Skip to main content
Log in

On the Mechanical Interplay Between Intra- and Inter-Synchronization During Collective Cell Migration: A Numerical Investigation

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

Collective cell migration is a fundamental process that takes place during several biological phenomena such as embryogenesis, immunity response, and tumorogenesis, but the mechanisms that regulate it are still unclear. Similarly to collective animal behavior, cells receive feedbacks in space and time, which control the direction of the migration and the synergy between the cells of the population, respectively. While in single cell migration intra-synchronization (i.e. the synchronization between the protrusion-contraction movement of the cell and the adhesion forces exerted by the cell to move forward) is a sufficient condition for an efficient migration, in collective cell migration the cells must communicate and coordinate their movement between each other in order to be as efficient as possible (i.e. inter-synchronization). Here, we propose a 2D mechanical model of a cell population, which is described as a continuum with embedded discrete cells with or without motility phenotype. The decomposition of the deformation gradient is employed to reproduce the cyclic active strains of each single cell (i.e. protrusion and contraction). We explore different modes of collective migration to investigate the mechanical interplay between intra- and inter-synchronization. The main objective of the paper is to evaluate the efficiency of the cell population in terms of covered distance and how the stress distribution inside the cohort and the single cells may in turn provide insights regarding such efficiency.

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

Similar content being viewed by others

References

  • Allena, R. (2013). Cell migration with multiple pseudopodia: temporal and spatial sensing models. Bull. Math. Biol., 75, 288–316.

    Article  MathSciNet  MATH  Google Scholar 

  • Allena, R., & Aubry, D. (2012). “Run-and-tumble” or “look-and-run”? A mechanical model to explore the behavior of a migrating amoeboid cell. J. Theor. Biol., 306, 15–31.

    Article  MathSciNet  Google Scholar 

  • Allena, R., Mouronval, A.-S., & Aubry, D. (2010). Simulation of multiple morphogenetic movements in the Drosophila embryo by a single 3D finite element model. J. Mech. Behav. Biomed. Mater., 3, 313–323.

    Article  Google Scholar 

  • Anand, R. J., Leaphart, C. L., Mollen, K. P., & Hackam, D. J. (2007). The role of the intestinal barrier in the pathogenesis of necrotizing enterocolitis. Shock, 27, 124–133.

    Article  Google Scholar 

  • Arciero, J. C., Mi, Q., Branca, M. F., Hackam, D. J., & Swigon, D. (2011). Continuum model of collective cell migration in wound healing and colony expansion. Biophys. J., 100, 535–543.

    Article  Google Scholar 

  • Bausch, A. R., Möller, W., & Sackmann, E. (1999). Measurement of local viscoelasticity and forces in living cells by magnetic tweezers. Biophys. J., 76, 573–579.

    Article  Google Scholar 

  • Borisy, G. G., & Svitkina, T. M. (2000). Acting machinery: pushing the envelope. Curr. Opin. Cell Biol., 12, 104–112.

    Article  Google Scholar 

  • Carlier, M. F., & Pantaloni, D. (1997). Control of actin dynamics in cell motility. J. Mol. Biol., 269, 459–467.

    Article  Google Scholar 

  • Carlsson, A. E., & Sept, D. (2008). Mathematical modeling of cell migration. Methods Cell Biol., 84, 911–937.

    Article  Google Scholar 

  • Chen, X., & Friedman, A. (2000). A free boundary problem arising in a model of wound healing. SIAM J. Math. Anal., 32, 778.

    Article  MathSciNet  MATH  Google Scholar 

  • Condeelis, J. (1993). Life at the leading edge: the formation of cell protrusions. Annu. Rev. Cell Biol., 9, 411–444.

    Article  Google Scholar 

  • Dong, C., Slattery, M. J., Rank, B. M., & You, J. (2002). In vitro characterization and micromechanics of tumor cell chemotactic protrusion, locomotion, and extravasation. Ann. Biomed. Eng., 30, 344–355.

    Article  Google Scholar 

  • Drury, J. L., & Dembo, M. (2001). Aspiration of human neutrophils: effects of shear thinning and cortical dissipation. Biophys. J., 81, 3166–3177.

