
This paper investigates the incompressible limit of a system modelling the growth of two cells population. The model describes the dynamics of cell densities, driven by pressure exclusion and cell proliferation. It has been shown that solutions to this system of partial differential equations have the segregation property, meaning that two population initially segregated remain segregated. This work is devoted to the incompressible limit of such system towards a free boundary Hele Shaw type model for two cell populations.
Citation: Pierre Degond, Sophie Hecht, Nicolas Vauchelet. Incompressible limit of a continuum model of tissue growth for two cell populations[J]. Networks and Heterogeneous Media, 2020, 15(1): 57-85. doi: 10.3934/nhm.2020003
[1] | Pierre Degond, Sophie Hecht, Nicolas Vauchelet . Incompressible limit of a continuum model of tissue growth for two cell populations. Networks and Heterogeneous Media, 2020, 15(1): 57-85. doi: 10.3934/nhm.2020003 |
[2] | M.A.J Chaplain, G. Lolas . Mathematical modelling of cancer invasion of tissue: dynamic heterogeneity. Networks and Heterogeneous Media, 2006, 1(3): 399-439. doi: 10.3934/nhm.2006.1.399 |
[3] | Andrea Tosin . Multiphase modeling and qualitative analysis of the growth of tumor cords. Networks and Heterogeneous Media, 2008, 3(1): 43-83. doi: 10.3934/nhm.2008.3.43 |
[4] | Peter V. Gordon, Cyrill B. Muratov . Self-similarity and long-time behavior of solutions of the diffusion equation with nonlinear absorption and a boundary source. Networks and Heterogeneous Media, 2012, 7(4): 767-780. doi: 10.3934/nhm.2012.7.767 |
[5] | Russell Betteridge, Markus R. Owen, H.M. Byrne, Tomás Alarcón, Philip K. Maini . The impact of cell crowding and active cell movement on vascular tumour growth. Networks and Heterogeneous Media, 2006, 1(4): 515-535. doi: 10.3934/nhm.2006.1.515 |
[6] | Urszula Ledzewicz, Heinz Schättler, Shuo Wang . On the role of tumor heterogeneity for optimal cancer chemotherapy. Networks and Heterogeneous Media, 2019, 14(1): 131-147. doi: 10.3934/nhm.2019007 |
[7] | Michiel Bertsch, Masayasu Mimura, Tohru Wakasa . Modeling contact inhibition of growth: Traveling waves. Networks and Heterogeneous Media, 2013, 8(1): 131-147. doi: 10.3934/nhm.2013.8.131 |
[8] | Nora Aïssiouene, Marie-Odile Bristeau, Edwige Godlewski, Jacques Sainte-Marie . A combined finite volume - finite element scheme for a dispersive shallow water system. Networks and Heterogeneous Media, 2016, 11(1): 1-27. doi: 10.3934/nhm.2016.11.1 |
[9] | Rinaldo M. Colombo, Graziano Guerra . A coupling between a non--linear 1D compressible--incompressible limit and the 1D p--system in the non smooth case. Networks and Heterogeneous Media, 2016, 11(2): 313-330. doi: 10.3934/nhm.2016.11.313 |
[10] | Marco Cicalese, Antonio DeSimone, Caterina Ida Zeppieri . Discrete-to-continuum limits for strain-alignment-coupled systems: Magnetostrictive solids, ferroelectric crystals and nematic elastomers. Networks and Heterogeneous Media, 2009, 4(4): 667-708. doi: 10.3934/nhm.2009.4.667 |
This paper investigates the incompressible limit of a system modelling the growth of two cells population. The model describes the dynamics of cell densities, driven by pressure exclusion and cell proliferation. It has been shown that solutions to this system of partial differential equations have the segregation property, meaning that two population initially segregated remain segregated. This work is devoted to the incompressible limit of such system towards a free boundary Hele Shaw type model for two cell populations.
Diversity is key in biology. It appears at all kind of level from the human scale to the microscopic scale, with million of cells types; each scales impacting on the others. During development, the coexistence of different cells types following different rules impact on the growth of tissue and then on the global structures. In a more specific case, this can be observed in cancerous tissue with the invasion of tumour cells in an healthy tissue creating abnormal growth. Furthermore, not all cancerous cells play the same role. They can be proliferative or quiescent depending on their positions, ages, ... To study the influence of these diverse cells on each others from a theoretical view, we introduce mathematical model for multiple populations. In this paper we are interesting in the global dynamics and interactions of the two populations, meaning that we focus specifically on continuous models.
In the already existing literature on macroscopic model, we distinguish two categories. The most common ones involved partial differential equations (PDE) in which cells are represented by densities. These models have been widely used to model growth of tissue [10,31], in particular for tumor growth [1,7,11,16]. Another way to model tissue growth is by considering free boundary models [17,18,21]. In these models the tissue is described by a domain and its growth and movement are driven by the motion of the boundary. The link between these two types of model has been been made via an incompressible limit in [23,24,26,28,29,30]. This link is interesting as both models have their advantages. On the one hand PDE relying models, also called mechanical models, are widely studied with many numerical and analytical tools. On the other hand free boundary models are closer to the biologic vision of the tissue and allow to study motion and dynamics of the tissue. This paper aims to extend the link between the mechanical and the free boundary models, in the case of multiple populations system.
In the specific case of multiple populations, several mathematical models have been already introduced. In particular in population dynamics, the famous Lotka-Volterra system [25] models the dynamics of a predator-prey system. This model has been extended to nonlinear diffusion Lotka-Volterra systems [3,4,5,9]. For the tumor growth modelling (see e.g. [14]), some models focus on mechanical property of tissues such as contact inhibition [2,6,20] and mutation [19]. They have been extended to multiple populations [19,32]. Solutions to these models may have some interesting spatial pattern known as segregation [3,12,27,32].
