• Keine Ergebnisse gefunden

with chemotaxis and active transport

N/A
N/A
Protected

Academic year: 2022

Aktie "with chemotaxis and active transport"

Copied!
29
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Universit¨ at Regensburg Mathematik

Well-posedness of a Cahn-Hilliard system modelling tumour growth

with chemotaxis and active transport

Harald Garcke and Kei Fong Lam

Preprint Nr. 17/2015

(2)

Well-posedness of a Cahn–Hilliard system modelling tumour growth with chemotaxis and active transport

Harald Garcke Kei Fong Lam November 19, 2015

Abstract

We consider a diffuse interface model for tumour growth consisting of a Cahn–

Hilliard equation with source terms coupled to a reaction-diffusion equation. The coupled system of partial differential equations models a tumour growing in the pres- ence of a nutrient species and surrounded by healthy tissue. The model also takes into account transport mechanisms such as chemotaxis and active transport. We establish well-posedness results for the tumour model and a variant with a quasi-static nutrient.

It will turn out that the presence of the source terms in the Cahn–Hilliard equation leads to new difficulties when one aims to derive a priori estimates. However, we are able to prove continuous dependence on initial and boundary data for the chemical potential and for the order parameter in strong norms.

Key words. Tumour growth; phase field model; Cahn–Hilliard equation; reaction- diffusion equations; chemotaxis; weak solutions; well-posedness.

AMS subject classification. 35K50, 35Q92, 35K57, 92B05.

1 Introduction

Several new diffuse interface models for tumour growth have been introduced recently in [7]. Amongst them is a Cahn–Hilliard equation coupled with a reaction-diffusion equation for a nutrient species. The model equations are given as

tϕ= div(m(ϕ)∇µ) + (λpσ−λa)h(ϕ) in Ω× (0, T), (1.1a) µ=AΨ(ϕ) −B∆ϕ−χϕσ in Ω× (0, T), (1.1b)

tσ= div(n(ϕ)(χσ∇σ−χϕ∇ϕ)) −λcσh(ϕ) in Ω× (0, T), (1.1c) 0= ∇ϕ⋅ν= ∇µ⋅ν on Γ× (0, T), (1.1d) n(ϕ)χσ∇σ⋅ν=K(σ−σ) on Γ× (0, T). (1.1e) Here, Ω⊂Rdis a bounded domain with boundary Γ∶=∂Ω, σ denotes the concentration of an unspecified chemical species that serves as a nutrient for the tumour,ϕ∈ [−1,1]denotes the difference in volume fractions, with{ϕ=1}representing unmixed tumour tissue, and {ϕ= −1}representing the surrounding healthy tissue, andµdenotes the chemical potential forϕ.

Fakult¨at f¨ur Mathematik, Universit¨at Regensburg, 93040 Regensburg, Germany ({Harald.Garcke, Kei-Fong.Lam}@mathematik.uni-regensburg.de).

(3)

The non-negative constants λp, λa represent the proliferation rate and the apoptosis rate of the tumour cells, respectively, and λc represents the consumption rate of the nutrient. Here we note that these are only active in the tumour regions, and the healthy tissue does not proliferate, or consume nutrient or undergo apoptosis.

In the system (1.1),A,B, andKdenote positive constants,m(ϕ)andn(ϕ)are positive mobilities forϕ and σ, respectively, Ψ(⋅) is a potential with two equal minima at±1,σ

denotes a nutrient supply on the boundary Γ, andh(ϕ) is an interpolation function with h(−1) =0 andh(1) =1. The simplest example ish(ϕ) = 1

2(1+ϕ).

We denote χσ > 0 as the diffusivity of the nutrient, and χϕ ≥ 0 can be seen as a parameter for transport mechanisms such as chemotaxis and active uptake. To see this, we note that in (1.1a) and (1.1c), the fluxes forϕand σ are given by

qϕ∶= −m(ϕ)∇µ= −m(ϕ)∇(AΨ(ϕ) −B∆ϕ−χϕσ), qσ∶= −n(ϕ)∇(χσσ−χϕϕ),

respectively. The termm(ϕ)∇(χϕσ)inqϕ models the chemotactic response, which drives the cells towards regions of high nutrient. Meanwhile, the termn(ϕ)∇(χϕϕ) inqσ drives the nutrients to regions of high ϕ, i.e., to the tumour cells, which indicates that the nutrient is actively moving towards the tumour cells. This term can be interpreted as the active transport mechanisms which move the nutrient towards the tumour colony, see [7]

for details.

We note that in (1.1), the mechanism of chemotaxis and active transport are connected via the parameter χϕ. To “decouple” the two mechanisms, we introduce the following choice for the mobilityn(ϕ)and diffusion coefficientχσ. For a positive constantη>0 and a positive mobilityD(ϕ), consider

n(ϕ) =ηD(ϕ)χ−1ϕ , χσ−1χϕ. (1.2) Then, the corresponding fluxes forϕand σ are now given as

qϕ∶= −m(ϕ)∇(AΨ(ϕ) −B∆ϕ−χϕσ),

qσ ∶= −D(ϕ)∇(σ−ηϕ), (1.3)

where the parameterχϕ controls the effects of chemotaxis, and the parameter η controls the effects of active transport.

We introduce the free energy N for the nutrient as N(ϕ, σ) = χσ

2 ∣σ∣2ϕσ(1−ϕ), (1.4) and its partial derivatives with respect toσ and ϕare given as

Nσσ+χϕ(1−ϕ), N= −χϕσ. (1.5) Note that, by the boundary condition∇ϕ⋅ν =0 on Γ, and the definition ofN (1.5), we have

∇N⋅ν=χσ∇σ⋅ν−χϕ∇ϕ⋅ν=χσ∇σ⋅ν on Γ.

