Skip to main content
Log in

Bridging the gap between individual-based and continuum models of growing cell populations

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

Continuum models for the spatial dynamics of growing cell populations have been widely used to investigate the mechanisms underpinning tissue development and tumour invasion. These models consist of nonlinear partial differential equations that describe the evolution of cellular densities in response to pressure gradients generated by population growth. Little prior work has explored the relation between such continuum models and related single-cell-based models. We present here a simple stochastic individual-based model for the spatial dynamics of multicellular systems whereby cells undergo pressure-driven movement and pressure-dependent proliferation. We show that nonlinear partial differential equations commonly used to model the spatial dynamics of growing cell populations can be formally derived from the branching random walk that underlies our discrete model. Moreover, we carry out a systematic comparison between the individual-based model and its continuum counterparts, both in the case of one single cell population and in the case of multiple cell populations with different biophysical properties. The outcomes of our comparative study demonstrate that the results of computational simulations of the individual-based model faithfully mirror the qualitative and quantitative properties of the solutions to the corresponding nonlinear partial differential equations. Ultimately, these results illustrate how the simple rules governing the dynamics of single cells in our individual-based model can lead to the emergence of complex spatial patterns of population growth observed in continuum models.

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
Fig. 7
Fig. 8

Similar content being viewed by others

References

  • Ambrosi D, Mollica F (2002) On the mechanics of a growing tumor. Int J Eng Sci 40(12):1297–1316

    MathSciNet  MATH  Google Scholar 

  • Ambrosi D, Preziosi L (2002) On the closure of mass balance models for tumor growth. Math Models Methods Appl Sci 12(05):737–754

    MathSciNet  MATH  Google Scholar 

  • Araujo RP, McElwain DS (2004) A history of the study of solid tumour growth: the contribution of mathematical modelling. Bull Math Biol 66(5):1039–1091

    MathSciNet  MATH  Google Scholar 

  • Basan M, Risler T, Joanny JF, Sastre-Garau X, Prost J (2009) Homeostatic competition drives tumor growth and metastasis nucleation. HFSP J 3(4):265–272

    Google Scholar 

  • Binder BJ, Landman KA (2009) Exclusion processes on a growing domain. J Theor Biol 259(3):541–551

    MathSciNet  MATH  Google Scholar 

  • Bresch D, Colin T, Grenier E, Ribba B, Saut O (2010) Computational modeling of solid tumor growth: the avascular stage. SIAM J Sci Comput 32(4):2321–2344

    MathSciNet  MATH  Google Scholar 

  • Brú A, Albertos S, Subiza JL, García-Asenjo JL, Brú I (2003) The universal dynamics of tumor growth. Biophys J 85(5):2948–2961

    Google Scholar 

  • Buttenschoen A, Hillen T, Gerisch A, Painter KJ (2018) A space-jump derivation for non-local models of cell-cell adhesion and non-local chemotaxis. J Math Biol 76(1–2):429–456

    MathSciNet  MATH  Google Scholar 

  • Byrne HM (2010) Dissecting cancer through mathematics: from the cell to the animal model. Nat Rev Cancer 10(3):221

    Google Scholar 

  • Byrne HM, Chaplain MAJ (1995) Growth of nonnecrotic tumors in the presence and absence of inhibitors. Math Biosci 130(2):151–181

    MATH  Google Scholar 

  • Byrne HM, Chaplain MAJ (1996) Growth of necrotic tumors in the presence and absence of inhibitors. Math Biosci 135(2):187–216

    MATH  Google Scholar 

  • Byrne HM, Chaplain MAJ (1997) Free boundary value problems associated with the growth and development of multicellular spheroids. Eur J Appl Math 8(6):639–658

    MathSciNet  MATH  Google Scholar 

  • Byrne HM, Drasdo D (2009) Individual-based and continuum models of growing cell populations: a comparison. J Math Biol 58(4–5):657

