Skip to main content

Advertisement

Log in

Reaction–Diffusion Finite Element Model of Lateral Line Primordium Migration to Explore Cell Leadership

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

Abstract

Collective cell migration plays a fundamental role in many biological phenomena such as immune response, embryogenesis and tumorigenesis. In the present work, we propose a reaction–diffusion finite element model of the lateral line primordium migration in zebrafish. The population is modelled as a continuum with embedded discrete motile cells, which are assumed to be viscoelastic and able to undergo large deformations. The Wnt/ß-catenin–FGF and cxcr4b–cxcr7b signalling pathways inside the cohort regulating the migration are described through coupled reaction–diffusion equations. The coupling between mechanics and the molecular scenario occurs in two ways. Firstly, the intensity of the protrusion–contraction movement of the cells depends on the cxcr4b concentration. Secondly, the intra-synchronization between the active deformations and the adhesion forces inside each cell is triggered by the cxcr4b–cxcr7b polarity. This influences the inter-synchronization between the cells and results in two main modes of migration: uncoordinated and coordinated. The main objectives of the work were (i) to validate our assumptions with respect to the experimental observations and (ii) to decipher the mechanical conditions leading to efficient migration of the primordium. To achieve the second goal, we will specifically focus on the role of the leader cells and their position inside the population.

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

Similar content being viewed by others

Notes

  1. Wnt/ß-catenin is the canonical Wnt pathway.

  2. FGF stands for fibroblast growth factor.

  3. apc is a protein regulating Wnt/ß-catenin signalling and ensuring association with the microtubules, which is necessary for normal migration, via the C terminus.

  4. SU5402 is a fibroblast growth factor receptor (FGFR)-specific tyrosine kinase inhibitor.

  5. dkk1 is a diffusible inhibitor of the Wnt/ß-catenin pathway.

References

  • Allena R (2013) Cell migration with multiple pseudopodia: temporal and spatial sensing models. Bull Math Biol 75(2):288–316. doi:10.1007/s11538-012-9806-1

    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. doi:10.1016/j.jtbi.2012.03.041

    Article  MathSciNet  Google Scholar 

  • Allena R, Aubry D, Sharpe J (2013) On the mechanical interplay between intra- and inter-synchronization during collective cell migration: a numerical investigation. Bull Math Biol 75(12):2575–2599. doi:10.1007/s11538-013-9908-4

    Article  MathSciNet  MATH  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(4):313–323. doi:10.1016/j.jmbbm.2010.01.001

    Article  Google Scholar 

  • Aman A, Piotrowski T (2008) Wnt/beta-catenin and Fgf signaling control collective cell migration by restricting chemokine receptor expression. Dev cell 15(5):749–761. doi:10.1016/j.devcel.2008.10.002

    Article  Google Scholar 

  • Aman A, Piotrowski T (2009) Multiple signaling interactions coordinate collective cell migration of the posterior lateral line primordium. Cell Adhes Migr 3(4):365–368

    Article  Google Scholar 

  • Aman A, Piotrowski T (2011) Cell–cell signaling interactions coordinate multiple cell behaviors that drive morphogenesis of the lateral line. Cell Adhes Migr 5(6):499–508. doi:10.4161/cam.5.6.19113

    Article  Google Scholar 

  • Anand RJ, Leaphart CL, Mollen KP, Hackam DJ (2007) The role of the intestinal barrier in the pathogenesis of necrotizing enterocolitis. Shock (Augusta, Ga) 27(2):124–133. doi:10.1097/01.shk.0000239774.02904.65

    Article  Google Scholar 

  • Arciero JC, Mi Q, Branca MF, Hackam DJ, Swigon D (2011) Continuum model of collective cell migration in wound healing and colony expansion. Biophys J 100(3):535–543. doi:10.1016/j.bpj.2010.11.083

    Article  Google Scholar 

  • Bausch AR, Möller W, Sackmann E (1999) Measurement of local viscoelasticity and forces in living cells by magnetic tweezers. Biophys J 76(1 Pt 1):573–579

    Article  Google Scholar 

  • Couzin ID, Krause J, Franks NR, Levin SA (2005) Effective leadership and decision-making in animal groups on the move. Nature 433(7025):513–516. doi:10.1038/nature03236