    Article  Google Scholar 

  • Farooqui, R., & Fenteany, G. (2005). Multiple rows of cells behind an epithelial wound edge extend cryptic lamellipodia to collectively drive cell-sheet movement. J. Cell Sci., 118, 51–63.

    Article  Google Scholar 

  • Fenteany, G., Janmey, P. A., & Stossel, T. P. (2000). Signaling pathways and cell mechanics involved in wound closure by epithelial cell sheets. Curr. Biol., 10, 831–838.

    Article  Google Scholar 

  • Flaherty, B., McGarry, J. P., & McHugh, P. E. (2007). Mathematical models of cell motility. Cell Biochem. Biophys., 49, 14–28.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Friedl, P., & Wolf, K. (2010). Plasticity of cell migration: a multiscale tuning model. J. Cell Biol., 188, 11–19.

    Article  Google Scholar 

  • Fukui, Y., Uyeda, T. Q. P., Kitayama, C., & Inoué, S. (2000). How well can an amoeba climb? Proc. Natl. Acad. Sci. USA, 97, 10020–10025.

    Article  Google Scholar 

  • Gaffney, E. A., Maini, P. K., McCaig, C. D., Zhao, M., & Forrester, J. V. (1999). Modelling corneal epithelial wound closure in the presence of physiological electric fields via a moving boundary formalism. IMA J. Math. Appl. Med. Biol., 16, 369–393.

    Article  MATH  Google Scholar 

  • Giannone, G., et al. (2007). Lamellipodial actin mechanically links myosin activity with adhesion-site formation. Cell, 128, 561–575.

    Article  Google Scholar 

  • Glowinski, R., & Pan, T.-W. (1992). Error estimates for fictitious domain/penalty/finite element methods. Calcolo, 29, 125–141.

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  • Graner, F., & Glazier, J. A. (1992). Simulation of biological cell sorting using a two-dimensional extended Potts model. Phys. Rev. Lett., 69, 2013–2016.

    Article  Google Scholar 

  • Holzapfel, G. A. (2000). Nonlinear solid mechanics: a continuum approach for engineering (1st ed.). New York: Wiley.

    Google Scholar 

  • Ilina, O., & Friedl, P. (2009). Mechanisms of collective cell migration at a glance. J. Cell Sci., 122, 3203–3208.

    Article  Google Scholar 

  • Laurent, V. M., et al. (2005). Gradient of rigidity in the lamellipodia of migrating cells revealed by atomic force microscopy. Biophys. J., 89, 667–675.

    Article  Google Scholar 

  • Lubarda, V. (2004). Constitutive theories based on the multiplicative decomposition of deformation gradient: thermoelasticity, elastoplasticity, and biomechanics. Appl. Mech. Rev., 57, 95–109.

    Article  Google Scholar 

  • Maini, P. K., McElwain, D. L. S., & Leavesley, D. I. (2004). Traveling wave model to interpret a wound-healing cell migration assay for human peritoneal mesothelial cells. Tissue Eng., 10, 475–482.

    Article  Google Scholar 

  • McLennan, R., et al. (2012). Multiscale mechanisms of cell migration during development: theory and experiment. Development. Available at: http://dev.biologists.org/content/early/2012/07/04/dev.081471 [Accessed April 27, 2013].

  • Meili, R., Alonso-Latorre, B., del Alamo, J. C., Firtel, R. A., & Lasheras, J. C. (2010). Myosin II is essential for the spatiotemporal organization of traction forces during cell motility. Mol. Biol. Cell, 21, 405–417.

    Article  Google Scholar 

  • Mogilner, A., & Rubinstein, B. (2005). The physics of filopodial protrusion. Biophys. J., 89, 782–795.

    Article  Google Scholar 

  • Murray, J. D. (2003). Mathematical biology II: spatial models and biomedical applications. Berlin: Springer.

    Google Scholar 

  • Phillipson, M., et al. (2006). Intraluminal crawling of neutrophils to emigration sites: a molecularly distinct process from adhesion in the recruitment cascade. J. Exp. Med., 203, 2569–2575.

    Article  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  • Serra-Picamal, X., et al. (2012). Mechanical waves during tissue expansion. Nat. Phys., 8, 628–634.

    Article  Google Scholar 

  • Sheetz, M. P., Felsenfeld, D., Galbraith, C. G., & Choquet, D. (1999). Cell migration as a five-step cycle. Biochem. Soc. Symp., 65, 233–243.