    MathSciNet  MATH  Google Scholar 

  • Byrne HM, Preziosi L (2003) Modelling solid tumour growth using the theory of mixtures. Math Med Biol 20(4):341–366

    MATH  Google Scholar 

  • Byrne HM, King JR, McElwain DS, Preziosi L (2003) A two-phase model of solid tumour growth. Appl Math Lett 16(4):567–573

    MathSciNet  MATH  Google Scholar 

  • Champagnat N, Méléard S (2007) Invasion and adaptive evolution for individual-based spatially structured populations. J Math Biol 55(2):147

    MathSciNet  MATH  Google Scholar 

  • Chaplain MAJ, Graziano L, Preziosi L (2006) Mathematical modelling of the loss of tissue compression responsiveness and its role in solid tumour development. Math Med Biol 23(3):197–229

    MATH  Google Scholar 

  • Chen C, Byrne H, King J (2001) The influence of growth-induced stress from the surrounding medium on the development of multicell spheroids. J Math Biol 43(3):191–220

    MathSciNet  MATH  Google Scholar 

  • Ciarletta P, Foret L, Amar MB (2011) The radial growth phase of malignant melanoma: multi-phase modelling, numerical simulations and linear stability analysis. J R Soc Interface 8(56):345–368

    Google Scholar 

  • Dancer EN, Hilhorst D, Mimura M, Peletier LA (1999) Spatial segregation limit of a competition-diffusion system. Eur J Appl Math 10(2):97–115

    MathSciNet  MATH  Google Scholar 

  • Deroulers C, Aubert M, Badoual M, Grammaticos B (2009) Modeling tumor cell migration: from microscopic to macroscopic models. Phys Rev E 79(3):031917

    MathSciNet  Google Scholar 

  • Drasdo D (2005) Coarse graining in simulated cell populations. Adv Complex Syst 8(02–03):319–363

    MathSciNet  MATH  Google Scholar 

  • Drasdo D, Hoehme S (2012) Modeling the impact of granular embedding media, and pulling versus pushing cells on growing cell clones. New J Phys 14(5):055025

    Google Scholar 

  • Dyson L, Maini PK, Baker RE (2012) Macroscopic limits of individual-based models for motile cell populations with volume exclusion. Phys Rev E 86(3):031903

    Google Scholar 

  • Engblom S, Wilson DB, Baker RE (2018) Scalable population-level modelling of biological cells incorporating mechanics and kinetics in continuous time. R Soc Open Sci 5(8):180379

    Google Scholar 

  • Fernando AE, Landman KA, Simpson MJ (2010) Nonlinear diffusion and exclusion processes with contact interactions. Phys Rev E 81(1):011903

    Google Scholar 

  • Greenspan H (1976) On the growth and stability of cell cultures and solid tumors. J Theor Biol 56(1):229–242

    MathSciNet  Google Scholar 

  • Hillen T, Othmer HG (2000) The diffusion limit of transport equations derived from velocity-jump processes. SIAM J Appl Math 61(3):751–775

    MathSciNet  MATH  Google Scholar 

  • Hillen T, Painter KJ (2009) A user’s guide to PDE models for chemotaxis. J Math Biol 58(1–2):183

    MathSciNet  MATH  Google Scholar 

  • Inoue M et al (1991) Derivation of a porous medium equation from many markovian particles and the propagation of chaos. Hiroshima Math J 21(1):85–110

    MathSciNet  MATH  Google Scholar 

  • Johnston ST, Simpson MJ, Baker RE (2012) Mean-field descriptions of collective migration with strong adhesion. Phys Rev E 85(5):051922

    Google Scholar 

  • Johnston ST, Baker RE, McElwain DS, Simpson MJ (2017) Co-operation, competition and crowding: a discrete framework linking Allee kinetics, nonlinear diffusion, shocks and sharp-fronted travelling waves. Sci Rep 7:42134

    Google Scholar 

  • Kim I, Požár N (2018) Porous medium equation to Hele-Shaw flow with general initial density. Trans Am Math Soc 370(2):873–909