Thus, by testing (1.1c) withN, (1.1b) with∂tϕ, (1.1a) withµ, and summing the resulting equations, one can show the following formal energy identity is satisfied,

d dt ∫

[AΨ(ϕ) +B

2 ∣∇ϕ∣2+ χσ

2 ∣σ∣2ϕσ(1−ϕ)]dx + ∫m(ϕ) ∣∇µ∣2+n(ϕ) ∣∇N2 dx+ ∫

ΓKN(σ−σ)dHd−1 + ∫−µ(λpσ−λa)h(ϕ) +λcσh(ϕ)Ndx =0,

(1.6)

(4)

whereHd−1is the(d−1)-dimensional Hausdorff measure. To derive useful a priori estimates from (1.6) we face a number of obstacles:

1. the presence of source termsµh(ϕ)(λa−λpσ) +Nλcσh(ϕ)deprives (1.6) of a Lya- punov structure, i.e., an inequality of the form dtd V ≤αV, for α≥0 and a suitable function V;

2. the termσ(1−ϕ) in the nutrient free energy N(ϕ, σ) can have a negative sign;

3. the presence of triple productsµσh(ϕ)and σh(ϕ)N.

One way to control the triple products with the usual H1-regularity expected fromσ, ϕ andµ is to assume thath(⋅)is bounded. The simplest choice is

h(ϕ) =min(0,max(1

2(ϕ+1),1)),

which ensuresh(−1) =0 andh(1) =1 as requested. By considering the bounded functions h(⋅), we can control the source terms µh(ϕ)(λa−λpσ) +Nλcσh(ϕ) in (1.6), and thus applications of H¨older’s inequality and Young’s inequality will lead to (see (3.12) below)

d dt ∫

[AΨ(ϕ) +B

2 ∣∇ϕ∣2σ

2 ∣σ∣2ϕσ(1−ϕ)]dx +k1(∥∇µ∥2L2(Ω)+ ∥∇N2L2(Ω)+ ∥σ∥2L2(Γ))

−k2∥σ∥2L2(Ω)−k3∥ϕ∥2L2(Ω)−k4∥∇ϕ∥2L2(Ω)≤C,

(1.7)

for some positive constantsk1, k2, k3, k4andC. The sign indefiniteness of the termχϕσ(1− ϕ) means that we have to first integrate (1.7) in time and then estimate with H¨older’s inequality and Young’s inequality. Thus, we obtain

A∥Ψ(ϕ)∥L1(Ω)+B

2∥∇ϕ∥2L2(Ω)+k5∥σ∥2L2(Ω)−k6∥ϕ∥2L2(Ω) +k1

T

0 (∥∇µ∥2L2(Ω)+ ∥∇N2L2(Ω)+ ∥σ∥2L2(Γ))dt

−k2∥σ∥2L2(0,T;L2(Ω))−k3∥ϕ∥2L2(0,T;L2(Ω))−k4∥∇ϕ∥2L2(0,T;L2(Ω))≤C,

(1.8)

for some positive constantsk5,k6 and C. A structural assumption (2.4) on the potential Ψ will allow us to control∥ϕ∥2L2(Ω) with∥Ψ∥L1(Ω) (see (3.16) below). This will lead to

(A−k7)∥Ψ(ϕ)∥L1(Ω)+ B

2∥∇ϕ∥2L2(Ω)+k5∥σ∥2L2(Ω)

+k1

T 0

(∥∇µ∥2L2(Ω)+ ∥∇N2L2(Ω)+ ∥σ∥2L2(Γ))dt

−k2∥σ∥2L2(0,T;L2(Ω))−k8∥Ψ(ϕ)∥L1(0,T;L1(Ω))−k4∥∇ϕ∥2L2(0,T;L2(Ω))≤C,

(1.9)

for some positive constants k7, k8 and C. To apply the integral version of Gronwall’s inequality, we have to assume that the constantAsatisfiesA>k7. This is needed in order to derive the usual a priori bounds for ϕ and µ in Cahn–Hilliard systems with source terms. However, we point out that, the constant A is often chosen to be A ∶= γ

ε, where γ>0 denotes the surface tension andε>0 is a small parameter related to the interfacial thickness. For sufficiently small values of εor sufficiently large surface tension γ, we see thatA>k7 will be satisfied, and thus it is not an unreasonable constraint.

(5)

Let us consider the nutrient equation (1.1c) with the specific choice of fluxes (1.2), leading to

tσ= div(D(ϕ)∇σ) −ηdiv(D(ϕ)∇ϕ) −λcσh(ϕ).

Performing a non-dimensionalisation leads to the following non-dimensionalised nutrient equation (here we reuse the same notation to denote the non-dimensionalised variables)

κ∂tσ=∆σ−θ∆ϕ−ασh(ϕ), (1.10) whereκ>0 represents the ratio between the nutrient diffusion time-scale and the tumour doubling time-scale, θ > 0 represents the ratio between the nutrient diffusion time-scale and the active transport time-scale, and α>0 represents the ratio between the nutrient diffusion time-scale and the nutrient consumption time-scale.