The two cell populations system under investigation in this paper is an extension on a simplest cell population model proposed in [10,29]. Let
∂tn+∇⋅(nv)=nG(p), on Rd×R+,v=−∇p,p=P(n). | (1) |
In [24,26,28,29,30], the pressure law is given by
∂tn0+Δp0=n0G(p0), in Rd×R+, |
with the relations
p20(Δp0+G(p0))=0, in Rd×R+. |
In the region where
The previous model has the particularity to derive from the free energy
E(n)=∫RP(n(x))dx. |
as a gradient flow for the Wasserstein metric. Using this property a model for two species of cells has been derived in the appendix of [15]. Let us denote
E(nϵ)=∫RP(n1ϵ(x)+n2ϵ(x))dx. |
Restricting to the one dimensional case, the system of equation deriving from this free energy is then defined by,
∂tn1ϵ−∂x(n1ϵ∂xpϵ)=n1ϵG1(pϵ), | (2) |
∂tn2ϵ−∂x(n2ϵ∂xpϵ)=n2ϵG2(pϵ), | (3) |
pϵ=P(nϵ)=ϵnϵ1−nϵ, | (4) |
nϵ=n1ϵ+n2ϵ, | (5) |
with
The model (2)-(5) has been first introduced in [14]. An interesting feature of this model is the preservation of the segregation of the species if they are initially segregated. Recently, in [22], the existence of solutions on
More precisely, we recall the precise statement of the main result in [2]:
Theorem 1.1 (Theorem 1.1 of [2]). Let
∃ζ0∈R such that n1iniϵ=niniϵ1x≤ζ0 and n2iniϵ=niniϵ1x≥ζ0, | (6) |
and
n1iniϵ,n2iniϵ≥0 and 0<A0≤n1iniϵ+n2iniϵ≤B0<1, | (7) |
then there exists
n1ϵ(t,x)=nϵ(t,x)1x≤ζϵ(t) and n2ϵ(t,x)=nϵ(t,x)1x≥ζϵ(t), | (8) |
and
{∂tnϵ−∂x(nϵ∂xpϵ)=nϵG1(pϵ) in the classical sense on{(t,x),x<ζϵ(t)},∂tnϵ−∂x(nϵ∂xpϵ)=nϵG2(pϵ) in the classical sense on {(t,x),x>ζϵ(t)},nϵ(t,ζϵ(t)−)=nϵ(t,ζϵ(t)+),ζ′ϵ(t)=−∂xpϵ(t,ζϵ(t)−)=−∂xpϵ(t,ζϵ(t)+),∂xnϵ(t,±L)=0 for t>0. | (9) |
In [2], the reaction terms are chosen to be affine decreasing. However, as mentioned by the authors, it is easy to verify that their proof can be extended to our system under a set of assumptions for the growth functions which will be defined later in this paper. We also mention that Theorem 1.1 above has been proved in [2] for a smooth pressure law whereas here we have to deal with a possible singularity when the density reaches the value
The aim of this paper is is to study the incompressible limit
When the two species are not in contact, the system is equivalent to the one population model [23], this is why we limit ourself in this paper to the case where the two populations are initially in contact. To use the solutions defined in [2], we restrict the space to a compact domain
We firstly remark that by adding (2) and (3), we get,
∂tnϵ−∂x(nϵ∂xpϵ)=n1ϵG1(pϵ)+n2ϵG2(pϵ) in (−L,L). | (10) |
Multiplying by
∂tpϵ−(p2ϵϵ+pϵ)∂xxpϵ−|∂xpϵ|2=1ϵ(pϵ+ϵ)2(n1ϵG1(pϵ)+n2ϵG2(pϵ)) in (−L,L). | (11) |
Formally, passing at the limit
−p20∂xxp0=p20(n10G1(p0)+n20G2(p0)) in (−L,L). |
In addition, passing formally to the limit
−p20∂xxp0={p20G1(p0) on Ω1(t),p20G2(p0) on Ω2(t). |
Then we obtain a free boundary problem of Hele-Shaw type: On
The outline of the paper is the following. In Section 2 we expose the main results of this paper, which are the convergence of the continuous model (2)-(5) when
In this paper we aim to prove the incompressible limit
{∃Gm>0,‖G1‖∞≤Gm,‖G2‖∞≤Gm,G′1,G′2<0, and ∃P1M,P2M>0,G1(P1M)=0 and G2(P2M)=0,∃γ>0,min(inf[0,P1M]|G′1|,inf[0,P2M]|G′2|)=γ,PM:=max(P1M,P2M),∃gm≥0, min(inf[0,PM]G1,inf[0,PM]G2)≥−gm. | (12) |
The set of assumptions on the growth rate is standard and similar to the one in e.g. [23]. Notice that the boundedness and the decay of
{0≤n1 iniϵ,0≤n2 iniϵ,n iniϵ=n1 iniϵ+n2 iniϵ,0<A0≤n iniϵ≤B0<1,∃ζ0∈(−L,L) such that n1 iniϵ=n iniϵ1x≤ζ0 and n2 iniϵ=n iniϵ1x≥ζ0,p iniϵ:=ϵn iniϵ1−niniϵ≤PM:=max(P1M,P2M),∂xn ini ϵ(±L)=0, |
{max(‖∂xn1 iniϵ‖L1(−L,L),‖∂xn2 iniϵ‖L1(−L,L))≤C and ‖∂xxp iniϵ‖L1(−L,L)≤C,∃n ini1,n ini 2∈L1+(−L,L), such that ‖n1 iniϵ−n ini1‖L1(−L,L)→0 and ‖n2 iniϵ−n ini2‖L1(−L,L)→0, as ϵ→0. | (13) |
These initial conditions imply that
From [2], we recover that at a fix
Remark 1. Considering
∂tniϵ=∂tnϵ(t,x)1x≤ζϵ(t)+nϵζ′ϵ(t)δx=ζϵ(t). |
Given (9), for all
∫R∂tniϵφdx=∫ζϵ(t)−∞∂tnϵφdx+nϵ(t,ζϵ(t))ζ′ϵ(t)φ(ζϵ(t))=∫ζϵ(t)−L∂x(nϵ∂xpϵ)φdx+∫L−LniϵGi(pϵ)φdx+nϵ(t,ζϵ(t))ζ′ϵ(t)φ(ζϵ(t))=−∫ζϵ(t)−Lnϵ∂xpϵ∂xφdx+nϵ(t,ζϵ(t))∂xpϵ(t,ζϵ(t))φ(ζϵ(t))+∫L−LniϵGi(pϵ)φdx−nϵ(t,ζϵ(t))∂xpϵ(t,ζϵ(t))φ(ζϵ(t))=∫L−Lniϵ∂xpϵ∂xφdx+∫L−LniϵGi(pϵ)φdx=∫L−L(∂x(niϵ∂xpϵ)+niϵGi(pϵ))φdx. |
Hence
Considering this particular solution, we are going to show the incompressible limit
Theorem 2.1. Let
0≤n10(t,x)≤1,0≤n20(t,x)≤1, a.e. in QT, | (14) |
0<A0e−gmt≤n0(t,x)≤1,0≤p0≤PM, a.e. in QT, | (15) |
∂tn0−∂xxp0=n10G1(p0)+n20G1(p0), in D′(QT), | (16) |
where
∂tn10−∂x(n10∂xp0)=n10G1(p0), in D′(QT), | (17) |
∂tn20−∂x(n20∂xp0)=n20G2(p0), in D′(QT), | (18) |
complemented with Neumann boundary conditions
(1−n0)p0=0, a.e. in QT, | (19) |
and
n10n20=0, a.e. in QT, | (20) |
and the complementary relation
p20(∂xxp0+n10G1(p0)+n20G2(p0))=0, in D′(QT). | (21) |
Thanks to (19), we may consider the domain
The proof of this convergence result is given in Section 3. It is straightforward to observe that adding (2) and (3) provides an equation on the total density similar to the one found in the one species case [23,29]. Then we use a similar strategy for the proof relying on a compatness method. However the presence of the two populations generate some technical difficulties. To overcome them, we use the segregation property. Notice that this paper is written in the specific case where the two species are separated by one interface, but could be generalised to many interfaces. Using the segregation of the species we are able to obtain a priori estimates on the densities, the pressure and their spatial derivatives. The proof of convergence follows from these new estimates. In order to obtain the complementary relation (21), we follow the approach proposed in [8] which allows us to obtained further regularity.
