stradal T (2011) Actin dynamics and turnover in cell motility. Curr Opin Cell Biol 23:569–578

    Article  Google Scholar 

  • Rubinstein B, Jacobson K, Mogilner A (2005) Multiscale two-dimensional modeling of a motile simple-shaped cell. Multiscale Model Simul 3(2):413–439

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Sakamoto Y, Prudhomme S, Zaman M (2011) Viscoelastic gel-strip model for the simulation of migrating cells. Ann Biomed Eng 39(11):2735–2749

    Article  Google Scholar 

  • Sarvestani A, Jabbari E (2009) Analysis of cell locomotion on ligand gradient substrates. Biotechnol Bioeng 103(2):424–429

    Article  Google Scholar 

  • Schaub S, Bohnet S, Laurent V, Meister J, Verkhovsky A (2007) Comparative maps of motion and assembly of filamentous actin and myosin II in migrating cells. Mol Biol Cell 18(10):3723–3732

    Article  Google Scholar 

  • Shao D, Levine H, Rappel W (2012) Coupling actin flow, adhesion, and morphology in a computational cell motility model. PNAS 109(18):6851–6856

    Article  Google Scholar 

  • Svitkina T, Verkhovsky A, McQuade K, Borisy G (1997) Analysis of the actin-myosin II system in fish epidermal keratocytes: mechanism of cell body translocation. J Cell Biol 139(2):397–415

    Article  Google Scholar 

  • Vallotton P, Gupton S, Waterman-Storer C, Danuser G (2004) Simultaneous mapping of filamentous actin flow and turnover in migrating cells by quantitative fluorescent speckle microscopy. PNAS 101(26):9660–9665

    Article  Google Scholar 

  • Vallotton P, Danuser G, Bohnet S, Meister J, Verkhovsky A (2005) Tracking retrograde flow in keratocytes: news from the front. Mol Biol Cell 16(3):1223–1231

    Article  Google Scholar 

  • Verkhovsky A, Svitkina T, Borisy G (1999) Self-polarization and directional motility of cytoplasm. Curr Biol 9(1):11–20

    Article  Google Scholar 

  • Wolgemuth C, Stajic J, Mogilner A (2011) Redundant mechanisms for stable cell locomotion revealed by minimal models. Biophys J 101(3):545–553

    Article  Google Scholar 

  • Wottawah F, Schinkinger S, Lincoln B, Ananthakrishnan R, Romeyke M, Guck J, Käs J (2005) Optical rheology of biological cells. Phys Rev Lett 94(9):98103

    Article  Google Scholar 

  • Wright G, Guy R, Du J, Fogelson A (2011) A high-resolution finite-difference method for simulating two-fluid, viscoelastic gel dynamics. J Non-Newtonian Fluid Mech 166(19):1137–1157

    Article  MATH  Google Scholar 

  • Yamaoka H, Matsushita S, Shimada Y, Adachi T (2012) Multiscale modeling and mechanics of filamentous actin cytoskeleton. Biomech Model Mechanobiol 11:291–302

    Article  Google Scholar 

  • Zajac M, Dacanay B, Mohler W, Wolgemuth C (2008) Depolymerization-driven flow in nematode spermatozoa relates crawling speed to size and shape. Biophys J 94(10):3810–3823

    Article  Google Scholar 

  • Ziebert F, Swaminathan S, Aranson I (2012) Model for self-polarization and motility of keratocyte fragments. J R Soc Interface 9:1084–1092

    Article  Google Scholar 

Download references

Acknowledgments

This publication is based on work supported by Award No. KUK-C1-013-04, made by King Abdullah University of Science and Technology (KAUST). S.L.W. is grateful for funding from the EPSRC in the form of an Advanced Research Fellowship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. M. Oliver.

Appendices

Appendix A: System reduction by potential hodograph transformation

Potential hodograph transformations have proved useful and informative in understanding this type of system (King and Oliver 2005; Kimpton et al. 2012). We perform a similar analysis here, seeking insight into the nature of the viscoelastic governing Eqs. (39)–(41). We introduce an integrated mass variable,

$$\begin{aligned} s&= \int \limits _0^{\chi (s,t)} \theta (\xi ,t)\,\mathrm {d}\xi ,\end{aligned}$$
(128)
$$\begin{aligned} \chi (0,t)&= 0, \end{aligned}$$
(129)

and transform all the dependent variables

$$\begin{aligned} \theta (\chi (s,t),t)&= \Theta (s,t),\end{aligned}$$
(130)
$$\begin{aligned} u(\chi (s,t),t)&= U(s,t),\end{aligned}$$
(131)
$$\begin{aligned} \sigma (\chi (s,t),t)&= \Lambda (s,t), \end{aligned}$$
(132)

