Processing math: 100%
Research article Special Issues

Global weak solutions and asymptotic limits of a Cahn–Hilliard–Darcy system modelling tumour growth

  • We study the existence of weak solutions to a Cahn-Hilliard-Darcy system coupled witha convection-reaction-diffusion equation through the fluxes, through the source terms and in Darcy’slaw. The system of equations arises from a mixture model for tumour growth accounting for transportmechanisms such as chemotaxis and active transport. We prove, via a Galerkin approximation, theexistence of global weak solutions in two and three dimensions, along with new regularity results forthe velocity field and for the pressure. Due to the coupling with the Darcy system, the time derivativeshave lower regularity compared to systems without Darcy flow, but in the two dimensional case weemploy a new regularity result for the velocity to obtain better integrability and temporal regularityfor the time derivatives. Then, we deduce the global existence of weak solutions for two variantsof the model; one where the velocity is zero and another where the chemotaxis and active transportmechanisms are absent.

    Citation: Harald Garcke, Kei Fong Lam. Global weak solutions and asymptotic limits of a Cahn–Hilliard–Darcy system modelling tumour growth[J]. AIMS Mathematics, 2016, 1(3): 318-360. doi: 10.3934/Math.2016.3.318

    Related Papers:

    [1] Joseph L. Shomberg . Well-posedness and global attractors for a non-isothermal viscous relaxationof nonlocal Cahn-Hilliard equations. AIMS Mathematics, 2016, 1(2): 102-136. doi: 10.3934/Math.2016.2.102
    [2] Jean De Dieu Mangoubi, Mayeul Evrard Isseret Goyaud, Daniel Moukoko . Pullback attractor for a nonautonomous parabolic Cahn-Hilliard phase-field system. AIMS Mathematics, 2023, 8(9): 22037-22066. doi: 10.3934/math.20231123
    [3] Dieunel DOR . On the modified of the one-dimensional Cahn-Hilliard equation with a source term. AIMS Mathematics, 2022, 7(8): 14672-14695. doi: 10.3934/math.2022807
    [4] Pierluigi Colli, Gianni Gilardi, Jürgen Sprekels . Distributed optimal control of a nonstandard nonlocal phase field system. AIMS Mathematics, 2016, 1(3): 225-260. doi: 10.3934/Math.2016.3.225
    [5] Aymard Christbert Nimi, Daniel Moukoko . Global attractor and exponential attractor for a Parabolic system of Cahn-Hilliard with a proliferation term. AIMS Mathematics, 2020, 5(2): 1383-1399. doi: 10.3934/math.2020095
    [6] Martin Stoll, Hamdullah Yücel . Symmetric interior penalty Galerkin method for fractional-in-space phase-field equations. AIMS Mathematics, 2018, 3(1): 66-95. doi: 10.3934/Math.2018.1.66
    [7] Taishi Motoda . Time periodic solutions of Cahn-Hilliard systems with dynamic boundary conditions. AIMS Mathematics, 2018, 3(2): 263-287. doi: 10.3934/Math.2018.2.263
    [8] Mingliang Liao, Danxia Wang, Chenhui Zhang, Hongen Jia . The error analysis for the Cahn-Hilliard phase field model of two-phase incompressible flows with variable density. AIMS Mathematics, 2023, 8(12): 31158-31185. doi: 10.3934/math.20231595
    [9] Alain Miranville . The Cahn–Hilliard equation and some of its variants. AIMS Mathematics, 2017, 2(3): 479-544. doi: 10.3934/Math.2017.2.479
    [10] Scala Riccardo, Schimperna Giulio . On the viscous Cahn-Hilliard equation with singular potential and inertial term. AIMS Mathematics, 2016, 1(1): 64-76. doi: 10.3934/Math.2016.1.64
  • We study the existence of weak solutions to a Cahn-Hilliard-Darcy system coupled witha convection-reaction-diffusion equation through the fluxes, through the source terms and in Darcy’slaw. The system of equations arises from a mixture model for tumour growth accounting for transportmechanisms such as chemotaxis and active transport. We prove, via a Galerkin approximation, theexistence of global weak solutions in two and three dimensions, along with new regularity results forthe velocity field and for the pressure. Due to the coupling with the Darcy system, the time derivativeshave lower regularity compared to systems without Darcy flow, but in the two dimensional case weemploy a new regularity result for the velocity to obtain better integrability and temporal regularityfor the time derivatives. Then, we deduce the global existence of weak solutions for two variantsof the model; one where the velocity is zero and another where the chemotaxis and active transportmechanisms are absent.