To complete the results on the asymptotic limit of the model, an uniqueness result for the Hele-Shaw free boundary model for two populations is provided in Proposition 1 in §3.4. The proof of this uniqueness result for the limiting problem is based on Hilbert's duality method.
This section is devoted to the proof of Theorem 2.1, whereas in Section 3.4 the uniqueness of the solution to the Hele Shaw system is established. We first establish some a priori estimates.
The following Lemma establishes the nonnegativity of the densities.
Lemma 3.1. Let
Proof. To show the nonnegativity we use the Stampaccchia method. We multiply (2) by
1n1ϵ<0∂tn1ϵ−1n1ϵ<0∂x(n1ϵ∂xpϵ)=1n1ϵ<0n1ϵG1(pϵ). |
With the above notation, it reads
∂t(n1ϵ)−−∂x((n1ϵ)−∂xpϵ)=(n1ϵ)−G1(pϵ). |
We may integrate in space thanks to the continuity of
ddt∫L−L(n1ϵ)−dx≤∫L−L(n1ϵ)−G1(pϵ)dx≤Gm∫L−L(n1ϵ)−dx. |
Then we integrate in time,
∫L−L(n1ϵ)−dx≤eGmt∫L−L(n1 iniϵ)−dx. |
With the initial condition
Remark 2. We notice that the positivity gives a formal proof of the segregation of any solution of (2)-(5). Indeed, defining
∂trϵ−∂xrϵ∂xpϵ−2rϵ∂xxpϵ=rϵ(G1(pϵ)+G2(pϵ)). |
Multiplying by
ddt∫R−R(rϵ)−dx−∫R−R(rϵ)−∂xxpϵdx≤2Gm∫R−R(rϵ)−. |
Given that
To show the compactness result we establish a priori estimate on the densities, pressure and their derivatives. We first compute the equation on the total density. As shown earlier
∂tnϵ−∂xxH(nϵ)=n1ϵG1(pϵ)+n2ϵG2(pϵ), | (22) |
with
We establish the following a priori estimates
Lemma 3.2. Let us assume that (12) and (13) hold. Let
n1ϵ,n2ϵ in L∞([0,T];L1∩L∞(−L,L));0≤pϵ≤PM,0<A0e−gmt≤nϵ(t)≤PMPM+ϵ≤1. |
Moreover, we have that
Proof. Comparison principle.
The usual comparison principle is not true for this system of equations. However we are able to show some comparison between the total density and
∂t(nϵ−nM)−∂xx(H(nϵ)−H(nM))≤n1ϵG1(P(nϵ))−nM1x≤ζϵ(t)G1(PM)+n2ϵG2(P(nϵ))−nM1x≥ζϵ(t)G2(PM), |
where we use the monotonicity of
Notice that, since the function
∂xxf(y)=f″(y)|∂xy|2+f′(y)∂xxy, |
so for
∂t(nϵ−nM)+−∂xx(H(nϵ)−H(nM))+≤(nϵ−nM)1x≤ζϵ(t)G1(P(nϵ))1nϵ−nM>0+(nϵ−nM)1x≥ζϵ(t)G2(P(nϵ))1nϵ−nM>0+nM(G1(P(nϵ))−G1(P(nM))+G2(P(nϵ))−G2(P(nM)))1nϵ−nM>0. |
Since the function
ddt∫L−L(nϵ−nM)+dx≤ ∂x(H(nϵ)−H(nM))+(L,t)−∂x(H(nϵ)−H(nM))+(−L,t)+∫ζϵ(t)−L(nϵ−nM)1nϵ−nM>0G1(P(nϵ))dx+∫Lζϵ(t)(nϵ−nM)1nϵ−nM>0G2(P(nϵ))dx. |
Given that
ddt∫L−L(nϵ−nM)+dx≤0. |
With the above comparison principle, we conclude that
Estimates from below.
From above, we deduce that the pressure is bounded by
∂tnϵ−∂xxH(nϵ)=n1ϵG1(P(nϵ))+n2ϵG2(P(nϵ))≥−nϵgm. |
Let us introduce
∂t(nm−nϵ)−∂xx(H(nm)−H(nϵ))≤−(nm−nϵ)gm. |
As above, for the comparison principle, we may use the positive part and the Kato inequality to deduce
∂t(nm−nϵ)+−∂xx(H(nm)−H(nϵ))+≤−(nm−nϵ)+gm. |
Integrating in space and in time as above, we deduce that
Integrating (22) on
ddt‖nϵ‖L1(−L,L)≤Gm‖nϵ‖L1(−L,L). |
Integrating in time, we deduce
‖nϵ‖L1(−L,L)≤eGmt‖n iniϵ‖L1(−L,L). |
Since
From the relation (4), we deduce
‖pϵ‖L1(−L,L)≤(ϵ+PM)∫L−L|nϵ|dx≤CeGmt‖n iniϵ‖L1(−L,L). |
Recalling (8), we can refomulate (10) by
∂tnϵ−∂xxH(nϵ)=nϵG(pϵ,t,x) | (23) |
with
∂xG(p,t,x)=(G1(p)−G2(p))δx=ζϵ(t)+G′1(p)∂xp1x≤ζϵ(t)+G′2(p)∂xp1x≥ζϵ(t). |
We derive (23) with respect to
∂t∂xnϵ−∂xx(∂xH(nϵ))= ∂xnϵG(pϵ,t,x)+nϵ(G1(pϵ)−G2(pϵ))δx=ζϵ(t)+nϵ(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))∂xpϵ. |
We multiply by
∂t|∂xnϵ|−∂xx(|∂xH(nϵ)|)≤ |∂xnϵ|G(pϵ,t,x)+nϵ(G1(pϵ)−G2(pϵ))δx=ζϵ(t)sign(∂xinϵ)+nϵ(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))|∂xpϵ|. |
We integrate in space on
∂t∫L−L|∂xnϵ|dx≤ Gm∫L−L|∂xnϵ|dx−γ∫L−Lnϵ|∂xpϵ|dx+nϵ(t,ζϵ(t))|G1(pϵ(t,ζϵ(t))−G2(pϵ(t,ζϵ(t))|. |
Using Gronwall's lemma and the uniform bound on
‖∂xnϵ(t)‖L1(−L,L)+γ∫t0∫L−Lnϵ|∂xpϵ|dxds≤CeGmt(‖∂xn ini ϵ‖L1(−L,L)+1). | (24) |
Given the estimate from below from Lemma 3.2, namely
γA0e−gmt‖∂xpϵ‖L1(QT)≤CeGmt(‖∂xn ini ϵ‖L1(−L,L)+1). | (25) |
Hence we have a uniform bound on
∂xn1ϵ=∂xnϵ1x≤ζϵ(t)+nϵδx=ζϵ(t), |
∂xn2ϵ=∂xnϵ1x≤ζϵ(t)−nϵδx=ζϵ(t). |
So
‖∂xn1ϵ‖L1(−L,L)=∫x≤ζϵ(t)∂xin1ϵdx+nϵ(t,ζϵ(t))≤‖∂xnϵ‖L1(−L,L)+‖n1ϵ‖∞, |
and
‖∂xn2ϵ‖L1(−L,L)=∫x≥ζϵ(t)∂xin2ϵdx−nϵ(t,ζϵ(t))≤‖∂xnϵ‖L1(−L,L)+‖n2ϵ‖∞. |
This concludes the proof.