    Article  Google Scholar 

  • Dambly-Chaudière C, Cubedo N, Ghysen A (2007) Control of cell migration in the development of the posterior lateral line: antagonistic interactions between the chemokine receptors CXCR4 and CXCR7/RDC1. BMC Dev Biol 7:23. doi:10.1186/1471-213X-7-23

    Article  Google Scholar 

  • David NB, Sapède D, Saint-Etienne L, Thisse C, Thisse B, Dambly-Chaudière C, Ghysen A (2002) Molecular basis of cell migration in the fish lateral line: role of the chemokine receptor CXCR4 and of its ligand, SDF1. Proc Natl Acad Sci 99(25):16297–16302. doi:10.1073/pnas.252339399

    Article  Google Scholar 

  • Di Costanzo E, Natalini R, Preziosi L (2014) A hybrid mathematical model for self-organizing cell migration in the zebrafish lateral line. J Math Biol. doi:10.1007/s00285-014-0812-9

  • Dong C, Slattery MJ, Rank BM, You J (2002) In vitro characterization and micromechanics of tumor cell chemotactic protrusion, locomotion, and extravasation. Ann Biomed Eng 30(3):344–355

    Article  Google Scholar 

  • Drury JL, Dembo M (2001) Aspiration of human neutrophils: effects of shear thinning and cortical dissipation. Biophys J 81(6):3166–3177

    Article  Google Scholar 

  • Friedl P, Gilmour D (2009) Collective cell migration in morphogenesis, regeneration and cancer. Nat Rev Mol Cell Biol 10(7):445–457. doi:10.1038/nrm2720

    Article  Google Scholar 

  • Fukui Y, Uyeda TQP, Kitayama C, Inoué S (2000) How well can an amoeba climb? Proc Natl Acad Sci 97(18):10020–10025. doi:10.1073/pnas.97.18.10020

    Article  Google Scholar 

  • Gamba L, Cubedo N, Ghysen A, Lutfalla G, Dambly-Chaudière C (2010) Estrogen receptor ESR1 controls cell migration by repressing chemokine receptor CXCR4 in the zebrafish posterior lateral line system. Proc Natl Acad Sci 107(14):6358–6363. doi:10.1073/pnas.0909998107

    Article  Google Scholar 

  • Gompel N, Cubedo N, Thisse C, Thisse B, Dambly-Chaudière C, Ghysen A (2001) Pattern formation in the lateral line of zebrafish. Mech Dev 105(1–2):69–77

    Article  Google Scholar 

  • Graner, Glazier (1992) Simulation of biological cell sorting using a two-dimensional extended Potts model. Phys Rev Lett 69(13):2013–2016

    Article  Google Scholar 

  • Haas P, Gilmour D (2006) Chemokine signaling mediates self-organizing tissue migration in the zebrafish lateral line. Dev Cell 10(5):673–680. doi:10.1016/j.devcel.2006.02.019

    Article  Google Scholar 

  • Holzapfel GA (2000) Nonlinear solid mechanics: a continuum approach for engineering, 1st edn. Wiley, Hoboken

    Google Scholar 

  • Ilina O, Friedl P (2009) Mechanisms of collective cell migration at a glance. J Cell Sci 122(18):3203–3208. doi:10.1242/jcs.036525

    Article  Google Scholar 

  • Kabla AJ (2011) Collective cell migration: leadership, invasion and segregation. arXiv:1108.4286 [physics, q-bio]. Retrieved from http://arxiv.org/abs/1108.4286

  • Laurent VM, Kasas S, Yersin A, Schäffer TE, Catsicas S, Dietler G, Meister J-J et al (2005) Gradient of rigidity in the lamellipodia of migrating cells revealed by atomic force microscopy. Biophys J 89(1):667–675. doi:10.1529/biophysj.104.052316

    Article  Google Scholar 

  • Lecaudey V, Cakan-Akdogan G, Norton WHJ, Gilmour D (2008) Dynamic Fgf signaling couples morphogenesis and migration in the zebrafish lateral line primordium. Development (Cambridge, England) 135(16):2695–2705. doi:10.1242/dev.025981

    Article  Google Scholar 

  • López-Schier H (2010) Fly fishing for collective cell migration. Curr Opin Genet Dev 20(4):428–432. doi:10.1016/j.gde.2010.04.006