    1. Introduction

    In recent years there has been an increased focus on the mathematical modelling and analysis of tumour growth. Many new models have been proposed and numerical simulations have been carried out to provide new and important insights on cancer research, see for instance [8] and [13, Chap. 3]. In this work we analyse a diffuse interface model proposed in [20], which models a mixture of tumour cells and healthy cells in the presence of an unspecified chemical species acting as a nutrient. More precisely, for a bounded domain ΩRd where the cells reside and T > 0, we consider the following set of equations,

    divv=ΓvinΩ×(0,T)=:Q, (1.1a)
    v=K(p(μ+χσ)φ) in Q, (1.1b)
    tφ+div(φv)=div(m(φ)μ)+Γφ in Q, (1.1c)
    μ=AΨ(φ)BΔφχσ in Q, (1.1d)
    tσ+div(σv)=div(n(φ)(Dσχφ))S in Q. (1.1e)

    Here, v denotes the volume-averaged velocity of the mixture, p denotes the pressure, σ denotes the concentration of the nutrient, φ[1,1] denotes the difference in volume fractions, with {φ=1} representing the unmixed tumour tissue, and {φ=1} representing the surrounding healthy tissue, and μ denotes the chemical potential for φ.

    The model treats the tumour and healthy cells as inertia-less fluids, leading to the appearance of a Darcy-type subsystem with a source term Γv. The order parameter φ satisfies a convective Cahn--Hilliard type equation with additional source term Γφ, and similarly, the nutrient concentration σ satisfies a convection-reaction-diffusion equation with a non-standard flux and a source term S. We refer the reader to [20, x2] for the derivation from thermodynamic principles, and to [20, x2.5] for a discussion regarding the choices for the source terms Γφ,Γv and S.

    The positive constants K and D denote the permeability of the mixture and the diffusivity of the nutrient, m(φ) and n(φ) are positive mobilities for φ and σ, respectively. The parameter χ0 regulates the chemotaxis effect (see [20] for more details), Ψ() is a potential with two equal minima at ±1, A and B denote two positive constants related to the thickness of the diffuse interface and the surface tension.

    We supplement the above with the following boundary and initial conditions

    nφ=nμ=0 on Ω×(0,T)=:Σ, (1.2a)
    vn=np=0 on Σ, (1.2b)
    n(φ)Dnσ=b(σσ) on Σ, (1.2c)
    φ(0)=φ0,σ(0)=σ0 on Ω. (1.2d)

    Here φ0, σ0 and σ are given functions and b>0 is a constant. We denote nf:=fn as the normal derivative of f at the boundary Ω, where n is the outer unit normal. Associated to (1.1) is the free energy density N(φ,σ) for the nutrient, which is defined as

    N(φ,σ):=D2|σ|2+χσ(1φ). (1.3)

    Note that

    N,σ:=Nσ=Dσ+χ(1φ),N,φ:=Nφ=χσ,

    so that (1.1) may also be written as

    divv=Γv, (1.4a)
    v=K(pμφ+N,φφ), (1.4b)
    tφ+div(φv)=div(m(φ)μ)+Γφ, (1.4c)
    μ=AΨ(φ)BΔφ+N,φ, (1.4d)
    tσ+div(σv)=div(n(φ)N,σ)S, (1.4e)

    which is the general phase field model proposed in [20]. In this work we do not aim to analyse such a model with a general free energy density N(φ,σ), but we will focus solely on the choice (1.3) and the corresponding model (1.1)-(1.2).