In practice, experimental parameters estimate that κ ≪ 1 (see for example [3, Sec- tion 4.3.2]) and we assume that the time-scale of nutrient active transport and nutrient consumption is of the same order as the time-scale of nutrient diffusion, i.e., θ ∼ O(1), α∼ O(1). This leads to the following quasi-static model,

tϕ= div(m(ϕ)∇µ) + (λpσ−λa)h(ϕ) in Ω× (0, T), (1.11a) µ=AΨ(ϕ) −B∆ϕ−χϕσ in Ω× (0, T), (1.11b) 0= div(D(ϕ)∇σ) −ηdiv(D(ϕ)∇ϕ) −λcσh(ϕ)in Ω× (0, T), (1.11c) 0= ∇ϕ⋅ν= ∇µ⋅ν on Γ× (0, T), (1.11d) D(ϕ)∇σ⋅ν=K(σ−σ) on Γ× (0, T). (1.11e) Note that the loss of the time derivative ∂tσ implies that an energy identity for (1.11) cannot be derived in a similar fashion as (1.6). However, if we test (1.11b) with ∂tϕ, (1.11a) withχϕσ+µ, (1.11c) withσ and add the resulting equations, we formally obtain

d dt ∫

[AΨ(ϕ) + B

2 ∣∇ϕ∣2] dx

+ ∫m(ϕ)(∣∇µ∣2ϕ∇µ⋅ ∇σ) +D(ϕ)(∣∇σ∣2−η∇ϕ⋅ ∇σ)dx + ∫

a−λpσ)h(ϕ)(χϕσ+µ) +λch(ϕ) ∣σ∣2 dx + ∫ΓK(∣σ∣2−σσ)dHd−1 =0.

(1.12)

Here, we point out that there are no terms with indefinite sign under the time derivative, and so we expect that there will not be a restriction on the constant A as in the model (1.1).

We now compare (1.1) with the other models for tumour growth studied in the litera- ture. In [8], the authors derived the following model,

tϕ= div(m(ϕ)∇µ) +P(ϕ)(χσσ+χϕ(1−ϕ) −µ), (1.13a)

µ=AΨ(ϕ) −B∆ϕ−χϕσ, (1.13b)

tσ= div(n(ϕ)(χσ∇σ−χϕ∇ϕ)) −P(ϕ)(χσσ+χϕ(1−ϕ) −µ), (1.13c) where we see that the chemical potentialsN and µenter as source terms in (1.13a) and (1.13c), andP(ϕ) is a non-negative function. Subsequently, if we consider

χσ=1, χϕ=0, n(ϕ) =m(ϕ) =1

(6)

in (1.13), then we obtain

tϕ=∆µ+P(ϕ)(σ−µ), (1.14a)

µ=AΨ(ϕ) −B∆ϕ, (1.14b)

tσ=∆σ−P(ϕ)(σ−µ). (1.14c) Furnishing (1.14) with homogeneous Neumann boundary conditions, the well-posedness of the system and the existence of the global attractor have been proved in [6] for large classes of nonlinearities Ψ andP.

The corresponding viscosity regularised version of (1.14) (where there is an extraα∂tµ term on the left-hand side of (1.14a) and an extra α∂tϕ term on the right-hand side of (1.14b) for positive constant α) has been studied in [4], where well-posedness is proved for a general class of potentials Ψ, and for a Lipschitz and globally bounded P. The asymptotic behaviour asα→0 is shown under more restrictions on Ψ (polynomial growth of order 4) and the authors proved that a sequence of weak solutions to the viscosity regularised system converges to the weak solution of (1.14).

For (1.14), there is a natural Lyapunov-type energy equality given as d

dt ∫

[AΨ(ϕ) +B

2 ∣∇ϕ∣2+1

2∣σ∣2]dx + ∥∇µ∥2L2(Ω)+ ∥∇σ∥2L2(Ω)+ ∫

P(ϕ)(σ−µ)2dx =0.

(1.15) Since all the terms are non-negative, the standard a priori estimates can be obtained even in the case where Ψ has polynomial growth of order 6 in three dimensions. In contrast, for (1.1) we have to assume that the derivative Ψ has linear growth, and thus restricting our class of potentials to those with at most quadratic growth (see Section 7 below).

The quasi-static model (1.11) bears the most resemblance to [5, Equations (68)-(70)]

when the active transport is neglected (i.e.,η=0). We note that the focus of study seems to be the linear stability of radial solutions to the resulting sharp interface limit when we set A= 1

ε and B =ε, and send ε→0. To the best of our knowledge, there are no results concerning the well-posedness of (1.11).

We also mention another class of models that describes tumour growth using a Cahn–

Hilliard–Darcy system,

divv= S, (1.16a)

v= −∇p+µ∇ϕ, (1.16b)

tϕ+div(vϕ) = ∇ ⋅ (m(ϕ)∇µ) + S, (1.16c)

µ=AΨ(ϕ) −B∆ϕ, (1.16d)

wherev denote a mixture velocity,pdenotes the pressure, andS denotes a mass exchange term. For the case where S = 0, the existence of strong solutions in 2D and 3D have been studied in [10], while for the case where S ≠ 0 is prescribed, existence of global weak solutions in 2D and 3D, and unique local strong solutions in 2D can be found in [9]. A related system, known as the Cahn–Hilliard–Brinkman system where an additional viscosity term is added to the left-hand side of the velocity equation (1.16b) and the mass exchangeS is set to zero, has been the subject of study in [2].

The structure of this paper is as follows. In Section 2, we state the assumptions and the well-posedness results for (1.1) and (1.11). In Section 3 we derive some useful estimates, and in Section 4, we prove the existence of weak solutions to (1.1) via a Galerkin procedure.

Continuous dependence on initial and boundary data for (1.1) is shown in Section 5. In Section 6, we outline the proof of well-posedness for (1.11), and in Section 7 we discuss the issue of the growth assumptions for the potential.