    Article  Google Scholar 

  • Lubarda V (2004) Constitutive theories based on the multiplicative decomposition of deformation gradient: thermoelasticity, elastoplasticity, and biomechanics. Appl Mech Rev 57(2):95–109

    Article  Google Scholar 

  • Nechiporuk A, Raible DW (2008) FGF-dependent mechanosensory organ patterning in zebrafish. Science (New York, NY) 320(5884):1774–1777. doi:10.1126/science.1156547

    Article  Google Scholar 

  • Rørth P (2007) Collective guidance of collective cell migration. Trends Cell Biol 17(12):575–579. doi:10.1016/j.tcb.2007.09.007

    Article  Google Scholar 

  • Serra-Picamal X, Conte V, Vincent R, Anon E, Tambe DT, Bazellieres E, Trepat X (2012) Mechanical waves during tissue expansion. Nat Phys 8(8):628–634

    Article  Google Scholar 

  • Sherratt JA, Murray JD (1990) Models of epidermal wound healing. Proc Biol Sci/R Soc 241(1300):29–36. doi:10.1098/rspb.1990.0061

    Article  Google Scholar 

  • Sherratt JA, Murray JD (1991) Mathematical analysis of a basic model for epidermal wound healing. J Math Biol 29(5):389–404

    Article  MATH  Google Scholar 

  • Streichan SJ, Valentin G, Gilmour D, Hufnagel L (2011) Collective cell migration guided by dynamically maintained gradients. Phys Biol 8(4):045004. doi:10.1088/1478-3975/8/4/045004

    Article  Google Scholar 

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

    Book  Google Scholar 

  • Tambe DT, Hardin CC, Angelini TE, Rajendran K, Park CY, Serra-Picamal X, Trepat X (2011) Collective cell guidance by cooperative intercellular forces. Nat Mater 10(6):469–475. doi:10.1038/nmat3025

    Article  Google Scholar 

  • Trepat X, Wasserman MR, Angelini TE, Millet E, Weitz DA, Butler JP, Fredberg JJ (2009) Physical forces during collective cell migration. Nat Phys 5(6):426–430. doi:10.1038/nphys1269

    Article  Google Scholar 

  • Valentin G, Haas P, Gilmour D (2007) The chemokine SDF1a coordinates tissue migration through the spatially restricted activation of Cxcr7 and Cxcr4b. Curr Biol 17(12):1026–1031. doi:10.1016/j.cub.2007.05.020

    Article  Google Scholar 

  • Vedel S, Tay S, Johnston DM, Bruus H, Quake SR (2013) Migration of cells in a social context. Proc Natl Acad Sci 110(1):129–134. doi:10.1073/pnas.1204291110

    Article  Google Scholar 

  • Vitorino P, Meyer T (2008) Modular control of endothelial sheet migration. Genes Dev 22(23):3268–3281. doi:10.1101/gad.1725808

    Article  Google Scholar 

  • Yamao M, Naoki H, Ishii S (2011) Multi-cellular logistics of collective cell migration. PLoS One 6(12):e27950. doi:10.1371/journal.pone.0027950

    Article  Google Scholar 

Download references

Acknowledgments

This work was initiated and partially completed while Rachele Allena was a visitor to the Mathematical Institute, Oxford. The authors are grateful to Dr. Tatjana Piotrowski and Pr. Denis Aubry for useful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to R. Allena.

Electronic supplementary material

Appendix

Appendix

1.1 LLP Geometry

The cell network \(\Omega _n \)is defined by a characteristic function \(h_n \left( {{\varvec{p}}} \right) \) as follows:

$$\begin{aligned} h_n \left( {{\varvec{p}}} \right) =\left\{ {{\begin{array}{ll} 1&{} { \mathrm{if} \left\| {{{\varvec{p}}}-2\cdot r_\mathrm{c} \cdot round\left( {p_x } \right) {{\varvec{i}}}_x -2 \cdot r_\mathrm{c} \cdot round\left( {p_y } \right) {{\varvec{i}}}_y } \right\| <r_c^2 } \\ 0&{} \mathrm{otherwise} \\ \end{array} }} \right. . \end{aligned}$$
(8)

with round being the classical integer function and \({{\varvec{p}}}=\left( {p_x ,p_y } \right) \) the initial position of any particle of the system .