and introduce \(\phi \) and \(I\) defined

$$\begin{aligned} \phi (s,t)&= \frac{1}{\Theta (s,t)},\end{aligned}$$
(133)
$$\begin{aligned} I(s,t)&= - \phi (s,t)\int \limits _0^s \phi (\eta ,t)J(1/\phi (\eta ,t))\,\text{ d }\eta , \end{aligned}$$
(134)

to obtain, after some manipulation

$$\begin{aligned} \frac{\partial U}{\partial s}&= \frac{\partial \phi }{\partial t} + \frac{\partial I}{\partial s},\end{aligned}$$
(135)
$$\begin{aligned} U&= A(\phi )\frac{\partial \phi }{\partial s} + B(\phi )\frac{\partial }{\partial s}\left( \frac{\Lambda }{\phi }\right) - C(\phi )V,\end{aligned}$$
(136)
$$\begin{aligned} \frac{1}{D\phi }\frac{\partial U}{\partial s}&= \frac{\partial \Lambda }{\partial t} + \frac{I}{\phi }\frac{\partial \Lambda }{\partial s} + \left( \frac{1}{D}-\frac{2}{\phi }\frac{\partial U}{\partial s} \right) \Lambda , \end{aligned}$$
(137)

where \(A,\,B\) and \(C\) are given by

$$\begin{aligned} A(\phi )&= \frac{\varPsi '(1/\phi )}{\phi ^3\left( \alpha (1/\phi ) + \beta (1/\phi )\right) },\end{aligned}$$
(138)
$$\begin{aligned} B(\phi )&= \frac{1}{\phi \left( \alpha (1/\phi ) + \beta (1/\phi )\right) },\end{aligned}$$
(139)
$$\begin{aligned} C(\phi )&= \frac{\beta (1/\phi )}{\left( \alpha (1/\phi ) + \beta (1/\phi )\right) }. \end{aligned}$$
(140)

It is not clear how (135)–(137) can be combined to eliminate \(U\) and \(\Lambda \), so we simplify the problem by considering the no polymerization/depolymerization case where \(J=0\). This means that \(I=0\) and \(\chi \) becomes a Lagrangian variable. The transformed viscoelastic stress balance becomes

$$\begin{aligned} \frac{\partial \Lambda }{\partial t} + E\Lambda = F, \end{aligned}$$
(141)

where

$$\begin{aligned} E(s,t)&= \frac{1}{D} - \frac{2}{\phi }\frac{\partial \phi }{\partial t},\end{aligned}$$
(142)
$$\begin{aligned} F(s,t)&= \frac{1}{D\phi }\frac{\partial \phi }{\partial t}, \end{aligned}$$
(143)

which has the solution

$$\begin{aligned} \Lambda = \phi ^2e^{-t/D} \left( \frac{\Lambda ^*}{\phi ^{*2}} + \frac{1}{D}\int \limits _0^t \frac{e^{\xi /D}}{\phi ^3}\frac{\partial \phi }{\partial \xi } \,\mathrm {d}\xi \right) , \end{aligned}$$
(144)

where \(\Lambda ^*(x)=\Lambda (x,0)\) is the initial condition for \(\Lambda \) obtained from \(\sigma ^*\) and \(\phi ^*(s)=\phi (s,0)\) is obtained from \(\theta ^*\), the initial condition for the network volume fraction. Now combining (135) and (136), along with \(I=0\), we see that

$$\begin{aligned} \frac{\partial \phi }{\partial t} = \frac{\partial }{\partial s}\left( A(\phi )\frac{\partial \phi }{\partial s} + B(\phi )\frac{\partial }{\partial s}\left( \frac{\Lambda }{\phi }\right) - C(\phi )V\right) . \end{aligned}$$
(145)

Finally we can substitute for \(\Lambda \) using (144) to obtain an integro-partial-differential equation for \(\phi \)

$$\begin{aligned} \frac{\partial \phi }{\partial t} \!=\! \frac{\partial }{\partial s}\left( A(\phi )\frac{\partial \phi }{\partial s} \!+\! B(\phi )\frac{\partial }{\partial s}\left( \phi e^{-t/D} \left( \frac{\Lambda ^*}{\phi ^{*2}} \!+\! \frac{1}{D}\int \limits _0^t \frac{e^{\xi /D}}{\phi ^3}\frac{\partial \phi }{\partial \xi } \,\mathrm {d}\xi \right) \right) \!-\! C(\phi )V\right) ,\nonumber \\ \end{aligned}$$
(146)

which depends on one extra unknown, \(V(t)\), which is determined by a transformed version of the integral constraint (44).