(7)

2 Main results

For any d∈N, let Ω⊂Rd denote a bounded domain with Lipschitz boundary Γ, and let T >0. We recall the Poincar´e inequalities (see for instance [14, Equations (1.35), (1.37a) and (1.37c)]): There exists a positive constantCP, depending only on Ω and the dimension d, such that for allf ∈H1(Ω),

∥f−f∥

L2(Ω)≤CP∥∇f∥L2(Ω), (2.1)

∥f∥L2(Ω)≤CP(∥∇f∥L2(Ω)+ ∥f∥L2(Γ)), (2.2) wheref ∶= 1

∣Ω∣fdx denotes the mean off.

Assumption 2.1. We assume that ϕ0 ∈ H1(Ω), σ0 ∈ L2(Ω), m, n, h, D ∈ C0(R), and there exist positive constantsh,m0, m1, D0, D1, n0 and n1, such that

m0≤m(t) ≤m1, n0≤n(t) ≤n1, D0≤D(t) ≤D1, 0≤h(t) ≤h ∀t∈R. (2.3) We assume that σ ∈ L2(0, T;L2(Γ)) and Ψ ∈ C1,1(R) is non-negative, continuously differentiable, with globally Lipschitz derivative and satisfies

Ψ(t) ≥R1∣t∣2−R2, ∣Ψ(t)∣ ≤R3(1+ ∣t∣), (2.4) for positive constants R2, R3 and a positive constant R1 such that

A>

2ϕ

χσR1. (2.5)

Definition 2.1. We call a triplet of functions (ϕ, µ, σ) a weak solution to (1.1) if σ, ϕ∈H1(0, T;(H1(Ω))) ∩L2(0, T;H1(Ω)),

µ∈L2(0, T;H1(Ω)) such that forζ, φ, ξ∈H1(Ω) and a.e. t∈ (0, T),

⟨∂tϕ, ζ⟩ = ∫

−m(ϕ)∇µ⋅ ∇ζ+ (λpσ−λa)h(ϕ)ζdx, (2.6a)

µφdx = ∫

(ϕ)φ+B∇ϕ⋅ ∇φ−χϕσφdx, (2.6b)

⟨∂tσ, ξ⟩ = ∫

−n(ϕ)(χσ∇σ−χϕ∇ϕ) ⋅ ∇ξ−λcσh(ϕ)ξdx (2.6c) + ∫ΓξK(σ−σ)dHd−1,

where ⟨⋅,⋅⟩ denotes the duality pairing between H1(Ω) and its dual (H1(Ω)).

Theorem 2.1(Existence of global weak solutions). Let Ω⊂Rdbe a bounded domain with Lipschitz boundary Γ and let T > 0. Suppose Assumption 2.1 is satisfied. Then, there exists a triplet of functions (ϕ, µ, σ) such that

ϕ∈L(0, T;H1(Ω)) ∩H1(0, T;(H1(Ω))), (2.7)

µ∈L2(0, T;H1(Ω)), (2.8)

σ∈L2(0, T;H1(Ω)) ∩L(0, T;L2(Ω)) ∩H1(0, T;(H1(Ω))), (2.9) and is a weak solution of (1.1)in the sense of Definition 2.1 with ϕ(0) =ϕ0,σ(0) =σ0 in L2(Ω).

(8)

The embedding of L2(0, T;H1(Ω)) ∩H1(0, T;(H1(Ω))) into C([0, T];L2(Ω)) guar- antees that the initial data are meaningful. We point out that the assumption (2.5) arises from using Young’s inequality to estimate the termχϕσ(1−ϕ)in (1.6), and is by no means an optimal assumption. See Remark 3.1 for more details. In addition, Theorem 2.1 gives existence of weak solutions in any dimension. This is thanks to the fact that Ψ has linear growth (see(2.4)2).

Next, we show continuous dependence on initial and boundary data and uniqueness of weak solutions under additional assumptions on the interpolation function h(⋅) and the mobilitiesm(⋅)and n(⋅).

Theorem 2.2 (Continuous dependence and uniqueness). Let d ≤ 4. Suppose h(⋅) ∈ C0,1(R), m(⋅) and n(⋅) are constant mobilities (without loss of generality we set m(⋅) = n(⋅) =1). For i=1,2, let

ϕi∈L(0, T;H1(Ω)) ∩H1(0, T;(H1(Ω))), µi∈L2(0, T;H1(Ω)),

σi∈L2(0, T;H1(Ω)) ∩L(0, T;L2(Ω)) ∩H1(0, T;(H1(Ω)))

denote two weak solutions of (1.1) satisfying (2.6)with corresponding initial dataϕi(0) = ϕ0,i∈H1(Ω), σi(0) =σ0,i∈L2(Ω), and boundary dataσ∞,i∈L2(0, T;L2(Γ)). Then,

sup

s∈[0,T]

(∥σ1(s) −σ2(s)∥2L2(Ω)+ ∥ϕ1(s) −ϕ2(s)∥2L2(Ω)) + ∥µ1−µ22L2(0,T;L2(Ω))+ ∥∇(σ1−σ2)∥2L2(0,T;L2(Ω)) + ∥σ1−σ22L2(0,T;L2(Γ))+ ∥∇(ϕ1−ϕ2)∥2L2(0,T;L2(Ω))

≤C(∥σ0,1−σ0,22L2(Ω)+ ∥ϕ0,1−ϕ0,22L2(Ω)+ ∥σ∞,1−σ∞,22L2(0,T;L2(Γ))), where the constantC depends on∥σiL(0,T;L2(Ω)), T, K,h, Ω, d, A,B, λp, λc, λaϕ, χσ, and Lh, LΨ which denote the Lipschitz constants of h and Ψ, respectively.