Lemma 3.3 (
Proof. For a given function
∂tnϵψ(nϵ)−∂x(nϵ∂xpϵ)ψ(nϵ)=(n1ϵG1(pϵ)+n2ϵG2(pϵ))ψ(nϵ). |
Integrating on
ddt∫L−LΨ(nϵ)dx+∫L−Lnϵ∂xnϵ⋅∂xpϵψ′(nϵ)dx=∫L−L(n1ϵG1(pϵ)+n2ϵG2(pϵ))ψ(nϵ)dx, |
where
ddt∫L−Lϵnϵln(nϵ1−nϵ)dx+∫L−L|∂xpϵ|2dx≤Gm∫L−Lϵnϵ|ln(nϵ)−ln(1−nϵ)+11−nϵ|dx. |
After integrating in time and using the expression of the pressure (4), we have
∫L−Lϵnϵln(pϵϵ)dx−∫L−Lϵn ini ϵln(n iniϵ1−n iniϵ)dx+∫T0∫L−L|∂xpϵ|2dxdt≤Gm∫T0∫L−L(ϵnϵ|ln(pϵϵ)|+pϵ)dx. |
Then, to prove that
∫L−Lϵnϵ|ln(pϵϵ)|dx≤∫L−Lϵnϵ|lnpϵ|dx+ϵ|ln(ϵ)|∫L−Lnϵdx≤∫L−L(1−nϵ)pϵ|lnpϵ|dx+ϵ|ln(ϵ)|∫L−Lnϵdx |
Since
∫L−L(1−nϵ)pϵ|ln(pϵ)|dx≤C∫L−L1pϵ>0dx≤2LC. |
This concludes the proof.
Lemma 3.4 (
Proof. Introduce
∂tpϵ=(p2ϵϵ+pϵ)wϵ+|∂xpϵ|2 in (−L,L). | (26) |
As for all
∂twϵ=∂xx(∂tpϵ)+∂t(G(pϵ,t,x))=∂xx((p2ϵϵ+pϵ)wϵ)+2∂x(∂xxpϵ∂xpϵ)+(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))∂tpϵ+ζ′ϵ(t)δx=ζϵ(t)(G1(pϵ)−G2(pϵ)). |
We recall that
∂twϵ=∂xx((p2ϵϵ+pϵ)wϵ)+2∂x(wϵ∂xpϵ)−2∂x(G(pϵ,t,x)∂xpϵ)+(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))((p2ϵϵ+pϵ)wϵ+|∂xpϵ|2)−∂xpϵ(ζϵ(t))δx=ζϵ(t)(G1(pϵ)−G2(pϵ)). | (27) |
Moreover, we have
∂x(G(pϵ,t,x)∂xpϵ)=G(pϵ,t,x)∂xxpϵ+(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))|∂xpϵ|2+(G2(pϵ)−G1(pϵ))∂xpϵδx=ζϵ(t)=G(pϵ,t,x)wϵ−(G(pϵ,t,x))2+(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))|∂xpϵ|2+(G2(pϵ)−G1(pϵ))∂xpϵδx=ζϵ(t). | (28) |
Then inserting (28) in (27), we get
∂twϵ=∂xx((p2ϵϵ+pϵ)wϵ)+2∂x(wϵ∂xpϵ)+(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))(p2ϵϵ+pϵ)wϵ+2(G(pϵ,t,x))2−2G(pϵ,t,x)wϵ+∂xpϵ(ζϵ(t))δx=ζϵ(t)(G1(pϵ)−G2(pϵ))−(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))|∂xpϵ|2. | (29) |
Since
∂t(wϵ)−≤∂xx((p2ϵϵ+pϵ)(wϵ)−)+2∂x((wϵ)−∂xpϵ)+(G′1(pϵ)1x≤ζϵ(t)+G′2(pϵ)1x≥ζϵ(t))(p2ϵϵ+pϵ)(wϵ)−−2(G(pϵ,t,x))21wϵ≤0−2G(pϵ,t,x)(wϵ)−−1wϵ≤0∂xpϵ(ζϵ(t))δx=ζϵ(t)(G1(pϵ)−G2(pϵ)). |
Using assumption (12), we get
∂t(wϵ)−≤∂xx((p2ϵϵ+pϵ)(wϵ)−)+2∂x((wϵ)−∂xpϵ)−γ(p2ϵϵ+pϵ)(wϵ)−−2G(pϵ,t,x)(wϵ)−+2Gm|∂xpϵ(ζϵ(t))|δx=ζϵ(t). | (30) |
We want to integrate on
∫L−L∂xx((p2ϵϵ+pϵ)(wϵ)−)(t,x)dx=∂x((p2ϵϵ+pϵ)(wϵ)−)(t,L)−∂x((p2ϵϵ+pϵ)(wϵ)−)(t,−L). |
Besides, by differentiating the equation of the pressure (26) we have
∂t(∂xpϵ)=∂x((p2ϵϵ+pϵ)wϵ)+2∂xpϵ∂xxpϵ. |
Then, given that the pressure is subject to homogeneous Neumann boundary conditions,
∂x((p2ϵϵ+pϵ)wϵ)(t,±L)=ddt(∂xpϵ(t,±L))−2∂xpϵ(t,±L)∂xxpϵ(t,±L)=0,∀t∈[0,T], |
and so
∂x((p2ϵϵ+pϵ)(wϵ)−)(t,±L)=−∂x((p2ϵϵ+pϵ)wϵ)(t,±L)1wϵ≤0=0,∀t∈[0,T]. |
Therefore, integrating (30) on
ddt∫L−L(wϵ)−dx+γ∫L−L(p2ϵϵ+pϵ)(wϵ)−dx≤−∫L−L2G(pϵ,t,x)(wϵ)−dx+2Gm|∂xpϵ(ζϵ(t))|. |
Note that we may integrate in space thanks to the regularity of the pressure given in Theorem 1.1. Indeed, thanks to the interface conditions at
|∂xpϵ(ζϵ(t))|≤∫ζϵ(t)−L|∂xxpϵ|dx≤∫L−L(|wϵ|+|G(pϵ,t,x)|)dx. |
Moreover, since
∫L−L|wϵ|dx=2∫L−L(wϵ)−dx+∫L−L(∂xxpϵ+G(pϵ))dx≤2∫L−L(wϵ)−dx+2LGm, |
where we use Neumann boundary condition and (12) for the last inequality. Then we get
ddt∫L−L(wϵ)−dx+γ∫L−L(p2ϵϵ+pϵ)(wϵ)−dx≤C(∫L−L(wϵ)−dx+1), | (31) |
with
γ∫T0∫L−L(p2ϵϵ+pϵ)(wϵ)−dxdt≤C(∫T0∫L−L(wϵ)−dxdt+T)≤CT. |
Since
‖∂tpϵ‖L1([0,T]×(−L,L))=∫T0ddt∫L−Lpϵdxdt+2∫T0∫L−L(∂tpϵ)−dxdt≤‖pϵ(T)‖L1(−L,L)+2∫T0∫L−L(p2ϵϵ+pϵ)(wϵ)−dxdt≤‖pϵ(T)‖L1(−L,L)+2CTγ<+∞. |
This concludes the proof.