    MathSciNet  MATH  Google Scholar 

  • Kim IC, Perthame B, Souganidis PE (2016) Free boundary problems for tumor growth: a viscosity solutions approach. Nonlinear Anal 138:207–228

    MathSciNet  MATH  Google Scholar 

  • Landman KA, Fernando AE (2011) Myopic random walkers and exclusion processes: single and multispecies. Phys A Stat Mech Appl 390(21–22):3742–3753

    Google Scholar 

  • LeVeque RJ (2002) Finite volume methods for hyperbolic problems. Cambridge University Press, Cambridge

  • Lorenzi T, Lorz A, Perthame B (2017) On interfaces between cell populations with different mobilities. Kinet Relat Models 10(1):299–311

    MathSciNet  MATH  Google Scholar 

  • Lowengrub JS, Frieboes HB, Jin F, Chuang YL, Li X, Macklin P, Wise SM, Cristini V (2009) Nonlinear modelling of cancer: bridging the gap between cells and tumours. Nonlinearity 23(1):R1

    MathSciNet  MATH  Google Scholar 

  • Lushnikov PM, Chen N, Alber M (2008) Macroscopic dynamics of biological cells interacting via chemotaxis and direct contact. Phys Rev E 78(6):061904

    Google Scholar 

  • Mellet A, Perthame B, Quiros F (2017) A Hele-Shaw problem for tumor growth. J Funct Anal 273(10):3061–3093

    MathSciNet  MATH  Google Scholar 

  • Mimura M, Sakaguchi H, Matsushita M (2000) Reaction-diffusion modelling of bacterial colony patterns. Phys A Stat Mech Appl 282(1–2):283–303

    Google Scholar 

  • Motsch S, Peurichard D (2018) From short-range repulsion to Hele-Shaw problem in a model of tumor growth. J Math Biol 76(1–2):205–234

    MathSciNet  MATH  Google Scholar 

  • Murray PJ, Edwards CM, Tindall MJ, Maini PK (2009) From a discrete to a continuum model of cell dynamics in one dimension. Phys Rev E 80(3):031912

    Google Scholar 

  • Murray PJ, Edwards CM, Tindall MJ, Maini PK (2012) Classifying general nonlinear force laws in cell-based models via the continuum limit. Phys Rev E 85(2):021921

    Google Scholar 

  • Oelschläger K (1989) On the derivation of reaction-diffusion equations as limit dynamics of systems of moderately interacting stochastic processes. Probab Theory Relat Fields 82(4):565–586

    MathSciNet  MATH  Google Scholar 

  • Oelschläger K (1990) Large systems of interacting particles and the porous medium equation. J Differ Equ 88(2):294–346

    MathSciNet  MATH  Google Scholar 

  • Othmer HG, Hillen T (2002) The diffusion limit of transport equations II: chemotaxis equations. SIAM J Appl Math 62(4):1222–1250

    MathSciNet  MATH  Google Scholar 

  • Othmer HG, Dunbar SR, Alt W (1988) Models of dispersal in biological systems. J Math Biol 26(3):263–298

    MathSciNet  MATH  Google Scholar 

  • Painter KJ, Sherratt JA (2003) Modelling the movement of interacting cell populations. J Theor Biol 225(3):327–339

    MathSciNet  Google Scholar 

  • Penington CJ, Hughes BD, Landman KA (2011) Building macroscale models from microscale probabilistic models: a general probabilistic approach for nonlinear diffusion and multispecies phenomena. Phys Rev E 84(4):041120

    Google Scholar 

  • Penington CJ, Hughes BD, Landman KA (2014) Interacting motile agents: taking a mean-field approach beyond monomers and nearest-neighbor steps. Phys Rev E 89(3):032714

    Google Scholar 

  • Perthame B (2014) Some mathematical aspects of tumor growth and therapy. In: International congress of mathematicians (ICM)