    Our goal in this work is to prove the existence of weak solutions (see Definition 2.1 below) of (1.1)-(1.2) in two and three dimensions. Moreover, one might expect that by setting Γv=0 and then sending b0 and K0, the weak solutions to (1.1)-(1.2) will converge (in some appropriate sense) to the weak solutions of

    tφ=div(m(φ)μ)+Γφ in Q, (1.5a)
    μ=AΨ(φ)BΔφχσ in Q, (1.5b)
    tσ=div(n(φ)(Dσχφ))S in Q, (1.5c)
    0=nφ=nμ=nσ on Σ. (1.5d)

    We denote (1.5) as the limit system of vanishing permeability, where the effects of the volume-averaged velocity are neglected. By substituting

    Γφ=S=f(φ)(Dσ+χ(1φ)μ) (1.6)

    for some non-negative function f(φ) leads to the model derived in [21]. The specific choices for Γφ and S in (1.6) are motivated by linear phenomenological laws for chemical reactions. The analysis of (1.5) with the parameters

    D=1,χ=0,n(φ)=m(φ)=1

    has been the subject of study in [5, 6, 7, 16], where well-posedness and long-time behaviour have been established for a large class of functions Ψ(φ) and f(φ). Alternatively, one may consider the following choice of source terms

    Γφ=h(φ)(λpσλa),S=λch(φ)σ, (1.7)

    where λp, λa, λc are non-negative constants representing the tumour proliferation rate, the apoptosis rate, and the nutrient consumption rate, respectively, and h(φ) is a non-negative interpolation function such that h(1)=0 and h(1)=1. The above choices for Γφ and S are motivated from the modelling of processes experienced by a young tumour.

    The well-posedness of model (1.5) with the choice (1.7) has been studied by the authors in [17] and [18] with the boundary conditions (1.2) (neglecting (1.2b)) in the former and for non-zero Dirichlet boundary conditions in the latter. It has been noted in [17] that the well-posedness result with the boundary conditions (1.2) requires Ψ to have at most quadratic growth, which is attributed to the presence of the source term Γφμ=h(φ)μ(λpσλa) when deriving useful a priori estimates. Meanwhile in [18] the Dirichlet boundary conditions and the application of the Poincaré inequality allows us to overcome this restriction and allow for Ψ to be a regular potential with polynomial growth of order less than 6, and by a Yosida approximation, the case where Ψ is a singular potential is also covered.

    We also mention the work of [19] that utilises a Schauder's fixed point argument to show existence of weak solutions for Ψ with quartic growth and Γφ,S as in (1.7). This is based on first deducing that σ is bounded by a comparison principle, leading to ΓφL(Ω). Then, the standard a priori estimates are derived for a Cahn--Hilliard equation with bounded source terms. The difference between [19] and [17, 18] is the absence of the chemotaxis and active transport mechanisms, i.e., χ=0, so that the comparison principle can be applied to the nutrient equation. We refer to [9] for the application of a similar procedure to a multi-species tumour model with logarithmic potentials.

    On the other hand, by sending b0 and χ0 in (1.1), we should obtain weak solutions of

    divv=Γv in Q, (1.8a)
    v=K(pμφ) in Q, (1.8b)
    tφ+div(φv)=div(m(φ)μ)+Γφ in Q, (1.8c)
    μ=AΨ(φ)BΔφ in Q, (1.8d)
    tσ+div(σv)=div(n(φ)Dσ)S in Q, (1.8e)
    0=nφ=nμ=nσ=vn on Σ. (1.8f)

    We denote (1.8) as the limit system of vanishing chemotaxis. If the source terms Γv and Γφ are independent of σ, then (1.8) consists of an independent Cahn--Hilliard--Darcy system and an equation for σ which is advected by the volume-averaged velocity field v. In the case where there is no nutrient and source terms, i.e., σ=Γv=Γφ=0, global existence of weak solutions in two and three dimensions has been established in [14] via the convergence of a fully discrete and energy stable implicit finite element scheme. For the well-posedness and long-time behaviour of strong solutions, we refer to [25]. Meanwhile, in the case where Γv=Γφ is prescribed, global weak existence and local strong well-posedness for (1.8) without nutrient is shown in [22].