The ECM domain \(\Omega _\mathrm{ECM}\) is identified by the characteristic function \(h_\mathrm{ECM} \left( {{\varvec{p}}} \right) \) which reads

$$\begin{aligned} h_\mathrm{ECM} \left( {{\varvec{p}}} \right) =1-h_n \left( {{\varvec{p}}} \right) . \end{aligned}$$
(9)

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

$$\begin{aligned} \left\{ {\begin{array}{l} 1\le i\le N_\mathrm{c} =i_\mathrm{max } \\ 1\le j\le n_\mathrm{c} \left( i \right) =n_\mathrm{{c,\max }} \sqrt{1-\left( {\frac{2i-1}{N_\mathrm{c} }} \right) ^{2}} \\ \end{array}} \right. \end{aligned}$$
(10)

with \(N_\mathrm{c} =\frac{L}{r_\mathrm{c}}, i_\mathrm{max } =18\) and \(n_\mathrm{{c,\max }} =\frac{l}{r_\mathrm{c} }\) being the number of cells along the two axes of the ellipse (Fig. 2b, c).

The domain \(\Omega _{\mathrm{c}_{i,j} } \) occupied by each cell \(c(i,j)\) is defined through a characteristic function as follows

$$\begin{aligned} h_{\mathrm{c}_{i,j} } \left( {{\varvec{p}}} \right) =\left\{ {\begin{array}{l} 1\quad \mathrm{if}\left\| {{{\varvec{p}}}-{{\varvec{c}}}_{i,j} } \right\| <r_c^2 \\ 0\quad \mathrm{otherwise} \\ \end{array}} \right. \end{aligned}$$
(11)

Each cell is equipped with a frontal \(\partial \Omega _{\mathrm{sf}_{i,j} }\) and a rear \(\partial \Omega _{\mathrm{sr}_{i,j}}\) adhesion region (Fig. 2d) described, respectively, by two characteristic functions

$$\begin{aligned} h_{\mathrm{sf}_{i,j} } \left( {{\varvec{p}}} \right)&= \left\{ {\begin{array}{ll} 1&{} \quad \mathrm{if}\left( {{{\varvec{p}}}-{{\varvec{c}}}_{i,j} ,{{\varvec{i}}}_x } \right) >l_\mathrm{f} \\ 0&{}\quad \mathrm{otherwise} \\ \end{array}} \right. \nonumber \\ h_{\mathrm{sr}_{i,j} } \left( {{\varvec{p}}} \right)&= \left\{ {\begin{array}{ll} 1&{} \quad \mathrm{if}\left( {{{\varvec{p}}}-{{\varvec{c}}}_{i,j} ,{{\varvec{i}}}_x } \right) <-l_r \\ 0&{}\quad \mathrm{otherwise} \\ \end{array}} \right. \end{aligned}$$
(12)

where \(\left( {{{\varvec{a}}},{{\varvec{b}}}} \right) \) defines the scalar product and \(l_\mathrm{f}\) and \(l_\mathrm{r}\) are the distances of \({{\varvec{c}}}_{i,j}\) from the frontal and rear adhesion surfaces, respectively.

The ellipse is divided into cell rows \(r(i)\) (Fig. 2b), which are numbered, similarly to the single cells, from the “stern” (left) to the “bow” (right) of the ellipse (\(1\le i\le N_\mathrm{c} =i_\mathrm{max } \)) (Fig. 2c) and are defined through a characteristic function as

$$\begin{aligned} h_{r_i } \left( {{\varvec{p}}} \right) =h_n \left( {{\varvec{p}}} \right) \left\{ {\begin{array}{ll} 1&{} \quad \mathrm{if}\left( {p_x -c_{i,j_x } } \right) <r_c \\ 0&{} \quad \mathrm{otherwise}. \\ \end{array}} \right. \end{aligned}$$
(13)

1.2 Leading and Trailing Edge of the LLP

The Wnt/ß-catenin–FGF network is mainly based on the spatial polarization of the LLP. We define the leading, \(\Omega _\mathrm{front}\), and the trailing, \(\Omega _\mathrm{rear}\), edges of the LLP through the characteristic functions \(h_\mathrm{front}\) and \(h_\mathrm{rear}\), respectively, as follows:

$$\begin{aligned} h_\mathrm{front} =\left\{ {\begin{array}{ll} 1&{} \quad \mathrm{if}\, p_x >p_{x0} \\ 0&{} \quad \mathrm{otherwise} \\ \end{array}} \right. \nonumber \\ h_\mathrm{rear} =\left\{ {\begin{array}{ll} 1&{} \quad \mathrm{if}\, p_x <p_{x0} \\ 0&{} \quad \mathrm{otherwise} \\ \end{array}} \right. \end{aligned}$$
(14)

where \(p_{x0}\) is the axial coordinate defining the boundary between the leading and the trailing edges.