  • Perthame B, Quirós F, Tang M, Vauchelet N (2014) Derivation of a Hele-Shaw type system from a cell model with active motion. Interface Free Bound 16(4):489–508

    MathSciNet  MATH  Google Scholar 

  • Perthame B, Quirós F, Vázquez JL (2014) The Hele-Shaw asymptotics for mechanical models of tumor growth. Arch Ration Mech Anal 212(1):93–127

    MathSciNet  MATH  Google Scholar 

  • Preziosi L (2003) Cancer modelling and simulation. CRC Press, Boca Roton

    MATH  Google Scholar 

  • Ranft J, Basan M, Elgeti J, Joanny JF, Prost J, Jülicher F (2010) Fluidization of tissues by cell division and apoptosis. Proc Nat Acad Sci USA 107(49):20863–20868

    Google Scholar 

  • Roose T, Chapman SJ, Maini PK (2007) Mathematical models of avascular tumor growth. SIAM Rev 49(2):179–208

    MathSciNet  MATH  Google Scholar 

  • Sherratt JA, Chaplain MAJ (2001) A new mathematical model for avascular tumour growth. J Math Biol 43(4):291–312

    MathSciNet  MATH  Google Scholar 

  • Simpson MJ, Merrifield A, Landman KA, Hughes BD (2007) Simulating invasion with cellular automata: connecting cell-scale and population-scale properties. Phys Rev E 76(2):021918

    Google Scholar 

  • Simpson MJ, Landman KA, Hughes BD (2010) Cell invasion with proliferation mechanisms motivated by time-lapse data. Phys A Stat Mech Appl 389(18):3779–3790

    Google Scholar 

  • Stevens A (2000) The derivation of chemotaxis equations as limit dynamics of moderately interacting stochastic many-particle systems. SIAM J Appl Math 61(1):183–212

    MathSciNet  MATH  Google Scholar 

  • Stevens A, Othmer HG (1997) Aggregation, blowup, and collapse: the ABC’s of taxis in reinforced random walks. SIAM J Appl Math 57(4):1044–1081

    MathSciNet  MATH  Google Scholar 

  • Tang M, Vauchelet N, Cheddadi I, Vignon-Clementel I, Drasdo D, Perthame B (2014) Composite waves for a cell population system modeling tumor growth and invasion. In: Partial differential equations: theory, control and approximation. Springer, pp 401–429

  • Van Liedekerke P, Palm M, Jagiella N, Drasdo D (2015) Simulating tissue mechanics with agent-based models: concepts, perspectives and some novel results. Comput Part Mech 2(4):401–444

    Google Scholar 

  • Wang YY, Lehuédé C, Laurent V, Dirat B, Dauvillier S, Bochet L, Le Gonidec S, Escourrou G, Valet P, Muller C (2012) Adipose tissue and breast epithelial cells: a dangerous dynamic duo in breast cancer. Cancer Lett 324(2):142–151

    Google Scholar 

  • Ward JP, King J (1997) Mathematical modelling of avascular-tumour growth. Math Med Biol 14(1):39–69

    MATH  Google Scholar 

  • Ward J, King J (1999) Mathematical modelling of avascular-tumour growth II: modelling growth saturation. Math Med Biol 16(2):171–211

    MATH  Google Scholar 

Download references

Acknowledgements

FRM is funded by the Engineering and Physical Sciences Research Council (EPSRC) (Grant No. EP/N014642/1). TL and FRM gratefully acknowledge Dirk Drasdo and Luís Neves de Almeida for insightful discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fiona R. Macfarlane.

Additional information

Publisher's Note

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

Electronic supplementary material

Appendices

Details of numerical simulations of the individual-based model

We use a uniform discretisation of the interval [0, 100] that consists of 1001 points as the spatial domain (i.e. the grid-step is \(\chi =0.1\)) and we choose the time-step \(\tau =2\times 10^{-3}\). We implement zero-flux boundary conditions by letting the attempted move of a cell be aborted if it requires moving out of the spatial domain. For all simulations, we use the definition (50) of G(p) and we perform numerical computations in Matlab. Further details specific either to the case of one single cell population or to the case of two cell populations are provided in the next sections.