    We also mention the work of [3] on the well-posedness and long-time behaviour of a related system also used in tumour growth, known as the Cahn--Hilliard--Brinkman system, where in (1.8) without nutrient an additional viscosity term is added to the left-hand side of the velocity equation (1.8b) and the mass exchange terms Γv and Γφ are set to zero. The well-posedness of a nonlocal variant of the Cahn--Hilliard--Brinkman system has been investigated in [10]. Furthermore, when K is a function depending on φ, the model (1.8) with σ=Γv=Γφ=0 is also referred to as the Hele--Shaw--Cahn--Hilliard model (see [23, 24]). In this setting, K(φ) represents the reciprocal of the viscosity of the fluid mixture. We refer to [30] concerning the strong well-posedness globally in time for two dimensions and locally in time for three dimensions when Ω is the d-dimensional torus. Global well-posedness in three dimensions under additional assumptions and long-time behaviour of solutions to the Hele--Shaw--Cahn--Hilliard model are investigated in [29].

    We point out that from the derivation of (1.1) in [20], the source terms Γv and Γφ are connected in the sense that Γv is related to sum of the mass exchange terms for the tumour and healthy cells, and Γφ is related to the difference between the mass exchange terms. Thus, if Γφ would depend on the primary variables φ, σ or μ, then one expects that Γv will also depend on the primary variables. Here, we are able to prove existence of weak solutions for Γφ of the form (2.1), which generalises the choices (1.6) and (1.7), but in exchange Γv has to be considered as a prescribed function. This is attributed to the presence of the source term Γv(φμ+D2|σ|2) when deriving useful a priori estimates. We see that if Γv depends on the primary variables, we obtain triplet products which cannot be controlled by the usual regularity of φ, μ and σ in the absence of a priori estimates.

    In this work we attempt to generalise the weak existence results for the models studied in [5, 16, 17, 18, 22, 25] by proving that the weak solutions of (1.1) with Γv=0 converge (in some appropriate sense) to the weak solutions of (1.5) as b0 and K0, and the weak solutions of (1.1) converge to the weak solutions of (1.8) as b0 and χ0.

    This paper is organised as follows. In Section 2 we state the main assumptions and the main results. In Section 3 we introduce a Galerkin procedure and derive some a priori estimates for the Galerkin ansatz in Section 4 for the case of three dimensions. We then pass to the limit in Section 5 to deduce the existence result for three dimensions, while in Section 6 we investigate the asymptotic behaviour of solutions to (1.1) as K0 and χ0. In Section 7, we outline the a priori estimates for two dimensions and show that the weak solutions for two dimensions yields better temporal regularity than the weak solutions for three dimensions. In Section 8 we discuss some of the issues present in the analysis of (1.1) using different formulations of Darcy's law and the pressure, and with different boundary conditions for the velocity and the pressure.

    Notation. For convenience, we will often use the notation Lp:=Lp(Ω) and Wk,p:=Wk,p(Ω) for any p[1,], k>0 to denote the standard Lebesgue spaces and Sobolev spaces equipped with the norms Lp and Wk,p. In the case p=2 we use Hk:=Wk,2 and the norm Hk. For the norms of Bochner spaces, we will use the notation Lp(X):=Lp(0,T;X) for Banach space X and p[1,]. Moreover, the dual space of a Banach space X will be denoted by X, and the duality pairing between X and X is denoted by ,X,X. For d=2 or 3, let Hd1 denote the (d1) dimensional Hausdorff measure on Ω, and we denote Rd-valued functions and any function spaces consisting of vector-valued/tensor-valued functions in boldface. We will use the notation Df to denote the weak derivative of the vector function f.