1.3 Description of Mutants

In the following, we define the reaction–diffusion equations that have been used to describe the molecular and chemokine patterns specific to each mutant as mentioned in Sect. 2.2.

  • apc embryo

    $$\begin{aligned} \frac{\partial \left[ W \right] }{\partial t}&= \underbrace{D_a \nabla ^{2}\left[ W \right] }_\mathrm{diffusion}+\underbrace{S_a \left[ W \right] \left( {1-\left[ W \right] } \right) }_{signalling}-\underbrace{R_\mathrm{a} \left[ W \right] \left[ F \right] }_\mathrm{reaction\, by\, dkk1}\end{aligned}$$
    (15)
    $$\begin{aligned} \frac{\partial \left[ F \right] }{\partial t}&= \underbrace{D_\mathrm{b} \nabla ^{2}\left[ F \right] }_\mathrm{diffusion}+\underbrace{P_\mathrm{b} \left[ W \right] \left( {1-\left[ F \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{b} \left[ F \right] \left[ W \right] }_\mathrm{reaction\, by\, sef}\end{aligned}$$
    (16)
    $$\begin{aligned} \frac{\partial \left[ {c_4 } \right] }{\partial t}&= \underbrace{P_\mathrm{c} \left[ {c_4 } \right] \left( {1-\left[ {c_4 } \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{c} \left[ {c_4 } \right] \left[ F \right] }_\mathrm{reaction\, by\, Fgf}\end{aligned}$$
    (17)
    $$\begin{aligned} \frac{\partial \left[ {c_7 } \right] }{\partial t}&= \underbrace{P_\mathrm{d} \left[ {c_7 } \right] \left( {1-\left[ {c_7 } \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{d} \left[ {c_7 } \right] \left[ W \right] }_\mathrm{reaction\, by\, Wnt} \end{aligned}$$
    (18)
  • SU5402 embryo

    $$\begin{aligned} \frac{\partial \left[ W \right] }{\partial t}&= \underbrace{D_\mathrm{a} \nabla ^{2}\left[ W \right] }_\mathrm{diffusion}+\underbrace{S_\mathrm{a} \left[ W \right] \left( {1-\left[ W \right] } \right) h_\mathrm{front} }_\mathrm{signalling}-\underbrace{R_\mathrm{a} \left[ W \right] \left[ F \right] h_\mathrm{rear} }_\mathrm{reaction\, by\, dkk1}\end{aligned}$$
    (19)
    $$\begin{aligned} \frac{\partial \left[ F \right] }{\partial t}&= \underbrace{D_\mathrm{b} \nabla ^{2}\left[ F \right] }_\mathrm{diffusion}-\underbrace{R_\mathrm{b} \left[ F \right] \left[ W \right] h_\mathrm{front} }_\mathrm{reaction\, by\, sef}\end{aligned}$$
    (20)
    $$\begin{aligned} \frac{\partial \left[ {c_4 } \right] }{\partial t}&= \underbrace{P_\mathrm{c} \left[ {c_4 } \right] \left( {1-\left[ {c_4 } \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{c} \left[ {c_4 } \right] \left[ F \right] }_{\mathrm{reaction}\, {by}\, Fgf}\end{aligned}$$
    (21)
    $$\begin{aligned} \frac{\partial \left[ {c_7 } \right] }{\partial t}&= \underbrace{P_\mathrm{d} \left[ {c_7 } \right] \left( {1-\left[ {c_7 } \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{d} \left[ {c_7 } \right] \left[ W \right] }_{\mathrm{reaction\, by}\, Wnt} \end{aligned}$$
    (22)
  • dkk1 embryo