We point out that Theorem 2.2 provides continuous dependence for the difference of the chemical potentials ∥µ1 −µ2L2(Ω×(0,T)) and also with a stronger norm ∥ϕ1(t) − ϕ2(t)∥L(0,T;L2(Ω)) for the difference of the order parameters. This is in contrast with the classical norm∥ϕ1(t) −ϕ2(t)∥L(0,T;(H1(Ω)))one obtains for the Cahn–Hilliard equation, compare [6, Theorem 2].

We will now consider the quasi-static system (1.11).

Definition 2.2. We call a triplet of functions (ϕ, µ, σ) a weak solution to (1.11) if ϕ∈H1(0, T;(H1(Ω))) ∩L2(0, T;H1(Ω)),

σ, µ∈L2(0, T;H1(Ω)) such that forζ, λ, ξ∈H1(Ω) and a.e. t∈ (0, T),

⟨∂tϕ, ζ⟩ = ∫

−m(ϕ)∇µ⋅ ∇ζ+ (λpσ−λa)h(ϕ)ζdx, (2.10a)

µλdx = ∫

(ϕ)λ+B∇ϕ⋅ ∇λ−χϕσλdx, (2.10b)

ΓξK(σ−σ)dHd−1 = ∫

D(ϕ)(∇σ−η∇ϕ) ⋅ ∇ξ+λcσh(ϕ)ξdx. (2.10c)

(9)

Theorem 2.3(Existence and regularity of global weak solutions). LetΩ⊂Rdbe a bounded domain with Lipschitz boundaryΓ and letT >0. Suppose Assumption 2.1 is satisfied, and let A be a positive constant which need not satisfy (2.5). Then, there exists a triplet of functions (ϕ, µ, σ) such that

ϕ∈L(0, T;H1(Ω)) ∩H1(0, T;(H1(Ω))), (2.11)

µ, σ∈L2(0, T;H1(Ω)), (2.12)

and is a weak solution of (1.11) in the sense of Definition 2.2 with ϕ(0) =ϕ0 in L2(Ω). Furthermore, ifσ∈L(0, T;L2(Γ)), then

σ∈L(0, T;H1(Ω)). (2.13)

In Section 6 we derive the a priori estimates and deduce the existence of approximate solutions on the Galerkin level. The proof of Theorem 2.3 then follows from standard compactness results. In Section 6.4, we show the continuous dependence on initial and boundary data and uniqueness under additional assumptions.

Theorem 2.4 (Continuous dependence and uniqueness). Let d ≤ 4. Suppose h(⋅) ∈ C0,1(R), m and D are constant mobilities (without loss of generality we set m = 1). Fori=1,2, let

ϕi∈L(0, T;H1(Ω)) ∩H1(0, T;(H1(Ω))), µi∈L2(0, T;H1(Ω)),

σi∈L(0, T;H1(Ω))

denote two weak solutions of (1.11)satisfying(2.10)with corresponding initial dataϕi(0) = ϕ0,i∈H1(Ω) and boundary dataσ∞,i∈L(0, T;L2(Γ)). Then,

sup

s∈[0,T]

∥ϕ1(s) −ϕ2(s)∥2L2(Ω)+ ∥µ1−µ22L2(0,T;L2(Ω))+ ∥∇(ϕ1−ϕ2)∥2L2(0,T;L2(Ω)) + ∥∇(σ1−σ2)∥2L2(0,T;L2(Ω))+ ∥σ1−σ22L2(0,T;L2(Γ))

≤C(∥ϕ0,1−ϕ0,22L2(Ω)+ ∥σ∞,1−σ∞,22L2(0,T;L2(Γ))),

where the constant C depends on ∥σiL(0,T;H1(Ω)), K,Ω, A, B, Lh, LΨp, λca, χϕ, andT.

3 Useful estimates

We will use a modified version of Gronwall’s inequality in integral form.

Lemma 3.1. Let α, β, uandv be real-valued functions defined onI ∶= [0, T]. Assume that α is integrable, β is nonnegative and continuous, u is continuous, v is nonnegative and integrable. Supposeu and v satisfy the integral inequality

u(s) + ∫

s

0 v(t)dt ≤α(s) + ∫

s

0 β(t)u(t)dt ∀s∈I. (3.1) Then, it holds that

u(s) + ∫

s

0 v(t)dt ≤α(s) + ∫

s

0 α(t)β(t)exp(∫

s

t β(r)dr)dt. (3.2)

(10)

This differs from the usual Gronwall’s inequality in integral form by an extra term

s

0 v(t)dt on the left-hand side.

Proof. Let

w(s) ∶=u(s) + ∫

s

0 v(t)dt. Then, by (3.1) and the non-negativity ofβ and v, it holds that

w(s) ≤α(s) + ∫

s

0 β(t)w(t)dt.

Applying the standard Gronwall’s inequality in integral form yields the required result.

Below we will derive the first a priori estimate for sufficiently smooth solutions to (1.1), in particular this will hold for the Galerkin approximations in Section 4.1. We choose to present this estimate here due to the length of the derivation.