    Useful preliminaries. For convenience, we recall the Poincaré inequality: There exists a positive constant Cp depending only on Ω such that, for all fH1,

    fˉfL2CpfL2, (1.9)

    where ˉf:=1|Ω|Ωf dx denotes the mean of f. The Gagliardo--Nirenberg interpolation inequality in dimension d is also useful (see [15, Thm. 10.1, p. 27], [11, Thm. 2.1] and [1, Thm. 5.8]): Let Ω be a bounded domain with Lipschitz boundary, and fWm,rLq, 1q,r. For any integer j, 0j<m, suppose there is αR such that

    1p=jd+(1rmd)α+1αq,jmα1.

    Then, there exists a positive constant C depending only on Ω, m, j, q, r, and α such that

    DjfLpCfαWm,rf1αLq. (1.10)

    We will also use the following Gronwall inequality in integral form (see [17, Lem. 3.1] for a proof): Let α,β,u and v be real-valued functions defined on [0,T]. Assume that α is integrable, β is non-negative and continuous, u is continuous, v is non-negative and integrable. If u and v satisfy the integral inequality

    u(s)+s0v(t) dtα(s)+s0β(t)u(t) dt for s(0,T],

    then it holds that

    u(s)+s0v(t) dtα(s)+s0β(t)α(t)exp(t0β(r) dr) dt. (1.11)

    To analyse the Darcy system, we introduce the spaces

    L20:={fL2:ˉf=0},H2N:={fH2:nf=0 on Ω},(H1)0:={f(H1):f,1H1=0}.

    Then, the Neumann-Laplacian operator ΔN:H1L20(H1)0 is positively defined and self-adjoint. In particular, by the Lax--Milgram theorem and the Poincaré inequality (1.9) with zero mean, the inverse operator (ΔN)1:(H1)0H1L20 is well-defined, and we set u:=(ΔN)1f for f(H1)0 if ˉu=0 and

    Δu=f in Ω, nu=0 on Ω.

    2. Main results

    We make the following assumptions.

    Assumption 2.1.

    (A1) The constants A, B, K, D, χ and b are positive and fixed.

    (A2) The mobilities m, n are continuous on R and satisfy

    m0m(t)m1, n0n(t)n1tR,

    for positive constants m0, m1, n0 and n1.

    (A3) Γφ and S are of the form

    Γφ(φ,μ,σ)=Λφ(φ,σ)Θφ(φ,σ)μ,S(φ,μ,σ)=ΛS(φ,σ)ΘS(φ,σ)μ, (2.1)

    where Θφ,ΘS:R2R are continuous bounded functions with Θφ non-negative, and Λφ,ΛS:R2R are continuous with linear growth

    |Θi(φ,σ)|R0,  |Λi(φ,σ)|R0(1+|φ|+|σ|) for i{φ,S}, (2.2)

    so that

    |Γφ|+|S|R0(1+|φ|+|μ|+|σ|), (2.3)

    for some positive constant R0.

    (A4) Γv is a prescribed function belonging to L4(0,T;L20).

    (A5) ΨC2(R) is a non-negative function satisfying

    Ψ(t)R1 |t|2R2tR (2.4)

    and either one of the following,

    1: if Θφ is non-negative and bounded, then

    Ψ(t)R3(1+|t|2),  |Ψ(t)|R4(1+|t|), |Ψ(t)|R4; (2.5)

    2: if Θφ is positive and bounded, that is,

    R0Θφ(t,s)R5>0t,sR, (2.6)

    then

    |Ψ(t)|R6(1+|t|q),q[0,4), (2.7)

    for some positive constants R1, R2, R3, R4, R5, R6. Furthermore we assume that

    A>2χ2DR1. (2.8)

    (A6) The initial and boundary data satisfy

    σL2(0,T;L2(Ω)),σ0L2,φ0H1.

    We point out that some of the above assumptions are based on previous works on the well-posedness of Cahn--Hilliard systems for tumour growth. For instance, (2.5) and (2.8) reflect the situation encountered in [17], where if Θφ=0, i.e., Γφ is independent of μ, then the derivation of the a priori estimate requires a quadratic potential. But in the case where (2.6) is satisfied, we can allow Ψ to be a regular potential with polynomial growth of order less than 6, and by a Yosida approximation, we can extend our existence results to the situation where Ψ is a singular potential, see for instance [18]. Moreover, the condition (2.8) is a technical assumption based on the fact that the second term of the nutrient free energy χσ(1φ) does not have a positive sign.