1.1 Setup of numerical simulations for the case of one single cell population

For consistency with Eq. (1), we set \(M=1\) and we drop the index \(h=1\) both from the functions and from the parameters of the individual-based model. We define the rate G in Eqs. (9)–(11) according to Eq. (50). We set the homeostatic pressure \(P=120 \times 10^4\) for the simulation results reported in Figs. 2 and 3, while we choose \(P=120 \times 10^5\) for the simulation results of Fig. 4. Moreover, we choose \(\beta = 4 \times 10^{-6}\) for the simulation results reported in Fig. 2, \(\beta = 4 \times 10^{-5}\) for the simulation results of Fig. 4, and \(\beta \in \left\{ 1.5 \times 10^{-6}, 4 \times 10^{-6}, 4 \times 10^{-5} \right\} \) for the simulation results of Fig. 3. We set \(\nu =0.02\) in Eqs. (13)–(15) and we define the pressure \(p^k_i\) according to the following barotropic relation

$$\begin{aligned} \varPi (\rho ^k_i) = K_{\gamma } \, (\rho ^k_i)^{\gamma } \quad \text {with} \quad K_{\gamma } = \frac{\gamma +1}{\gamma } \quad \text {and} \quad \gamma >1, \end{aligned}$$

which satisfies conditions (3). We let \(\gamma \in \left\{ 1.2, 1.5, 2 \right\} \) for the simulation results of Fig. 4, while we choose \(\gamma = 1.2\) for the simulation results reported in Figs. 2 and 3. We use the initial cell density

$$\begin{aligned} \rho ^0_{i} = A \, \exp {\left( - b \, x_i^{2}\right) } \quad \text {with} \quad A = 2\times 10^{4} \quad \text{ and } \quad b=4\times 10^{-3}. \end{aligned}$$

Additionally, we impose compact support on the initial condition by ensuring \(\rho _{i}^{0}=0\) for all \(x_{i} \ge 50\).

The results presented in Figs. 2 and 3 correspond to the average over three simulations of our individual-based model, while the results in Fig. 4 correspond to one single simulation when \(\gamma =1.5\) or \(\gamma =2\) and the average over two simulations when \(\gamma =1.2\).

1.2 Setup of numerical simulations for the case of two cell populations: Figs. 5 and 6

For consistency with the system of Eq. (7), we choose \(M=2\), and we set \(G_1 \equiv G\) and \(G_2 \equiv 0\) in Eqs. (9)–(11), where G is defined according to Eq. (50) with the homeostatic pressure \(P=10 \times 10^4\) and the factor \(\beta = 4 \times 10^{-5}\). We set \(\nu _1 = 0.01\) and \(\nu _2 = 0.5\) in Eqs. (13)–(15) for the simulation results reported in Fig. 5, while we consider \(\nu _1 = 0.5\) and \(\nu _2 = 0.01\) for the simulation results of Fig. 6. We define the pressure \(p^k_i\) according to the following simplified barotropic relation

$$\begin{aligned} \varPi (\rho ^k_i) = K \, \rho ^k_i \quad \text {with} \quad K = 2, \end{aligned}$$

which satisfies conditions (3). We make use of the initial cell densities

$$\begin{aligned} \rho ^0_{1 i} = A_{1} \exp {\left( -b_{1} \, x_i^{2}\right) } \quad \text{ and } \quad \rho ^0_{2 i}=\left\{ \begin{array}{ll} 0 , \quad \text {for } x_i \le 13, \\ \\ A_{2} \exp {\left( -b_{2} (x_i-14)^{2}\right) }, \quad \text {for } x_i \in (13,29), \\ \\ 0 , \quad \text {for } x_i \ge 29, \end{array} \right. \end{aligned}$$
(52)

where

$$\begin{aligned} A_{1}=1.25\times 10^{4}, \quad b_{1}=0.06, \quad A_{2}=2.5\times 10^{4} \quad \text{ and } \quad b_{2}=6\times 10^{-3}. \end{aligned}$$

The results presented in Figs. 5 and 6 correspond to one single simulation of our individual-based model.