    Google Scholar 

  • Sherratt, J. A., & Murray, J. D. (1990). Models of epidermal wound healing. Proc. - Royal Soc., Biol. Sci., 241, 29–36.

    Article  Google Scholar 

  • Sherratt, J. A., & Murray, J. D. (1991). Mathematical analysis of a basic model for epidermal wound healing. J. Math. Biol., 29, 389–404.

    Article  MATH  Google Scholar 

  • Soofi, S. S., Last, J. A., Liliensiek, S. J., Nealey, P. F., & Murphy, C. J. (2009). The elastic modulus of MatrigelTM as determined by atomic force microscopy. J. Struct. Biol., 167, 216–219.

    Article  Google Scholar 

  • Sumpter, D. J. (2006). The principles of collective animal behaviour. Philos. Trans. R. Soc. Lond. B, Biol. Sci., 361, 5–22.

    Article  Google Scholar 

  • Szabo, B., et al. (2006). Phase transition in the collective migration of tissue cells: experiment and model. arXiv:q-bio/0611045. Available at: http://arxiv.org/abs/q-bio/0611045. Accessed April 27, 2013.

  • Taber, L. A. (2004). Nonlinear theory of elasticity: applications in biomechanics. Singapore: World Scientific.

    Book  Google Scholar 

  • Taber, L. A., Shi, Y., Yang, L., & Bayly, P. V. (2011). A poroelastic model for cell crawling including mechanical coupling between cytoskeletal contraction and actin polymerization. J. Mech. Mater. Struct., 6, 569–589.

    Article  Google Scholar 

  • Tambe, D. T., et al. (2011). Collective cell guidance by cooperative intercellular forces. Nat. Mater., 10, 469–475.

    Article  Google Scholar 

  • Theriot, J. A., & Mitchison, T. J. (1991). Actin microfilament dynamics in locomoting cells. Nature, 352, 126–131. Published online: 11 July 1991. doi:10.1038/352126a0.

    Article  Google Scholar 

  • Trepat, X., et al. (2009). Physical forces during collective cell migration. Nat. Phys., 5, 426–430.

    Article  Google Scholar 

  • Vedel, S., Tay, S., Johnston, D. M., Bruus, H., & Quake, S. R. (2013). Migration of cells in a social context. Proc. Natl. Acad. Sci. USA, 110, 129–134.

    Article  Google Scholar 

  • Vennat, E., Aubry, D., & Degrange, M. (2010). Collagen fiber network infiltration: permeability and capillary infiltration. Transp. Porous Media, 84, 717–733.

    Article  MathSciNet  Google Scholar 

  • Vicsek, T., Czirók, A., Ben-Jacob, E., Cohen, I., & Shochet, O. (1995). Novel type of phase transition in a system of self-driven particles. Phys. Rev. Lett., 75, 1226–1229.

    Article  Google Scholar 

  • Wagh, A. A., et al. (2008). Localized elasticity measured in epithelial cells migrating at a wound edge using atomic force microscopy. Am. J. Physiol., Lung Cell. Mol. Physiol., 295, L54–60.

    Article  Google Scholar 

  • Weijer, C. J. (2009). Collective cell migration in development. J. Cell Sci., 122, 3215–3223.

    Article  Google Scholar 

  • Xue, C., Friedman, A., & Sen, C. K. (2009). A mathematical model of ischemic cutaneous wounds. Proc. Natl. Acad. Sci. USA, 106, 16782–16787.

    Article  Google Scholar 

  • Yamao, M., Naoki, H., & Ishii, S. (2011). Multi-cellular logistics of collective cell migration. PLoS ONE 6, e27950.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. Allena.

Electronic Supplementary Material

Appendix

Appendix

1.1 A.1 Geometry of the cell population

The cells network Ω n is described through a characteristic function h n (p), which reads

$$ h_n (\boldsymbol{p})= \left \{\begin{array}{l@{\quad}l} 1 & \mbox{if}\ \| \boldsymbol{p} - 2 \cdot r_c \cdot \mathit{round}(\boldsymbol{p}_x)\boldsymbol{i}_x- 2\cdot r_c \cdot \mathit{round} (\boldsymbol{p}_y)\boldsymbol{i}_y\|< r_c^2 \\ 0 & \mbox{otherwise} \end{array} \right . $$
(9)

with round being the classical integer function and p(p x ,p y ) the initial position of any particle of the system.