    Meanwhile, the linearity of the source terms Γφ and S with respect to the chemical potential μ assumed in (2.1) is a technical assumption based on the expectation that, at best, we have weak convergence for Galerkin approximation to μ, which is in contrast with φ and σ where we might expect a.e convergence and strong convergence for the Galerkin approximations. Moreover, if we consider

    Θφ(φ,σ)=ΘS(φ,σ)=f(φ),Λφ(φ,σ)=ΛS(φ,σ)=f(φ)(Dσ+χ(1φ)),

    for a non-negative function f(φ), then we obtain the source terms in [5, 16, 21].

    Compared to the set-up in [22], in (A4) we prescribe a higher temporal regularity for the prescribed source term Γv. This is needed when we estimate the source term ΓvD2|σ|2 in the absence of a priori estimates, see Section 4.1.2 for more details. The mean zero condition is a consequence of the no-flux boundary condition vn=0 on Ω and the divergence equation (1.1a). In particular, we can express the Darcy subsystem (1.1a)-(1.1b) as an elliptic equation for the pressure p:

    Δp=1KΓvdiv((μ+χσ)φ) in Ω, (2.9a)
    np=0 on Ω. (2.9b)

    Solutions to (2.9) are uniquely determined up to an arbitrary additive function that may only depend on time, and thus without loss of generality, we impose the condition ˉp=1|Ω|Ωp dx=0 to (2.9). We may then define p as

    p=(ΔN)1(1KΓvdiv((μ+χσ)φ)), (2.10)

    if 1KΓvdiv((μ+χσ)φ)(H1)0.

    Remark 2.1. In the case Γv=0, one can also consider the assumption

    SN,σΓφμ=S(Dσ+χ(1φ))Γφμ0 (2.11)

    instead of (2.6), which holds automatically if Γφ and S are chosen to be of the form (1.6). In fact this property is used in [5, 16].

    We make the following definition.

    Definition 2.1 (Weak solutions for 3D). We call a quintuple (φ,μ,σ,v,p) a weak solution to (1.1)-(1.2) if

    φL(0,T;H1)L2(0,T;H3)W1,85(0,T;(H1)),σL(0,T;L2)L2(0,T;H1)W1,54(0,T;(W1,5)),μL2(0,T;H1),pL85(0,T;H1L20),vL2(0,T;L2),

    such that φ(0)=φ0,

    σ0,ζH1,(H1)=σ(0),ζH1,(H1)ζH1,

    and

    tφ,ζH1,(H1)=Ωm(φ)μζ+Γφζ+φvζ dx, (2.12a)
    Ωμζ dx=ΩAΨ(φ)ζ+Bφζχσζ dx, (2.12b)
    tσ,ϕW1,5,(W1,5)=Ωn(φ)(Dσχφ)ϕSϕ+σvϕdx+Ωb(σσ)ϕ dHd1, (2.12c)
    Ωpζ dx=Ω1KΓvζ+(μ+χσ)φζ dx, (2.12d)
    Ωv ζ dx=ΩK(p(μ+χσ)φ) ζ dx, (2.12e)

    for a.e. t(0,T) and for all ζH1, ϕW1,5, and ζL2.

    Neglecting the nutrient σ, we observe that our choice of function spaces for (φ,μ,p,v) coincide with those in [22, Defn. 2.1(i)]. In contrast to the usual L2(0,T;(H1))-regularity (see [5, 16]) we obtain a less regular time derivative tφ. The drop in the time regularity from 2 to 85 is attributed to the convection term div(φv) belonging to L85(0,T;(H1)). The same is true for the regularity for the time derivative tσ in L54(0,T;(W1,5)) as the convection term div(σv) lies in the same space. We refer the reader to the end of Section 4.3 for a calculation motivating the choice of function spaces for div(σv) and tσ. Furthermore, the embedding of L(0,T;H1)W1,85(0,T;(H1)) into C0([0,T];L2) from [28, x8, Cor. 4] guarantees that the initial condition for φ is meaningful. However, for σ we have the embedding L(0,T;L2)W1,54(0,T;(W1,5))⊂⊂C0([0,T];(H1)), and so σ(0) makes sense as a function in (H1). Thus, the initial condition σ0 is attained as an equality in (H1). We now state the existence result for (1.1)-(1.2).