1.3 Setup of numerical simulations for the case of two populations: Figs. 7 and 8

For consistency with the system of Eq. (7), we choose \(M=2\), and we set \(G_1 \equiv G\) and \(G_2 \equiv 0\) in Eqs. (9)–(11), where G is defined according to Eq. (50) with the homeostatic pressure \(P=10 \times 10^4\) and the factor \(\beta = 4 \times 10^{-5}\). We set \(\nu _1 = 0.01\) and \(\nu _2 = 0.5\) in Eqs. (13)–(15) for the simulation results reported in Fig. 7, while we consider \(\nu _1 = 0.5\) and \(\nu _2 = 0.01\) for the simulation results of Fig. 8. We define the pressure \(p^k_i\) according to the following barotropic relation

$$\begin{aligned} \varPi (\rho ^k_i) = q \, (\rho _i^k- \rho ^*)_+ \quad \text{ where } \quad q=10 \quad \text{ and } \quad \rho ^* = r \, P \; \text { with } \; r=10^{-3}, \end{aligned}$$

which satisfies conditions (51). We make use of the initial cell densities (52) with

$$\begin{aligned} A_{1}=12.5\times 10^{4}, \quad b_{1}=0.06, \quad A_{2}=25\times 10^{4} \quad \text{ and } \quad b_{2}=6\times 10^{-3}. \end{aligned}$$

The results presented in Figs. 7 and 8 correspond to one single simulation of our individual-based model.

Details of numerical simulations of the continuum models

We let \(x \in [0,100]\) and we construct numerical solutions for Eq. (1) and for the system of Eq. (7) complemented with zero Neumann boundary conditions. We use a finite volume method based on a time-splitting between the conservative and nonconservative parts. For the conservative parts, transport terms are approximated through an upwind scheme whereby the cell edge states are calculated by means of a high-order extrapolation procedure (LeVeque 2002), while the forward Euler method is used to approximate the nonconservative parts. We consider a uniform discretisation of the interval [0, 100] that consists of 1001 points and we perform numerical computations in Matlab. For all simulations, we use the definition (50) of G(p). Further details specific either to the case of one cell population or to the case of two cell populations are provided in the next sections.

1.1 Setup of numerical simulations for Eq. (1)

The rate G is defined according to Eq. (50) with the homeostatic pressure \(P=120 \times 10^4\) for the numerical solutions reported in Figs. 2 and 3, while we choose \(P=120 \times 10^5\) for the numerical solutions of Fig. 4. Moreover, we choose \(\beta = 4 \times 10^{-6}\) for the numerical solutions reported in Fig. 2, \(\beta = 4 \times 10^{-5}\) for the numerical solutions of Fig. 4, and \(\beta \in \left\{ 1.5 \times 10^{-6}, 4 \times 10^{-6}, 4 \times 10^{-5} \right\} \) for the numerical solutions reported in Fig. 3. We define the pressure p according to the following barotropic relation

$$\begin{aligned} \varPi (\rho ) = K_{\gamma } \, \rho ^{\gamma } \quad \text {with} \quad K_{\gamma } = \frac{\gamma +1}{\gamma } \quad \text {and} \quad \gamma >1, \end{aligned}$$

which satisfies conditions (3). We let \(\gamma \in \left\{ 1.2, 1.5, 2 \right\} \) for the numerical solutions of Fig. 4, while we choose \(\gamma = 1.2\) for the numerical solutions reported in Figs. 2 and 3. Given the parameter values used for the individual-based model in the case of one single cell population, we choose the mobility \(\mu = 4.166\times 10^{-7}\) for the numerical solutions reported in Figs. 2 and 3, while we set \(\mu = 4.166\times 10^{-8}\) for the numerical solutions of Fig. 4. In this way, both values of \(\mu \) satisfy condition (21) for \(h=1\). We impose the initial condition

$$\begin{aligned} \rho (0,x) = A \, \exp {\left( - b \, x^{2}\right) } \quad \text {with} \quad A = 2\times 10^{4} \quad \text{ and } \quad b=4\times 10^{-3}. \end{aligned}$$

Additionally, we impose compact support on the initial condition by ensuring \(\rho (0,x)=0\) for all \(x \ge 50\).