Consequently, the ECM domain Ω ECM is defined by a characteristic function as follows:

$$ h_{\mathrm{ECM}} ( \boldsymbol{p} ) = 1 - h_{n} ( \boldsymbol{p} ) $$
(10)

Each cell inside the population is indicated as c(i,j) where the indices i and j vary as follows:

$$ \left \{ \begin{aligned} &1 \le i \le N_{c} = i_{\max} \\ &1 \le j \le n_{c} ( i ) = n_{c,\max} \sqrt{1 - \biggl( \frac{2i - 1}{N_{c}} \biggr)^{2}} \end{aligned} \right . $$
(11)

with \(N_{c} = \frac{L}{r_{c}}\) and \(n_{c,\max} = \frac{l}{r_{c}}\) being the number of cells along the two axes of the ellipse (Fig. 1c).

The domain \(\varOmega_{c_{i,j}}\) of each cell c(i,j) is defined through a characteristic function as follows:

$$ h_{c_{i,j}} ( \boldsymbol{p} ) = \left \{ \begin{array}{l@{\quad}l} 1 & \mbox{if}\ \Vert \boldsymbol{p} - \boldsymbol{c}_{i,j} \Vert < r_{c}^{2} \\ 0& \mbox{otherwise} \end{array} \right . $$
(12)

Each cell is equipped with a frontal \(\partial \varOmega_{sf_{i,j}}\) and a rear \(\partial \varOmega_{sr_{i,j}}\) adhesion region (Fig. 1d) described by two characteristic functions as

$$\begin{aligned} \begin{aligned} h_{sf_{i,j}} ( \boldsymbol{p} ) &= \left \{ \begin{array}{l@{\quad}l} 1 & \mbox{if}\ ( \boldsymbol{p} - \boldsymbol{c}_{i,j},\boldsymbol{i}_{x} ) > l_{f} \\ 0& \mbox{otherwise} \end{array} \right . \\ h_{sr_{i,j}} ( \boldsymbol{p} ) &= \left \{ \begin{array}{l@{\quad}l} 1 & \mbox{if}\ ( \boldsymbol{p} - \boldsymbol{c}_{i,j},\boldsymbol{i}_{x} ) < l_{r} \\ 0 & \mbox{otherwise} \end{array} \right . \end{aligned} \end{aligned}$$
(13)

where (a,b) defines the scalar product between two vectors, l f and l r are the distances of c i,j from the frontal and rear adhesion surfaces, respectively.

The ellipse is divided into cell rows r(i) (Fig. 1b), which are numbered, similarly to the single cells, from the stern (left) to the bow (right) of the ellipse (1≤iN c =i max) (Fig. 1c) and are defined through a characteristic function as

$$ h_{r_{i}} ( \boldsymbol{p} ) = h_{n} ( \boldsymbol{p} ) \left \{ \begin{array}{l@{\quad}l} 1 & \mbox{if}\ ( \boldsymbol{p}_{x} - \boldsymbol{c}_{i,j_{x}} ) < r_{c} \\ 0 & \mbox{otherwise} \end{array} \right . $$
(14)

1.2 A.2 Constitutive Model of the Cells

As mentioned in Sect. 2.2, the behavior of the active and quiescent cells is described through a generalized viscoelastic 2D Maxwell model (Allena 2013; Allena and Aubry 2012). Since the cells within the cohort may undergo large rotations and deformations during their locomotion, a fully non-linear tensorial approach is required.

For the active cells, the Cauchy stress σ a is assumed to be the sum of the solid (σ a,s ) and the fluid (σ a,f ) Cauchy stresses, while the deformation gradient F a is equal to the solid (F a,s ) and the fluid (F a,f ) deformation gradients.