Lemma 3.2. Suppose Assumption 2.1 is satisfied. Let ϕ, σ ∈ C1([0, T];H1(Ω)), µ ∈ C0([0, T];H1(Ω)) be such that the triplet (ϕ, µ, σ) satisfies (2.6) with ϕ(0) = ϕ0 and σ(0) = σ0. Then, there exists a positive constant C depending on T, Ω, Γ, d, R1, R2, R3, the parameters λp, λa, λc, χσ, χϕ, h, m0, n0, A, B, K, the initial-boundary data

∥σL2(0,T;L2(Γ)), ∥ϕ(0)∥H1(Ω) and ∥σ(0)∥L2(Ω), such that for all s∈ (0, T],

∥Ψ(ϕ(s))∥L1(Ω)+ ∥ϕ(s)∥2H1(Ω)+ ∥σ(s)∥2L2(Ω)

+ ∥∇µ∥2L2(0,s;L2(Ω))+ ∥∇σ∥2L2(0,s;L2(Ω))+ ∥σ∥2L2(0,s;L2(Γ)) ≤C. (3.3) Proof. Let us denote

c0∶= ∫

[AΨ(ϕ0) +B

2 ∣∇ϕ02σ

2 ∣σ02ϕσ0(1−ϕ0)]dx (3.4) as the initial energy. Then, by the assumption on theϕ0 and σ0, H¨older’s inequality and Young’s inequality we see thatc0 is bounded.

Substituting ζ =µ, φ=∂tϕ, and ξ =χσσ+χϕ(1−ϕ) =N into (2.6) and adding the resulting equations together, we obtain

d dt ∫

[AΨ(ϕ) +B

2 ∣∇ϕ∣2σ

2 ∣σ∣2ϕσ(1−ϕ)]dx + ∫m(ϕ) ∣∇µ∣2+n(ϕ) ∣χσ∇σ−χϕ∇ϕ∣2 dx + ∫

Γσ∣σ∣2 dHd−1 + ∫h(ϕ) (λcσ(χσσ+χϕ(1−ϕ)) − (λpσ−λa)µ)dx

− ∫ΓK(χσσ+χϕ(1−ϕ))σ−Kχϕ(1−ϕ)σdHd−1 =0.

(3.5)

We first estimate the mean µ using (2.6b) by considering φ = 1 and using the growth condition (2.4), leading to

∥µ∥2L2(Ω)= ∣µ∣2∣Ω∣ = ∣Ω∣−1∣∫

(ϕ) −χϕσdx∣

2

≤ ∣Ω∣−1(AR3∣Ω∣ +AR3∥ϕ∥L2(Ω)∣Ω∣

1

2ϕ∥σ∥L2(Ω)∣Ω∣

1 2)

2

≤3∣Ω∣−1(A2R23∣Ω∣2+A2R23∥ϕ∥2L2(Ω)∣Ω∣ +χ2ϕ∥σ∥2L2(Ω)∣Ω∣).

(11)

Employing the Poincar´e inequality (2.1) we have

∥µ∥2L2(Ω)≤2CP2∥∇µ∥2L2(Ω)+2∥µ∥2L2(Ω)

≤2CP2∥∇µ∥2L2(Ω)+6(A2R23∣Ω∣ +A2R23∥ϕ∥2L2(Ω)2ϕ∥σ∥2L2(Ω)). (3.6) Then, by H¨older’s inequality and Young’s inequality, we can estimate the source term involvingµas follows,

∣∫

−h(ϕ)(λpσ−λa)µdx∣ ≤hp∥σ∥L2(Ω)a∣Ω∣

1

2) ∥µ∥L2(Ω)

≤ h2λ2p

4a1 ∥σ∥2L2(Ω)+C(a2, λa, h,∣Ω∣) + (a1+a2)∥µ∥2L2(Ω)

≤2CP2(a1+a2)∥∇µ∥2L2(Ω)+C(a1, a2, λa, h,∣Ω∣, A, R3) + (

h2λ2p

4a1 +6(a1+a22ϕ) ∥σ∥2L2(Ω)+6A2R23(a1+a2)∥ϕ∥2L2(Ω),

(3.7)

for some positive constantsa1 anda2 yet to be determined. For the term involvingλc, we obtain from H¨older’s inequality and Young’s inequality

∣∫λch(ϕ)σ(χσσ+χϕ(1−ϕ))dx∣

≤λchσ∥σ∥2L2(Ω)ϕ∥ϕ∥L2(Ω)∥σ∥L2(Ω)ϕ∥σ∥L1(Ω))

≤λchσ+a4+ a3χϕ

2 ) ∥σ∥2L2(Ω)ch

χϕ

2a3∥ϕ∥2L2(Ω)+C(∣Ω∣, λc, h, χσ, χϕ, a4), (3.8)

for some positive constants a3 and a4 yet to be determined. For the terms involving the boundary integral, we have by H¨older’s inequality, Young’s inequality and the trace theorem,

∣∫Γχϕ(1−ϕ)σ−χσσσ−χϕ(1−ϕ)σdHd−1

≤χϕ(∥σ∥L1(Γ)+ ∥ϕ∥L2(Γ)∥σ∥L2(Γ)) +χσ∥σ∥L2(Γ)∥σL2(Γ)

ϕ∥σL1(Γ)ϕ∥ϕ∥L2(Γ)∥σL2(Γ)

≤ (a5σ

2 ) ∥σ∥2L2(Γ)+ ( χ2ϕσ

+a6) ∥ϕ∥2L2(Γ)+C(a5, a6, χϕ, χσ,∣Γ∣) (1+ ∥σ2L2(Γ))

≤ (a5σ

2 ) ∥σ∥2L2(Γ)+Ctr2 ( χ2ϕ

σ +a6) ∥ϕ∥2H1(Ω)+C(1+ ∥σ2L2(Γ)),

(3.9)

for some positive constantsa5 anda6yet to be determined. Here,Ctris the constant from the trace theorem which depends only on Ω andd,

∥f∥L2(Γ)≤Ctr∥f∥H1(Ω) ∀f ∈H1(Ω).