In the last paragraph we have found a priori estimates for the densities and their space derivatives. To use a compactness argument, we need to obtain estimates on the time derivative. To do so, we are going to use the Aubin Lions theorem [33]. More precisely, we have
Lemma 3.5. Assume that (12) and (13) hold. Let
Proof. According to Lemma 3.3,
W1,1(−L,L)⊂L1(−L,L)⊂W−1,2(−L,L). |
The Aubin Lions theorem implies that
As a consequence, up to extraction of subsequences,
From Lemma 3.5, we have the convergence, up to subsequences, of
Lemma 3.6. Assume that(12) and (13) hold. Let
Proof. We recall that
∂tnϵ−∂xx(pϵ−ϵln(pϵ+ϵ))=n1ϵG1(pϵ)+n2ϵG2(pϵ). |
From the uniform bounds on
ϵlnϵ≤ϵln(pϵ+ϵ)≤ϵln(PM+ϵ). |
Thus, the term in the Laplacian converges strongly to
∂tn0−∂xxp0=n10G1(p0)+n20G2(p0). |
Moreover, let
−∫T0∫L−Ln1ϵ∂tϕdtdx−∫L−Ln1iniϵ(x)ϕ(0,x)dx+∫T0∫L−Ln1ϵ∂xpϵ∂xϕdxdt=∫T0∫L−Ln1ϵG1(pϵ)ϕdxdt. |
Due to the strong convergence of
−∫T0∫L−Ln10∂tϕdtdx−∫L−Lnini1(x)ϕ(0,x)dx+∫T0∫L−Ln10∂xp0∂xϕdxdt=∫T0∫L−Ln10G1(p0)ϕdxdt, |
for any test function
Passing into the limit in the relation
(1−n0)p0=0. |
We can also pass to the limit for the segregation and deduce
To conclude the proof of Theorem 2.1, we are left to establish the relation (21).
In this section we prove the following results.
Lemma 3.7. Assume that (12) and (13) hold. Let
Proof. The approach is based on the previous work [23] and the recent publication [8]. In the weak sense, the complementary relation is equivalent to
∬QT(−2ϕp0|∂xp0|2−p20∂xp0∂xϕ+ϕp20G(p0))dxdt=0,∀ϕ∈D((0,T)×(−L,L)). | (32) |
Multiplying the pressure equation (11) by
ϵ∂tpϵ−pϵ(ϵ+pϵ)∂xxpϵ−ϵ|∂xpϵ|2=pϵ(ϵ+pϵ)G(pϵ), |
where we recall the definition
∬QTp2ϵϕ(∂xxpϵ+G(pϵ))dxdt=ϵ∬QTϕ(∂tpϵ−|∂xpϵ|2−pϵ(∂xxpϵ+G(pϵ))dxdt=ϵ∬QT(ϕ∂tpϵ+pϵ∂xpϵ∂xϕ−ϕpϵG(pϵ))dxdt. |
Therefore, the estimates of Lemmas 3.2 and 3.4 yield
|∬QTp2ϵϕ(∂xxpϵ+G(pϵ))dxdt|≤ϵ(‖ϕ‖L∞‖∂tpϵ‖L1(QT)+‖∂xϕ‖L∞PM‖∂xpϵ‖L1(QT)+‖ϕ‖L∞Gm‖pϵ‖L1(QT))→ϵ→00. |
Moreover, we have
∬QTp2ϵϕ(∂xxpϵ+G(pϵ))dxdt=∬QT(−2ϕpϵ|∂xpϵ|2)dxdt−∬QT(p2ϵ∂xpϵ∂xϕ+p2ϵG(pϵ)ϕ)dxdt=Iϵ+IIϵ. | (33) |
We want to study the convergence of the terms
IIϵ→ϵ→0−∬QT(p20∂xp0∂xϕ+p20G(p0)ϕ)dxdt. |
The convergence of the term
∫L−L|∂xxpϵ|dx=∫L−L∂xxpϵdx+2∫L−L(∂xxpϵ)−dx≤2∫L−L((wϵ)−+Gm)dx≤C. | (34) |
Therefore
∫T−h0∫L−L|∂xpϵ(t+h,x)−∂xpϵ(t,x)|dxdt→h→0+0, | (35) |
uniformly when
‖∂xωη‖L1(R)=∫R|1η2ω′(xη)|dx=1η‖ω′‖L1(R). |
Moreover, for any
‖f−f∗ωη‖L1(−L,L)≤Cη‖f′‖L1(−L,L). | (36) |
Then, we compute,
∫T−h0∫L−L|uh,ϵ(t,x)|dxdt≤∫T−h0∫L−L|uh,ϵ(t,x)−uh,ϵ(t,⋅)∗ωη(x)|dxdt+∫T−h0∫L−L|uh,ϵ(t,⋅)∗ωη(x)|dxdt. |
For the first term of the right hand side, we get with (36),
∫T−h0∫L−L|uh,ϵ(t,x)−uh,ϵ(t,⋅)∗ωη(x)|dxdt≤Cη∫T−h0∫L−L|∂xuh,ϵ|dxdt≤2Cη‖∂xxpϵ‖L1(QT)→η→00, |
uniformly with respect to
∫T−h0∫L−L|uh,ϵ(t,⋅)∗ωη(x)|dxdt=∫T−h0∫L−L|(pϵ(t+h)−pϵ(t))∗∂xωη(x)|dxdt≤‖∂xωη‖L1(R)∫T−h0∫L−L∫t+ht|∂tpϵ(s,x)|dsdxdt≤Chη‖∂tpϵ‖L1(QT), |
which is uniformly bounded with respect to
In this section, we focus on the uniqueness of solutions to the limiting problem (16)–(20). We first observe that from (16) and (20), we have
∂tn0−∂xx(n0p0)=n10G1(p0)+n20G2(p0), in D′(QT). | (37) |
Since we have the segregation property given by (20), we deduce that the support of
∂tn10−∂xx(n10p0)=n10G1(p0), in D′(QT), | (38) |
∂tn20−∂xx(n20p0)=n20G2(p0), in D′(QT). | (39) |
We are going to prove that system (38)–(39) complemented with the segregation property (20) and the relation (19) admits an unique solution. More precisely our result reads:
Proposition 1. Let us assume that assumptions (12) on
Proof. We follow the idea developped in [29] and adapt the Hilbert's duality method. Consider two solutions
∂t(n10−~n10)−∂xx(q1−~q1)=n10G1(p0)−~n10G1(~p0), in D′(QT),∂t(n20−~n20)−∂xx(q2−~q2)=n20G2(p0)−~n20G2(~p0), in D′(QT). |
We first observe that on the set
∬QT[(ni0−~ni0)∂tψi+(qi−~qi)∂xxψi+(ni0Gi(qi)−~ni0Gi(~qi))ψi]dxdt=0. | (40) |
This can be rewritten as, for
∬QT(ni0−~ni0+qi−~qi)(Ai∂tψi+Bi∂xxψi+AiGi(qi)ψi−CiBiψi)dxdt=0, | (41) |
where
Ai=ni0−~ni0ni0−~ni0+qi−~qi,Bi=qi−~qini0−~ni0+qi−~qi,Ci=−~ni0Gi(qi)−Gi(~qi)qi−~qi, |
and we define
The idea of the Hilbert's duality method consists in solving the dual problem, which is defined here by, for any smooth function
{Ai∂tψi+Bi∂xxψi+AiGi(qi)ψi−CiBiψi=AiΦi, in QT,∂ψi(±L)=0 in (0,T),ψi(⋅,T)=0 in (−L,L). | (42) |
If such a system admits a smooth solution, then, by choosing
∬QT(ni0−~ni0+qi−~qi)AiΦidxdt=0. |
From the expression of
∬QT(ni0−~ni0)Φidxdt=0, |
for any smooth function
However, the dual problem (42) is not uniformly parabolic and its coefficients are not smooth. Then, in order to make this step rigorous, a regularization procedure is required. It can be done exactly as in [29,p 109-110]. For the sake of completeness of this paper, this regularizing procedure is recalled in Appendix A.