The decomposition of the deformation gradient (Allena et al. 2010; Lubarda 2004) is used to describe the solid deformation tensor F a,s which is then equal to

$$ \boldsymbol{F}_{a,s} = \boldsymbol{F}_{a,se} \boldsymbol{F}_{a,sa} $$
(15)

where F a,se is the elastic strain tensor responsible for the stress generation and F a,sa is the active strain tensor responsible for the pulsating movement (protrusion-contraction) of each cell. Similarly, the fluid deformation tensor F a,f is the multiplicative decomposition of the fluid-elastic (F a,fe ) and the fluid-viscoelastic (F a,fv ) gradients.

Both the solid σ a,se and the fluid-elastic σ a,fe Cauchy’s stresses are given by isotropic hyperelastic models \(\bar{\sigma}_{a,se}\) and \(\bar{\sigma}_{a,fe}\), respectively, as

$$ \begin{aligned} &\sigma_{a,se} = \bar{\sigma}_{a,se} ( \boldsymbol{e}_{a,se} ) \\ &\sigma_{a,fe} = \bar{\sigma}_{a,fe} ( \boldsymbol{e}_{a,fe} ) \end{aligned} $$
(16)

with e a,se and e a,fe the Euler–Almansi strain tensors for the solid-elastic and the fluid-elastic phases respectively. Additionally, σ a,fe has to be expressed in the actual configuration according to the multiplicative decomposition described above. Finally, the strain rate \(\dot{\boldsymbol{e}}_{a,fv}\) is related to the deviator part of the fluid-viscous stress \(\sigma_{a,fv}^{D}\) as follows:

$$ \dot{\boldsymbol{e}}_{a,fv} = \frac{\sigma_{a,fv}^{D}}{\mu_{a,fv}} $$
(17)

where μ a,fv is the viscosity.

For the quiescent cells, the same equations can be applied but one has to notice that the solid deformation gradient can now be written as

$$ \boldsymbol{F}_{q,s} = \boldsymbol{F}_{q,se} \boldsymbol{F}_{q,sa} = \boldsymbol{F}_{q,se}\boldsymbol{I} $$
(18)

since no active strains take place in these cells (Sect. 2.2).

1.3 A.3 Numerical Implementation of the Constitutive Law

In this section, we provide the general framework of the numerical approach. For further details, we refer the reader to similar works and applications proposed by Glowinski and Pan (1992) and Vennat et al. (2010).

The cell population is modeled as a continuum (the ellipse). Each one of the three internal regions (active cells Ω a,k , quiescent cells Ω q,k , and ECM Ω ECM) is represented by a level-set function (h a,k ,h q,k , and h ECM, respectively). The constitutive behavior of the active and the quiescent cells is described through a 2D generalized Maxwell model, while the ECM is described by a viscoelastic material. In the finite element formulation, the Cauchy stress σ and the viscous strain rate \(\dot{\boldsymbol{e}}_{fv}\) are computed at each point p of the continuum taking into account the contributions of the three regions as follows:

$$ \begin{aligned} &\sigma ( \boldsymbol{p} ) = h_{\mathrm{ECM}} ( \boldsymbol{p} ) \sigma_{\mathrm{ECM}} ( \boldsymbol{p} ) + \sum h_{a,k} ( \boldsymbol{p} )\sigma_{a,k} ( \boldsymbol{p} ) + \sum h_{q,k} ( \boldsymbol{p} )\sigma_{q,k} ( \boldsymbol{p} ) \\ &\dot{\boldsymbol{e}}_{fv} ( \boldsymbol{p} ) = h_{\mathrm{ECM}} ( \boldsymbol{p} )\dot{\boldsymbol{e}}_{\mathrm{ECM}_{fv}} ( \boldsymbol{p} ) + \sum h_{a,k} ( \boldsymbol{p} )\dot{\boldsymbol{e}}_{a,k_{fv}} ( \boldsymbol{p} ) + \sum h_{q,k} ( \boldsymbol{p} )\dot{ \boldsymbol{e}}_{q,k_{fv}} ( \boldsymbol{p} ) \end{aligned} $$
(19)

where \(\dot{\boldsymbol{e}}_{\mathrm{ECM}_{fv}} ( \boldsymbol{p} )\), \(\dot{\boldsymbol{e}}_{a,k_{fv}} ( \boldsymbol{p} )\), and \(\dot{\boldsymbol{e}}_{q,k_{fv}} ( \boldsymbol{p} )\) are the viscous strain rates for the correspondent domains. Thus, the finite element mesh is not adapted to each sub-region of the continuum, but everything is handled via the level set functions, which allow localizing the mechanical behavior.