    $$\begin{aligned} \frac{\partial \left[ W \right] }{\partial t}&= \underbrace{D_\mathrm{a} \nabla ^{2}\left[ W \right] }_\mathrm{diffusion}+\underbrace{S_\mathrm{a} \left[ W \right] \left( {1-\left[ W \right] } \right) h_\mathrm{front} }_\mathrm{signalling}-\underbrace{R_\mathrm{a} \left[ W \right] \left[ F \right] }_{\mathrm{reaction}\, \mathrm{by}\, dkk1}\end{aligned}$$
    (23)
    $$\begin{aligned} \frac{\partial \left[ F \right] }{\partial t}&= \underbrace{D_\mathrm{b} \nabla ^{2}\left[ F \right] }_\mathrm{diffusion}+P_\mathrm{b} \left[ W \right] \left( {1-\left[ F \right] } \right) -\underbrace{R_\mathrm{b} \left[ F \right] \left[ W \right] h_\mathrm{front} }_{\mathrm{reaction\, by}\, sef}\end{aligned}$$
    (24)
    $$\begin{aligned} \frac{\partial \left[ {c_4 } \right] }{\partial t}&= \underbrace{P_\mathrm{c} \left[ {c_4 } \right] \left( {1-\left[ {c_4 } \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{c} \left[ {c_4 } \right] \left[ F \right] }_{\mathrm{reaction\, by}\, Fgf}\end{aligned}$$
    (25)
    $$\begin{aligned} \frac{\partial \left[ {c_7 } \right] }{\partial t}&= \underbrace{P_\mathrm{d} \left[ {c_7 } \right] \left( {1-\left[ {c_7 } \right] } \right) }_\mathrm{production}-\underbrace{R_\mathrm{d} \left[ {c_7 } \right] \left[ W \right] }_{\mathrm{reaction\, by}\, Wnt} \end{aligned}$$
    (26)

1.4 Constitutive Model

As mentioned in Sect. 2.3, the behaviour of the cells is described through a generalized viscoelastic 2D Maxwell model (Allena 2013; Allena and Aubry 2012).

The Cauchy stress, \(\varvec{\sigma }\), is assumed to be the sum of the solid (\(\varvec{\sigma }_\mathrm{s}\)) and the fluid (\(\varvec{\sigma }_\mathrm{f}\)) Cauchy stresses, while the deformation gradient \({{\varvec{F}}}\) is equal to the solid (\({{\varvec{F}}}_\mathrm{s}\)) and the fluid (\({{\varvec{F}}}_\mathrm{f}\)) deformation gradients.

The decomposition of the deformation gradient (Allena et al. 2010; Lubarda 2004) is used to describe the solid deformation tensor, \({{\varvec{F}}}_\mathrm{s}\), which is then given by

$$\begin{aligned} {{\varvec{F}}}_\mathrm{s} ={{\varvec{F}}}_\mathrm{se} {{\varvec{F}}}_\mathrm{sa} \end{aligned}$$
(27)

where \({{\varvec{F}}}_\mathrm{se} \) is the elastic deformation tensor responsible for the stress generation and \({{\varvec{F}}}_\mathrm{sa}\) is the active deformation tensor responsible for the pulsating movement (protrusion–contraction) of each cell. Similarly, the fluid deformation tensor \({{\varvec{F}}}_\mathrm{f}\) is the multiplicative decomposition of the fluid elastic (\({{\varvec{F}}}_\mathrm{fe}\)) and the fluid viscoelastic (\({{\varvec{F}}}_\mathrm{fv}\)) gradients.

Both the solid \(\varvec{\sigma }_\mathrm{se}\) and the fluid elastic \(\varvec{\sigma }_\mathrm{fe}\) Cauchy stresses are given by isotropic hyperelastic models \(\bar{{\varvec{\sigma }}}_\mathrm{se}\) and \(\bar{{\varvec{\sigma }}}_\mathrm{fe}\), respectively, as

$$\begin{aligned} \varvec{\sigma }_\mathrm{se}&= \bar{{\varvec{\sigma }}}_\mathrm{se} \left( {{{\varvec{e}}}_\mathrm{se} } \right) \nonumber \\ \varvec{\sigma }_\mathrm{fe}&= \bar{{\varvec{\sigma }}}_\mathrm{fe} \left( {{{\varvec{e}}}_\mathrm{fe} } \right) \end{aligned}$$
(28)

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

$$\begin{aligned} \dot{{{\varvec{e}}}}_\mathrm{fv} =\frac{\varvec{\sigma }_\mathrm{fv}^D }{{\mu }_\mathrm{fv} } \end{aligned}$$
(29)

where \(\mu _\mathrm{fv} \) is the viscosity and the dot is the derivative with respect to time.