Lemma 3.8. Under assumptions (12), we have
Proof. We observe that, for
By the same token, we show that, for
Finally, the bound on
The numerical simulations are performed using a finite volume method similar as the one proposed in [13,15]. The scheme used for the conservative part is a classical explicit upwind scheme. To facilitate the reading of this paper, we recall here the scheme used. We divide the computational domain into finite-volume cells
−L=x1/2<x3/2<...<xj−1/2<xj+1/2<...<xMx−1/2<xMx+1/2=L, |
and define the cell average of functions
ˉnβj(t)=1Δx∫Cjnβ(t,x)dx,β∈{1,2}. |
The scheme is obtained by integrating system (2)-(3) over
ˉnk+1βj=−Fkβ,j+1/2−Fkβ,j−1/2Δx+ˉnk+1βjGβ(pkj)for β=1,2, | (43) |
where
Fkβ,j+1/2=(ukβj+1/2)+ˉnkβj+(ukβj+1/2)−ˉnkβj+1,β∈{1,2}, |
where
uβkj+1/2={−pkj+1−pkjΔx,∀j∈{2,...,Mx−1},0, otherwise , |
with the discretized pressure
pkj=ϵnkj1−nkj,nkj=ˉnk1j+ˉnk2j. |
We use the usual notation
In order to illustrate the time dynamics for the model, we plot in Fig 1 the densities computed thanks to the above scheme for
n ini1(x)=0.51[−L;0.25](x) and n ini2(x)=0.51[0.25;L](x), | (44) |
with
G1(p)=10(1−p) and G2(p)=10(1−p/2). | (45) |
We recall that we have defined the parameters
N1Mϵ=p−1(P1M)=P1Mϵ+P1M and N2Mϵ=p−1(P2M)=P2Mϵ+P2M. | (46) |
Since the growth functions are different, clearly
In Fig 1 the red and blue species are initially segregated and equal to 0.5. At first the dynamics is driven by the growth term, so the two species grow and reach their respective maximal packing values
In order to illustrate our main result on the limit
n ini1(x)=1[−L;0.25](x) and n ini2(x)=1[0.25;L](x), |
and the growth function (45). The analytical expressions of the solution to the limiting Hele-Shaw system is computed in [15].
Fig 2 displays the time dynamics of the densities for different values of
We observe in Fig. 2 that the time dynamics of the numerical solutions is similar for each case and follows the dynamics presented above for the case
One interested application of this study is tissue development. Since we consider a system with two populations of cells, we can for example consider the case of tumour with proliferative cells, whose density is denoted
Solution of the limiting Hele-Shaw problem. We assume that initially the tumor is a spheroid centered in 0 and is composed by a spherical core representing the quiescent cells surrounded by a ring representing the proliferative cells. Then, we are looking for particular solution of the limiting Hele-Shaw problem (2)-(3) under the form:
n1(t,x)=1Ω1(t)(x) withΩ1(t)={n1(x,t)=1}=B[−R1(t),R1(t)], |
n2(t,x)=1Ω2(t)(x) withΩ2(t)={n2(x,t)=1}=B(−L,L)∖B[−R1(t),R1(t)]. |
The radius
{R′1(t)=−∂xp(R1(t)),R1(0)=R01, |
where
−∂xxp=n1G1(p)+n2G2(p)inΩ1(t)∪Ω2(t). |
Such functions
∂tn1=R′1(t)(δx=R1(t)−δx=−R1(t)),∂x(n1∂xp)=(δx=R1(t)−δx=−R1(t))∂xp+1[−R1(t),R1(t)]∂xxp. |
Since
∂tn1−∂x(n1∂xp)=1[−R1(t),R1(t)]G1(p)=n1G1(p). |
By applying the same computation on
∂tn2−∂x(n2∂xp)=n2G2(p). |
Analytical solution. As this paper is reduced to the case of dimension 1, we can compute the exact solution of the limiting Hele-Shaw problem (2)-(3) with this initial configuration for some simple expression of the growth terms
G1(p)=g1(P1M−p) and G2(p)=g2(P2M−p). |
This choice means that as the pressure increases, the tumor will grow more slowly, until the pressure reach a critical value (
p(x,t)={(P1M−P2M)√g2sinh(√g2(R1(t)−L))cosh(√g1x)λon Ω1(t),(P1M−P2M)√g1cosh(√g2(x−L))sinh(√g1R1(t))λon Ω2(t). |
with
λ=√g1cosh(√g2(R1−L))sinh(√g1R1)−√g2sinh(√g2(R1−L))cosh(√g1R1), |
Computing the derivatives at the interface
R′1(t)=−√g1g2(P1M−P2M)sinh(√g2(R1(t)−L))cosh(√g1R1(t))λ. | (47) |
We are interested in the study of the evolution of
Numerical simulations. Finally we show some simulations of the mechanical problem for the case of spheroid tumor growth. We run the simulations with
n1=0.51B[−R1(t),R1(t)] and n2=0.51B[−L,L]∖[−R1(t),R1(t)], |
with
R1(0)=0.5 and R2(0)=1.5. |
We fix the parameter
G1(p)=10(1−p) and G2(p)=10(1−p/2), | (48) |
In a second example we display an example where the species
G1(p)=10(4−p) and G2(p)=10(1−p/2). | (49) |
In Fig 3, we display the time dynamics of the densities of these two examples at different time step: (ⅰ)
In this appendix we prove rigorously Proposition 1 using a regularization procedure for the dual problem 42. We follow closely the ideas in [29,p 109-110] which are recall here for the sake of completness of this paper. Since the coefficients
{‖Ai−Aki‖L2(QT)<αik,1k<Aki≤1,‖Bi−Bki‖L2(QT)<βik,1k<Bki≤1,‖Ci−Cki‖L2(QT)<δ1,ik,0≤Cki≤M1,i,‖∂tCki‖L1(QT)≤K1,i,‖Gi(qi)−Gki‖L2(QT)<δ2,ik,|Gki|<M2,i,‖∂xGki‖L2(QT)≤K2,i, |
for some constant
{∂tψki+BkiAki∂xxψki+Gkiψki−CkiBkiAkiψki=Φi, in QT,∂xψki(±L)=0 in (0,T),ψki(⋅,T)=0 in (−L,L). | (50) |
As the coefficients
Using (41) and (50), for
∬QT(ni0−~ni0)Φidxdt=I1,i−I2,i−I3,i+I4,i, |
where