When the network volume fraction is in the contractile regime \(A<0\), conversely \(A>0\) when the network is in the swelling regime. Therefore, the potential hodograph transform is informative in identifying that contraction of the network can be viewed as a destabilising, backwards diffusion term which drives the cell motion. We see explicitly that the viscoelastic network rheology introduces a term which is nonlocal in time. We also see that the timescale over which the network’s flow history effects the current network configuration decreases as \(D\rightarrow 0\).

Appendix B: Convergence graphs

In order to ensure that our code converges as the spatial and temporal discretization is refined we solve the initial value problem (39)–(50) on a sequence of grids. We first fix the time step at \(10^{-3}\) and double the number of space elements successively from 200 to 64,000. The maximum absolute difference between the crawling velocity of the strip, \(V\), calculated on successive grids is plotted against the size of the spatial element \(\varDelta x\) in Fig. 10a. We see that the difference decreases as the spatial grid is refined, so the code converges as the number of spatial elements increases. The spatial convergence of the numerical approximation to \(V\) given by the code is order 3/2. Temporal convergence is similarly demonstrated in Fig. 10b by plotting the difference in crawling velocity of the strip as a function of step size. The temporal convergence of the numerical approximation to \(V\) given by the code was shown to be first order by solving the same problem on a grid with \(800\) space elements for a range of time steps doubling from \(1.25^{-4}\) to \(4\times 10^{-3}\). Full details of the initial conditions and parameters used are given in the caption.

Fig. 10
figure 10

a The rate of spatial convergence for initial conditions \(\theta ^* = \theta _E + 10^{-4}\cos (\pi x),\,\sigma ^* = 10^{-4}\cos (\pi x)\) and the parameter values \(\theta _L = 0.01,\,\theta _E = 0.02,\,\theta _R = 0.03,\,\varPsi ^* = 15{,}000,\,\alpha ^* = 0.4,\,\beta ^* = 3\) and \(D = 0.1\). b The rate of temporal convergence for the viscoelastic code with the same initial conditions and parameters as in a except that here \(\varPsi ^*=12{,}000\)

Appendix C: Myosin driven network contraction

Some parameter choices in the viscoelastic formulation lead to an ill-posed system. A key feature of cell motility that our model currently neglects is the distribution and transport of myosin. Myosin can bind to and move along actin filaments, it can also detach from the network and be transported through the cytosol. A number of modellers have produced models for cell motility that track the distribution and binding of myosin, see e.g. (Rubinstein et al. 2009) who implement an advection–reaction equation for bound myosin density and an advection–reaction–diffusion equation for unbound myosin density. The unbound, free myosin is assumed to be spatially uniform and a reaction–advection equation is used for the bound myosin concentration in Larripa and Mogilner (2006), Mogilner et al. (2001) and Wolgemuth et al. (2011). Alt et al. (2010) present equations to model the transport of bound myosin with the actin network, advection and diffusion of free myosin and binding/unbinding kinetics within the two-phase flow framework. Their calculations are done in the limit where their is no advection of bound myosin and both diffusion of free myosin and binding/unbinding kinetics are sufficiently fast that free myosin is spatially uniform and the binding dynamics are at equilibrium.

Here we consider a simple scenario in which myosin is advected with the cell’s actin network. Our governing equations are given in (39)–(44) along with

$$\begin{aligned} \frac{\partial }{\partial t}(m\theta ) + \frac{\partial }{\partial x}(m\theta u)=0, \end{aligned}$$

where \(m\) is the non-dimensional concentration of bound myosin per unit volume of actin network. An initial condition is required: \(m(x,0)=m^*(x)\), for \(0<x<1\). The aim of this section is not to develop a sophisticated model for myosin transport; that will be the topic of a future paper. Instead we demonstrate that allowing myosin concentration as opposed to network volume fraction drive the network contraction is enough to eliminate the ill-posedness seen in some parameter regimes. We replace (48) with a constitutive law that depends on both the network volume fraction and myosin concentration, a candidate constitutive law taken from Alt et al. (2010) would be

$$\begin{aligned} \varPsi = -S^*ln(1-\theta )-C^*\theta m, \end{aligned}$$
(147)

where \(S^*\) is a nondimensional swelling strength and \(C^*\) is a nondimensional contractile strength, chosen so that \(\varPsi _{\theta }(\theta _E,M)>0\) and \(\varPsi _{m}(\theta _E,M)<0\), where \(M\) is the spatially uniform equilibrium concentration of myosin.