1.5 Coordinated and Uncoordinated Migration

The characteristic functions \(h_\mathrm{c}\) and \(h_\mathrm{uc}\) are expressed as follows

$$\begin{aligned} h_\mathrm{c}&= \left\{ {\begin{array}{ll} 1&{} \mathrm{if} \left\{ {\begin{array}{c} \left( {\left[ {c_4 } \right] >\left[ {c_\mathrm{max } } \right] } \right) h_\mathrm{front} \wedge \left( {\left[ {c_4 } \right] <\left[ {c_\mathrm{min } } \right] } \right) h_\mathrm{rear} \\ \wedge \\ \left( {\left[ {c_7 } \right] >\left[ {c_\mathrm{max } } \right] } \right) h_\mathrm{rear} \wedge \left( {\left[ {c_7 } \right] <\left[ {c_\mathrm{min } } \right] } \right) h_\mathrm{front} \\ \end{array}} \right. \\ 0&{} \mathrm{otherwise} \\ \end{array}} \right. \nonumber \\ h_\mathrm{uc}&= \left\{ {\begin{array}{ll} 1&{} \mathrm{if} \left\{ {\begin{array}{c} \left( {\left[ {c_4 } \right] <\left[ {c_\mathrm{max } } \right] } \right) h_\mathrm{front} \wedge \left( {\left[ {c_4 } \right] >\left[ {c_\mathrm{max } } \right] } \right) h_\mathrm{rear} \\ \wedge \\ \left( {\left[ {c_7 } \right] <\left[ {c_\mathrm{max } } \right] } \right) h_\mathrm{rear} \wedge \left( {\left[ {c_7 } \right] >\left[ {c_\mathrm{min } } \right] } \right) h_\mathrm{front} \\ \end{array}} \right. \\ 0&{} \mathrm{otherwise} \\ \end{array}} \right. \end{aligned}$$
(30)

with \(\wedge \) being the Boolean operator AND and \(c_\mathrm{max}\) and \(c_\mathrm{min}\) being two thresholds fixed here to 0.9 and 0.2, respectively.

The terms \(e_{a,c}\) and \(e_{a,uc}\) describe the cyclic deformation of protrusion–contraction, and they read

$$\begin{aligned} e_{a,c}&= \frac{\left[ {c_4 } \right] }{\alpha _c }\sin \left( {2\pi \frac{t-\frac{T}{2}\left( {i_\mathrm{max } -i} \right) }{T}} \right) h_{r_i } \left( {{\varvec{p}}} \right) h_{wave} \left( {{{\varvec{p}}},t} \right) \nonumber \\ e_{a,uc}&= \frac{\left[ {c_4 } \right] }{\alpha _{uc_{ij} } }\sin \left( {2\pi \frac{t}{T_{uc_{ij} } }} \right) \end{aligned}$$
(31)

where \(t\) is time.

For the coordinated migration, \(\alpha _c\) is set to 2 and \(T\) indicates the duration of a migration period which has been fixed here to 60 s (Allena and Aubry 2012; Dong et al. 2002). Additionally, a wave progressively covers the LLP from the “bow” to the “stern” to activate, one by one, the cell row \(r(i) \) with a velocity equal to \(\frac{2t}{T}\). The wave is expressed by the characteristic function \(h_{wave} \left( {{{\varvec{p}}},t} \right) \) as follows:

$$\begin{aligned} h_{wave} \left( {{{\varvec{p}}},t} \right) =\left\{ {\begin{array}{ll} 1&{} \mathrm{if}\left( {2L-r_c } \right) -p_x -2r_c \frac{2t}{T}<0 \\ 0&{} \mathrm{otherwise}. \\ \end{array}} \right. \end{aligned}$$
(32)

For the uncoordinated migration, \(\alpha _{uc_{ij}}\) and \(T_{uc_{ij}}\) may vary between 0 and 1 and between 60 and 120 s, respectively, for each cell \(c(i,j)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Allena, R., Maini, P.K. Reaction–Diffusion Finite Element Model of Lateral Line Primordium Migration to Explore Cell Leadership. Bull Math Biol 76, 3028–3050 (2014). https://doi.org/10.1007/s11538-014-0043-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-0043-7

Keywords

Navigation