I1,i=∬QT(ni0−~ni0+qi−~qi)BkiAki(Ai−Aki)(Δψki−Ckiψki)dxdt,I2,i=∬QT(ni0−~ni0+qi−~qi)(Bi−Bki)(Δψki−Ckiψki)dxdt,I3,i=∬QT(ni0−~ni0)(Gi(qi)−Gki)ψkidxdt,I4,i=∬QT(ni0−~ni0+qi−~qi)Bi(Ci−Cki)ψkidxdt. |
We intend to show that at the limit
● As
‖ψki‖L∞(QT)≤κ1, |
where
● Multipling (50) by
12‖∂xψki(t)‖2L2(−L,L)+∬Ω×(t,T)BkiAki|∂xxψki−Ckiψki|2dxdt=−∫L−L(Cki(ψki)22)(t)dx+∬(−L,L)×(t,T)(−∂tCki(ψki)22−Gki|∂xψki|2−ψki∂xGki∂xψki+CkiGki(ψki)2+ψki∂xxΦi−ΦiCkiψki)dxdt≤K(1−t+∫Tt‖∂xψki(s)‖2L2(−L,L)ds), | (51) |
with
sup0≤t≤T‖∂xψki‖L2(QT)≤κ2, |
with
● Using (51), we get
‖(BkiAki)1/2(∂xxψki−Ckiψki)‖L2(QT)≤κ3, |
with
We use these bounds to prove the convergence of the integrals
I1,i=˜K∬QTBkiAki|Ai−Aki||∂xxψki−Ckiψki|dxdt≤˜K‖(BkiAki)1/2(Ai−Aki)‖L2(QT)≤˜Kk1/2‖(Ai−Aki)‖L2(QT)≤˜Kαk−1/2I2,i=˜K∬QT|Bi−Bki||∂xxψki−Ckiψki|dxdt≤˜K‖(Akik1/2Bki)1/2(Bi−Bki)‖L2(QT)≤˜Kk1/2‖(Bi−Bki)‖L2(QT)≤˜Kβk−1/2,I3,i=∬QT|ni0−~ni0||G1(q1)−Gki||ψki|dxdt≤˜K‖(Gi(qi)−Gki)‖L2(QT)≤˜Kδ2,in,I4,i=˜K∬QTBi|Ci−Cki||ψki|dxdt≤˜K‖(Ci−Cki)‖L2(QT)≤˜Kn. |
where
limk→+∞∬QT(ni0−~ni0)Φidxdt=0, |
for any smooth function
∬QT[(qi−~qi)∂xxψi+ni0(Gi(qi)−Gi(~qi))ψi]dxdt=0. |
By using
PD acknowledges support by the Engineering and Physical Sciences Research Council (EPSRC) under grants no. EP/M006883/1, EP/N014529/1 and EP/P013651/1, by the Royal Society and the Wolfson Foundation through a Royal Society Wolfson Research Merit Award no. WM130048 and by the National Science Foundation (NSF) under grant no. RNMS11-07444 (KI-Net). PD is on leave from CNRS, Institut de Mathématiques de Toulouse, France. SH acknowledge support from the Francis Crick Institute which receives its core funding from Cancer Research UK (FC001204), the UK Medical Research Council (FC001204), and the Well- come Trust (FC001204). N.V. acknowledges partial support from the ANR blanche project Kibord No ANR-13-BS01-0004 funded by the French Ministry of Research. Part of this work has been done while N.V. was a CNRS fellow at Imperial College, he is really grateful to the CNRS and to Imperial College for the opportunity of this visit. PD, SH and NV would like to thanks Jean-Paul Vincent for stimulating discussion.
No new data was collected in the course of this research.
[1] |
A history of the study of solid tumour growth: The contribution of mathematical modelling,. D.L.S. Bull. Math. Biol. (2004) 66: 1039-1091. ![]() |
[2] |
A free boundary problem arising in a simplified tumour growth model of contact inhibition,. Interfaces Free Bound. (2010) 12: 235-250. ![]() |
[3] |
On interacting populations that disperse to avoid crowding: the effect of a sedentary colony,. Q. Appl. Math. (1984) 19: 1-12. ![]() |
[4] |
On a degenerate diffusion equation of the form c(z)t = ϕ(zx)x with application to population dynamics. J. Differ. Equ. (1987) 67: 56-89. ![]() |
[5] |
On interacting populations that disperse to avoid crowding: The case of equal dispersal velocities,. Nonlinear Anal. Theory Methods Appl. (1987) 11: 493-499. ![]() |
[6] |
A non linear parabolic-hyperbolic system for contact inhibition of cell growth,. Differ. Equ. Appl. (2010) 4: 137-157. ![]() |
[7] |
Computational modeling of solid tumor growth: The avascular stage,. SIAM J. Sci. Comput. (2010) 32: 2321-2344. ![]() |
[8] | F. Bubba, B. Perthame, C. Pouchol and M. Schmidtchen, Hele-shaw limit for a system of two reaction-(cross-)diffusion equations for living tissues, preprint, arXiv: 1901.01692. |
[9] |
Epidemic models with spatial spread due to population migration,. J. Math. Biol. (1983) 16: 181-198. ![]() |
[10] |
Individual-based and continuum models of growing cell populations: A comparison,. J. Math. Biol. (2008) 58: 657-687. ![]() |
[11] |
Growth of necrotic tumors in the presence and absence of inhibitors. Math. Biosci. (1996) 135: 187-216. ![]() |
[12] |
A. J. Carrillo, S. Fagioli, F. Santambrogio and M. Schmidtchen, Splitting schemes and segregation in reaction-(cross-) diffusion systems, SIAM J. Math. Anal., 50 (2018), 5695–5718, arXiv: 1711.05434. doi: 10.1137/17M1158379
![]() |
[13] |
A finite-volume method for nonlinear nonlocal equations with a gradient flow structure,. Commun. Comput. Phys. (2015) 17: 233-258. ![]() |
[14] |
Mathematical modelling of the loss of tissue compression responsiveness and its role in solid tumour development. Math Med Biol. (2006) 23: 197-229. ![]() |
[15] | Incompressible limit of a continuum model of tissue growth with segregation for two cell populations. Math. Biosci. Eng. (2019) 16: 5804-5835. |
[16] |
P. Ciarletta, L. Foret and M. Ben Amar, The radial growth phase of malignant melanoma: Multi-phase modelling, numerical simulations and linear stability analysis, J. R. Soc. Interface, 8 (2011), 345–368, URL http://www.ncbi.nlm.nih.gov/pmc/articles/PMC3030817/. doi: 10.1098/rsif.2010.0285
![]() |
[17] |
Asymptotic behaviour of solutions of a multidimensional moving boundary problem modeling tumor growth. Comm. Partial Differential Equations (2008) 33: 636-655. ![]() |