1.2 Setup of numerical simulations for the system of Eq. (7): Figs. 5 and 6

The rate G is defined according to Eq. (50) with the homeostatic pressure \(P=10 \times 10^4\) and \(\beta = 4 \times 10^{-5}\). Given the parameter values used for the individual-based model in the case of two cell populations, we choose the mobilities \(\mu _1 = 2.5\times 10^{-7}\) and \(\mu _2 = 1.25\times 10^{-5}\) for the numerical solutions reported in Fig. 5, and the mobilities \(\mu _1 = 1.25\times 10^{-5}\) and \(\mu _2 = 2.5\times 10^{-7}\) for the numerical solutions of Fig. 6. This ensures that conditions (21) for \(h=1,2\) are satisfied. We define the pressure p according to the following simplified barotropic relation

$$\begin{aligned} \varPi (\rho ) = K \, \rho \quad \text {with} \quad K = 2, \end{aligned}$$

which satisfies conditions (3). We impose the initial conditions

$$\begin{aligned} \rho ^0_{1}(0,x) = A_{1} \exp {\left( -b_{1} \, x^{2}\right) } \quad \text{ and } \quad \rho ^0_{2}=\left\{ \begin{array}{ll} 0 , \quad \text {for } x \le 13, \\ \\ A_{2} \exp {\left( -b_{2} (x-14)^{2}\right) }, \quad \text {for } x \in (13,29), \\ \\ 0 , \quad \text {for } x \ge 29, \end{array} \right. \end{aligned}$$
(53)

where

$$\begin{aligned} A_{1}=1.25\times 10^{4}, \quad b_{1}=0.06, \quad A_{2}=2.5\times 10^{4} \quad \text{ and } \quad b_{2}=6\times 10^{-3}. \end{aligned}$$

1.3 Setup of numerical simulations for the system of Eq. (7): Figs. 7 and 8

The rate G is defined according to Eq. (50) with the homeostatic pressure \(P=10 \times 10^4\) and \(\beta = 4 \times 10^{-5}\). Given the parameter values used for the individual-based model in the case of two cell populations, we choose the mobilities \(\mu _1 = 2.5\times 10^{-7}\) and \(\mu _2 = 1.25\times 10^{-5}\) for the numerical solutions reported in Fig. 7, and the mobilities \(\mu _1 = 1.25\times 10^{-5}\) and \(\mu _2 = 2.5\times 10^{-7}\) for the numerical solutions of Fig. 8. This ensures that conditions (21) for \(h=1,2\) are satisfied. We define the pressure p according to the following barotropic relation

$$\begin{aligned} \varPi (\rho ) = q \, (\rho - \rho ^*)_+ \quad \text{ where } \quad q=10 \quad \text{ and } \quad \rho ^* = r \, P \; \text { with } \; r=10^{-3}, \end{aligned}$$

which satisfies conditions (51). We impose the initial conditions (53) with

$$\begin{aligned} A_{1}=12.5\times 10^{4}, \quad b_{1}=0.06, \quad A_{2}=25\times 10^{4} \quad \text{ and } \quad b_{2}=6\times 10^{-3}. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chaplain, M.A.J., Lorenzi, T. & Macfarlane, F.R. Bridging the gap between individual-based and continuum models of growing cell populations. J. Math. Biol. 80, 343–371 (2020). https://doi.org/10.1007/s00285-019-01391-y

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-019-01391-y

Keywords

Mathematics Subject Classification

Navigation