Employing the estimates (3.7), (3.8), and (3.9) into (3.5), and using the lower bounds of

(12)

m(⋅) andn(⋅), we have d

dt ∫

[AΨ(ϕ) + B

2 ∣∇ϕ∣2+ χσ

2 ∣σ∣2ϕσ(1−ϕ)]dx + ∫

(m0−2CP2(a1+a2)) ∣∇µ∣2+n0∣χσ∇σ−χϕ∇ϕ∣2 dx +K∫

Γσ−a5− χσ

2 ) ∣σ∣2 dHd−1 −K∫

Ctr2 ( χ2ϕ

σ +a6) ∣∇ϕ∣2 dx

− ∫

( h2λ2p

4a1

+6(a1+a22ϕchσ+a4+a3χϕ

2 )) ∣σ∣2 dx

− ∫

(6A2R23(a1+a2) +λch

χϕ

2a3 +KCtr2( χ2ϕ

σ +a6)) ∣ϕ∣2 dx

≤C(1+ ∥σ2L2(Γ)),

(3.10)

whereC is independent of ϕ,σ andµ. By the triangle inequality, Minkowski’s inequality and Young’s inequality, we see that

∥χσ∇σ∥2L2(Ω)≤ (∥∇NL2(Ω)+ ∥χϕ∇ϕ∥L2(Ω))2≤2∥∇N2L2(Ω)+2∥χϕ∇ϕ∥2L2(Ω). (3.11) We now choose the constants{ai}6i=1 to be

a1=a2= m0

8CP2, a5= χσ

4 , a3=a4=a6=1, and denote

c1∶=

m0

2 , c2∶=Kχσ

4 , c3 ∶=KCtr2 ( χ2ϕ

σ +1) +χ2ϕn0, c4∶=

2h2λ2pCP2 m0

+ 3m0

2CP2χ2ϕchσ+1+ χϕ

2 ), c5∶=

3m0

2CP2A2R23ch

χϕ

2 +KCtr2 ( χ2ϕσ +1),

where the additionalχ2ϕn0 in the constantc3 comes from (3.11). Then (3.10) becomes d

dt ∫

[AΨ(ϕ) +B

2 ∣∇ϕ∣2σ

2 ∣σ∣2ϕσ(1−ϕ)]dx + ∫c1∣∇µ∣2+n0χ2σ

2 ∣∇σ∣2dx + ∫

Γc2∣σ∣2dHd−1

− ∫c4∣σ∣2+c5∣ϕ∣2+c3∣∇ϕ∣2 dx ≤C(1+ ∥σ2L2(Γ)).

(3.12)

Upon integrating with respect totfrom 0 tos∈ (0, T] gives

[AΨ(ϕ(x, s)) +B

2 ∣∇ϕ(x, s)∣2σ

2 ∣σ(x, s)∣2ϕσ(x, s)(1−ϕ(x, s))]dx +c1∥∇µ∥2L2(0,s;L2(Ω))+n0χ2σ

2 ∥∇σ∥2L2(0,s;L2(Ω))+c2∥σ∥2L2(0,s;L2(Γ))

−c4∥σ∥2L2(0,s;L2(Ω))−c5∥ϕ∥2L2(0,s;L2(Ω))−c3∥∇ϕ∥2L2(0,s;L2(Ω))

≤c0+C(s+ ∥σ2L2(0,s;L2(Γ))),

(3.13)

(13)

where the constantc0 is defined in (3.4). By H¨older’s inequality and Young’s inequality, we have

∣∫χϕσ(1−ϕ)dx∣ ≤χϕ∥σ∥L1(Ω)ϕ∥σ∥L2(Ω)∥ϕ∥L2(Ω)

≤ χσ

8 ∥σ∥2L2(Ω)+C(χσ,∣Ω∣, χϕ) + χσ

8 ∥σ∥2L2(Ω)+ 2χ2ϕ

χσ ∥ϕ∥2L2(Ω),

(3.14)

and thus from (3.13) we deduce that A∥Ψ(ϕ(s))∥L1(Ω)+

B

2∥∇ϕ(s)∥2L2(Ω)+ χσ

4 ∥σ(s)∥2L2(Ω)− 2χ2ϕ

χσ ∥ϕ(s)∥2L2(Ω)

+c1∥∇µ∥2L2(0,s;L2(Ω))+ n0χ2σ

2 ∥∇σ∥2L2(0,s;L2(Ω))+c2∥σ∥2L2(0,s;L2(Γ))

−c4∥σ∥2L2(0,s;L2(Ω))−c5∥ϕ∥2L2(0,s;L2(Ω))−c3∥∇ϕ∥2L2(0,s;L2(Ω))

≤c0+C(1+T+ ∥σ2L2(0,T;L2(Γ))).