[18] |
Stability and instability of Liapunov-Schmidt and Hopf bifurcation for a free boundary problem arising in a tumor model. Trans. Am. Math. Soc. (2008) 360: 5291-5342. ![]() |
[19] |
G. Galiano, On a cross-diffusion population model deduced from mutation and splitting of a single species, Comput. Math. Appl., 64 (2012), 1927–1936, URL http://www.sciencedirect.com/science/article/pii/S0898122112002507. doi: 10.1016/j.camwa.2012.03.045
![]() |
[20] |
G. Galiano, S. Shmarev and J. Velasco, Existence and multiplicity of segregated solutions to a cell-growth contact inhibition problem, Discrete Contin. Dyn. Syst., 35 (2015), 1479–1501, URL http://aimsciences.org/journals/displayArticlesnew.jsp?paperID=10564. doi: 10.3934/dcds.2015.35.1479
![]() |
[21] |
Models for the growth of a solid tumor by diffusion. Stud. Appl. Math. (1972) 51: 317-340. ![]() |
[22] |
A two species hyperbolic-parabolic model of tissue growth. Comm. Partial Differential Equations (2019) 44: 1605-1618. ![]() |
[23] |
S. Hecht and N. Vauchelet, Incompressible limit of a mechanical model for tissue growth with non-overlapping constraint, Commun. Math. Sci., 15 (2017), 1913–1932, URL http://www.ncbi.nlm.nih.gov/pmc/articles/PMC5669502/. doi: 10.4310/CMS.2017.v15.n7.a6
![]() |
[24] |
Porous medium equation to Hele-Shaw flow with general initial density. Trans. Amer. Math. Soc. (2018) 370: 873-909. ![]() |
[25] |
Contribution to the theory of periodic reactions,. J. Chem. Biol. Phys. (1909) 14: 271-274. ![]() |
[26] |
A Hele-Shaw problem for tumor growth,. J. Funct. Anal. (2017) 273: 3061-3093. ![]() |
[27] |
Spatial segregation in competitive interaction-diffusion equations. J. Math. Biol. (1980) 9: 49-64. ![]() |
[28] |
Derivation of a Hele-Shaw type system from a cell model with active motion. Interfaces Free Bound. (2014) 16: 489-508. ![]() |
[29] |
The Hele–Shaw asymptotics for mechanical models of tumor growth. Arch. Ration. Mech. Anal. (2014) 212: 93-127. ![]() |
[30] |
B. Perthame and N. Vauchelet, Incompressible limit of a mechanical model of tumour growth with viscosity, Philos. Trans. Roy. Soc. A, Math. Phys. Eng. Sci., 373 (2015), 20140283, 16pp, URL http://www.ncbi.nlm.nih.gov/pmc/articles/PMC4535270/. doi: 10.1098/rsta.2014.0283
![]() |
[31] |
Fluidization of tissues by cell division and apoptosis. Proc. Natl. Acad. Sci. (2010) 107: 20863-20868. ![]() |
[32] |
Spatial segregation of interacting species. J. of Theor. Biol. (1979) 79: 83-99. ![]() |
[33] |
J. Simon, Compact sets in the space Lp(0, T; B), Ann. Mat. Pura Appl. (4), 146 (1987), 65–96. doi: 10.1007/BF01762360
![]() |
1. | Jian-Guo Liu, Xiangsheng Xu, Existence and Incompressible Limit of a Tissue Growth Model with Autophagy, 2021, 53, 0036-1410, 5215, 10.1137/21M1405253 | |
2. | Noemi David, Xinran Ruan, An asymptotic preserving scheme for a tumor growth model of porous medium type, 2022, 56, 2822-7840, 121, 10.1051/m2an/2021080 | |
3. | Tomasz Dębiec, Benoît Perthame, Markus Schmidtchen, Nicolas Vauchelet, Incompressible limit for a two-species model with coupling through Brinkman's law in any dimension, 2021, 145, 00217824, 204, 10.1016/j.matpur.2020.11.002 | |
4. | Noemi David, Benoît Perthame, Free boundary limit of a tumor growth model with nutrient, 2021, 155, 00217824, 62, 10.1016/j.matpur.2021.01.007 | |
5. | Pierre Degond, Sophie Hecht, Michèle Romanos, Ariane Trescases, Multi-species viscous models for tissue growth: incompressible limit and qualitative behaviour, 2022, 85, 0303-6812, 10.1007/s00285-022-01784-6 | |
6. | Tomasz Dębiec, Markus Schmidtchen, Incompressible Limit for a Two-Species Tumour Model with Coupling Through Brinkman’s Law in One Dimension, 2020, 169, 0167-8019, 593, 10.1007/s10440-020-00313-1 | |
7. | Xu'an Dou, Jian-Guo Liu, Zhennan Zhou, A tumor growth model with autophagy: The reaction-(cross-)diffusion system and its free boundary limit, 2023, 28, 1531-3492, 1964, 10.3934/dcdsb.2022154 | |
8. | Inwon Kim, Antoine Mellet, Incompressible Limit of a Porous Media Equation with Bistable and Monostable Reaction Term, 2023, 55, 0036-1410, 5318, 10.1137/22M1517147 | |
9. | Noemi David, Markus Schmidtchen, On the incompressible limit for a tumour growth model incorporating convective effects, 2024, 77, 0010-3640, 2613, 10.1002/cpa.22178 | |
10. | M. I. A. Barbosa, J. Belinha, R. M. Natal Jorge, A. X. Carvalho, Computationally modelling cell proliferation: A review, 2023, 39, 2040-7939, 10.1002/cnm.3715 | |
11. | Inwon Kim, Antoine Mellet, Jeremy Sheung-Him Wu, 2024, Chapter 6, 978-3-031-73422-9, 217, 10.1007/978-3-031-73423-6_6 | |
12. | Charles Elbar, Jakub Skrzeczkowski, On the Inviscid Limit Connecting Brinkman’s and Darcy’s Models of Tissue Growth with Nonlinear Pressure, 2025, 27, 1422-6928, 10.1007/s00021-025-00933-3 |