To see that coupling our viscoelastic model to this simple model for myosin transport eliminates the ill-posed behaviour seen in some parameter regimes for our viscoelastic model, we consider a linear stability analysis on an unbounded domain. We consider the perturbations \(\theta =\theta _E+\varepsilon \theta _1+\cdots ,\,\sigma =\varepsilon \sigma _1+\cdots ,\,m=M+\varepsilon m_1+\cdots ,\,u=\varepsilon u_1+\cdots \) and \(V=\varepsilon V_1+\cdots \) and obtain at leading order

$$\begin{aligned} \frac{\partial \theta _1}{\partial t}+\theta _E\frac{\partial u_1}{\partial x}&= -\theta _1,\\ -\varPsi _{\theta E}\frac{\partial \theta _1}{\partial x}-\varPsi _{mE}\frac{\partial m_1}{\partial x}+\theta _E\frac{\partial \sigma _1}{\partial x}&= \alpha _E u_1 + \beta _E(u_1+V_1),\\ D\frac{\partial \sigma _1}{\partial t}+\sigma _1&= \frac{\partial u_1}{\partial x},\\ M\frac{\partial \theta _1}{\partial t}+\theta _E\frac{\partial m_1}{\partial t} + \theta _EM\frac{\partial u_1}{\partial x}&= 0, \end{aligned}$$

where a subscript \(E\) indicates a function is evaluated at the steady state. After some manipulation we can eliminate the variables in favour of \(\theta _1\), which is governed by

$$\begin{aligned}&\left( D\frac{\partial }{\partial t} +1 \right) \left( (\alpha _E+\beta _E)\left( \frac{\partial ^2\theta _1}{\partial t^2}+\frac{\partial \theta _1}{\partial t}\right) - \varPsi _{\theta E}\theta _E\frac{\partial ^3\theta _1}{\partial x^2t} - \varPsi _{mE}M\frac{\partial ^2\theta _1}{\partial x^2} \right) \\&\qquad \qquad \qquad \qquad =\theta _E\frac{\partial ^2}{\partial x^2} \left( \frac{\partial ^2\theta _1}{\partial t^2}+\frac{\partial \theta _1}{\partial t}\right) . \end{aligned}$$

Now we consider a perturbation to the steady state, \(\theta _1\propto \exp (\omega t + ikx)\) and obtain the following cubic equation for the growth rate of a perturbation as a function of the wavenumber of the perturbation

$$\begin{aligned}&D(\alpha _E+\beta _E)\omega ^3 + \left[ (\alpha _E+\beta _E)(1+D)+(1+D\varPsi _{\theta E}\theta _E)k^2\right] \omega ^2\\&\qquad + \left[ \alpha _E+\beta _E+(1+DM\varPsi _{mE}+\varPsi _{\theta E}\theta _E)k^2\right] \omega + M\varPsi _{mE}k^2 =0. \end{aligned}$$

Seeking dominant balances in the large \(k\) limit reveals that one root decays, with

$$\begin{aligned} \omega \sim -\frac{1+D\varPsi _{\theta E}\theta _E}{D(\alpha _E+\beta _E)}k^2, \end{aligned}$$

while the two remaining roots tend to constants, \(\omega _{\infty }\), satisfying

$$\begin{aligned} (1+D\varPsi _{\theta E}\theta _E)\omega _{\infty }^2 + (1+DM\varPsi _{mE} +\varPsi _{\theta E}\theta _E)\omega _{\infty } + M\varPsi _{mE} = 0. \end{aligned}$$

Labelling the coefficients according to \(a\omega ^2+b\omega +c=0\), we note that \(a>0\) and \(c<0\) so the root of the discriminant, \(b^2-4ac\), is bigger than \(b\) in modulus and regardless of the sign on \(b\) there is always one positive real root and one negative real root. Therefore, coupling a simple advection model for myosin to our viscoelastic system allows us to drive network contraction through a slightly different physical mechanism with \(\varPsi _{\theta E}>0,\,\varPsi _{mE}<0\). This analysis demonstrates that on an unbounded domain the resulting system of equations will have a bounded largest growth rate that plateaus to a constant for large wavenumber perturbations, so that the governing equations are always well-posed.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Kimpton, L.S., Whiteley, J.P., Waters, S.L. et al. On a poroviscoelastic model for cell crawling. J. Math. Biol. 70, 133–171 (2015). https://doi.org/10.1007/s00285-014-0755-1

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-014-0755-1

Keywords

Mathematics Subject Classification (2000)

Navigation