(3.15)

Now, by (2.4) we have

∥ϕ∥2L2(Ω)= ∫

∣ϕ∣2dx ≤ 1 R1(∫

Ψ(ϕ)dx +R2∣Ω∣) = 1 R1

∥Ψ(ϕ)∥L1(Ω)+R2 R1

∣Ω∣, (3.16) and for anys∈ (0, T],

∥ϕ∥2L2(0,s;L2(Ω))≤ 1

R1∥Ψ(ϕ)∥L1(0,s;L1(Ω))+R2

R1∣Ω∣s. (3.17) Thus, using (3.16) and (3.17), we obtain from (3.15)

(A− 2χ2ϕ χσR1

)∥Ψ(ϕ(s))∥L1(Ω)+B

2∥∇ϕ(s)∥2L2(Ω)σ

4 ∥σ(s)∥2L2(Ω)

− c5

R1∥Ψ(ϕ(s))∥L1(0,s;L1(Ω))−c3∥∇ϕ∥2L2(0,s;L2(Ω))−c4∥σ∥2L2(0,s;L2(Ω)) +c1∥∇µ∥2L2(0,s∶L2(Ω))+n0χ2σ

2 ∥∇σ∥2L2(0,s;L2(Ω))+c2∥σ∥2L2(0,s;L2(Γ))

≤C(1+T+ ∥σ2L2(0,T;L2(Γ))) =∶c,

(3.18)

for some positive constantc independent of s∈ (0, T],µ(s),σ(s), andϕ(s). Let cmin∶=min(A−

2ϕ χσR1,B

2,χσ

4 ), cmax∶=max(c5/R1, c3, c4). (3.19) Then,cmin>0 by assumption (see (2.5)), and we obtain from (3.18) that,

cmin(∥Ψ(ϕ(s))∥L1(Ω)+ ∥∇ϕ(s)∥2L2(Ω)+ ∥σ(s)∥2L2(Ω)) +c1∥∇µ∥2L2(0,s;L2(Ω))+n0χ2σ

2 ∥∇σ∥2L2(0,s;L2(Ω))+c2∥σ∥2L2(0,s;L2(Γ))

≤ ∫

s

0 cmax(∥Ψ(ϕ)∥L1(Ω)+ ∥∇ϕ∥2L2(Ω)+ ∥σ∥2L2(Ω))dt +c.

(3.20)

(14)

Substituting

u(s) = ∥Ψ(ϕ(s))∥L1(Ω)+ ∥∇ϕ(s)∥2L2(Ω)+ ∥σ(s)∥2L2(Ω), (3.21) v(t) = 1

cmin

(c1∥∇µ∥2L2(Ω)+n0χ2σ

2 ∥∇σ∥2L2(Ω)+c2∥σ∥2L2(Γ)), (3.22) α(s) = c

cmin, β(t) = cmax

cmin, (3.23)

into Lemma 3.1, we obtain from (3.20)

∥Ψ(ϕ(s))∥L1(Ω)+ ∥∇ϕ(s)∥2L2(Ω)+ ∥σ(s)∥2L2(Ω) + 1

cmin

(c1∥∇µ∥2L2(0,s;L2(Ω))+n0χ2σ

2 ∥∇σ∥2L2(0,s;L2(Ω))+c2∥σ∥2L2(0,s;L2(Γ)))

≤ c

cmin + ∫

s 0

ccmax

c2min exp(cmax

cmin

(s−t))dt < ∞ ∀s∈ (0, T].

(3.24)

Together with (3.16), we find that there exists a positive constantC not depending onϕ, µand σ such that,

∥Ψ(ϕ(s))∥L1(Ω)+ ∥ϕ(s)∥2H1(Ω)+ ∥σ(s)∥2L2(Ω)

+ ∥∇µ∥2L2(0,s;L2(Ω))+ ∥∇σ∥2L2(0,s;L2(Ω))+ ∥σ∥2L2(0,s;L2(Γ)) ≤C, (3.25) for alls∈ (0, T].

Remark 3.1. The necessity of (2.5)comes from the fact that in (3.12), we cannot apply H¨older’s inequality and Young’s inequality like in (3.14) to estimate the term

d dt ∫

χϕσ(1−ϕ)dx, as inequalities are not preserved under differentiation.

4 Global weak solutions

4.1 Galerkin approximation

We obtain global weak solutions via a suitable Galerkin procedure. Consider a basis {wi}i∈

NofH1(Ω)which is orthonormal with respect to theL2-inner product, and without loss of generality, we assume w1 is constant and hence ∫widx = 0 for all i ≥ 2. In the following we take {wi}i∈N to be eigenfunctions for the Laplacian with homogeneous Neumann boundary conditions,

−∆wiiwi in Ω, (4.1a)

∇wi⋅ν=0 on Γ, (4.1b)

where Λi is the eigenvalue corresponding to wi. It is well-known that the {wi}i∈N can be chosen as an orthonormal basis ofL2(Ω) and then forms an orthogonal basis of H1(Ω). As constant functions are eigenfunctions, w1 can be chosen as a constant function with Λ1=0 (see for instance [12, Theorem 8.4]). Let

Wk∶=span{w1, . . . , wk} ⊂H1(Ω)

Referenzen

ÄHNLICHE DOKUMENTE

[r]

Using these brain-on-chip devices, we show that human fetal NPCs are able to respond to shallow CXCL12 gradients only in the presence of an environment consisting of neuronal

Strongly nonlinear parabolic system, cross-diffusion, segrega- tion, existence of weak solutions, uniqueness of solutions, entropy-type estimates.. 1991 Mathematics

Thus, if χ ϕ is positive, then (1.4e) 2 tells us that the tumour cells will experience a higher level of nutrient concentration than the healthy cells on the interface, which

A first phase field model for non-matched densities and a divergence free velocity field which in addition fulfills local and hence global free energy inequalities has been derived

These simpler equations are handled with higher order discretization and the results are coupled with an operator- splitting method together.. We describe the discretization methods

Nonlocal phase separation models, viscous phase separation models, Cahn- Hilliard equation, integrodifferential equations, initial value problems, nonlinear evolution equations..

We investigate the nonlinear stability of the system by introducing frictional damping and Cattaneo’s type heat conduction as the dissipative mecha- nism, and prove the global