Then such a constitutive behavior is implemented in the dynamics equation (Eq. (1)), which involves the aforementioned stress, the displacement acceleration a and the adhesion forces f adh. This equation is first transformed into the weak form of the problem (i.e. principle of the virtual works) and it is then discretized by finite elements. Accordingly, cells/ECM or cells/cells mutual forces are automatically equilibrated in a weak sense although in general, the cell boundaries intersect the finite elements edges. In fact, the level set functions describing the sub-regions of the system are defined independently from the finite element mesh.

1.4 A.4 Traveling Wave with Pulse Signal (or Worm-Like Migration)

In this mode of migration, a traveling wave spans the cell population and successively activate and de-activate the cell rows r(i). Thus, only one row is active at the time and the spatial coordinate \(c_{i, j_{x}}\) defining its position changes every migration cycle T (Fig. 2), so that the active cell network h a,2(p,t) is defined as

$$ h_{a,2} ( \boldsymbol{p},t ) = h_{n} ( \boldsymbol{p} ) \cdot \left \{ \begin{array}{l@{\quad}l} 1 & \mbox{if}\ ( \vert \boldsymbol{p}_{x} - c_{i,j_{x}} ( t ) \vert < r_{c} ) \\ 0 & \mbox{otherwise} \end{array} \right . $$
(20)

where \(c_{i, j_{x}}(t)\) reads

$$ c_{i,j_{t}} ( t ) = ( 2L - r_{c} ) - 2r_{c} \cdot \mathit{round} \biggl( \frac{t}{T} - 0.5 \biggr) + 2L \cdot \mathit{round} \biggl( \frac{t}{T \cdot N_{c}} - 0.5 \biggr) $$
(21)

In this case, also the quiescent domain h q,2 varies in space and time as h q,2(p,t)=h n (p)−h a,2(p,t).

Fig. 2
figure 2

Trend of the traveling wave, which defines the spatial coordinate \(c_{i,j_{x}}(t)\) of the temporary active row for the worm-like migration (only 1 h is represented)

Fig. 3
figure 3

Random distribution of the active cells network h a,3(p,t) between 0–6 h (a), 6 h–12 h (b), 12 h–18 h (c) and 18 h–24 h (d) for the tsunami-like migration. In red the active cells (τ i,j =1) and in blue the quiescent cells (τ i,j =0) (Color figure online)

Fig. 4
figure 4

Random values of the cyclic component e a0,ij of the active strain e a0 (a) and of the migration period T ij (b) for the random migration

Fig. 5
figure 5

The distance covered by the cell population after 24 h for the different modes of migration: chemoattractant migration (blue), worm-like migration (red), tsunami-like migration (green), chemoattractant tsunami-like migration (purple), random migration (pink) (Color figure online)

1.5 A.5 Traveling Wave with Random Unit Step Signal (or Tsunami-Like Migration)

The characteristic function h a,3(p,t) defining the active cells network for this mode of migration reads

$$ h_{a,3} ( \boldsymbol{p},t ) = \tau_{i,j} \cdot h_{r_{i}} ( \boldsymbol{p} ) \cdot h_{sw} ( \boldsymbol{p},t ) $$
(22)

where h sw (p,t) describes the progressive wave, which gradually covers the population with a velocity equal to \(\frac{2t}{T}\) and is expressed as

$$ h_{sw} ( \boldsymbol{p},t ) = \left \{ \begin{array}{l@{\quad}l} 1 & \mbox{if}\ ( 2L - r_{c} ) - \boldsymbol{p}_{x} - 2r_{c}\frac{2t}{T} < 0 \\ 0 & \mbox{otherwise} \end{array} \right . $$
(23)

Similarly to the previous case, the quiescent cells domain h q,3 reads h q,3(p,t)=h n (p)−h a,3(p,t).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Allena, R., Aubry, D. & Sharpe, J. On the Mechanical Interplay Between Intra- and Inter-Synchronization During Collective Cell Migration: A Numerical Investigation. Bull Math Biol 75, 2575–2599 (2013). https://doi.org/10.1007/s11538-013-9908-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9908-4

Keywords

Navigation