    Theorem 2.1 (Existence of weak solutions in 3D and energy inequality). Let ΩR3 be a bounded domain with C3-boundary Ω. Suppose Assumption 2.1 is satisfied. Then, there exists a weak solution quintuple (φ,μ,σ,v,p) to (1.1)-(1.2) in the sense of Definition 2.1 with

    pL87(0,T;H2),vL87(0,T;H1), (2.13)

    and in addition satisfies

    φL(H1)L2(H3)W1,85((H1))+σL(L2)W1,54((W1,5)))L2(H1)+μL2(H1)+b12σL2(L2(Ω))+pL85(H1)L87(H2)+K12(vL2(L2)L87(H1)+div(φv)L85((H1))+div(σv)L54((W1,5)))C, (2.14)

    where the constant C does not depend on (φ,μ,σ,v,p) and is uniformly bounded for b,χ(0,1] and is also uniformly bounded for K(0,1] when Γv=0.

    The regularity result (2.13) is new compared to estimates for weak solutions in [22], which arises from a deeper study of the Darcy subsystem, and can be obtained even in the absence of the nutrient. We mention that higher regularity estimates for the pressure p in L2(0,T;H2) and the velocity v in L2(0,T;H1) are also established in [22, but these are for strong solutions local in time in three dimensions and global in time for two dimensions.

    We now investigate the situation in two dimensions, where the Sobolev embeddings in two dimensions yields better integrability exponents.

    Theorem 2.2 (Existence of weak solutions in 2D). Let ΩR2 be a bounded domain with C3-boundary Ω. Suppose Assumption 2.1is satisfied. Then, there exists a quintuple (φ,μ,σ,v,p) to (1.1)-(1.2) with the following regularity

    φL(0,T;H1)L2(0,T;H3)W1,w(0,T;(H1)),μL2(0,T;H1),σL2(0,T;H1)L(0,T;L2)W1,r(0,T;(H1)),pLk(0,T;H1L20)Lq(0,T;H2),vL2(0,T;L2)Lq(0,T;H1),

    for

    1k<2,1q<43,1<r<87,43w<2,

    such that (2.12a), (2.12b), (2.12d), (2.12e) and

    tσ,ζH1,(H1)=Ωn(φ)(Dσχφ)ζSζ+σvζ dx+Ωb(σσ)ζ dHd1

    are satisfied for a.e. t(0,T), for all ζH1, and all ζL2. Furthermore, the initial conditions φ(0)=φ0 and σ(0)=σ0 are attained as in Definition 2.1, and an analogous inequality to (2.14) also holds.

    The proof of Theorem 2.2 is similar to that of Theorem 2.1, and hence the details are omitted. In Section 7 we will only present the derivation of a priori estimates. It is due to the better exponents for embeddings in two dimensions and the regularity result for the velocity that we obtain better regularities for the time derivatives tφ and tσ, namely tσ(t) belongs to the dual space (H1) for a.e. t(0,T). Furthermore, as mentioned in Remark 7.1 below, if we only have vL2(0,T;L2), then the convection term div(σv) and the time derivative tσ would only belong to the dual space L43(0,T;(W1,4)). However, even with the improved temporal regularity, as tσL2(0,T;(H1)), we do not have a continuous embedding into the space C0([0,T];L2) and so σ(0) may not be well-defined as an element of L2.

    We now state the two asymptotic limits of (1.1) for three dimensions, and note that analogous asymptotic limits also hold for two dimensions.

    Theorem 2.3 (Limit of vanishing permeability). For b,K(0,1], we denote a weak solution to (1.1)-(1.2) with Γv=0 and initial conditions (φ0,σ0) by (φK,μK,σK,vK,pK). Then, as b0 and K0, it holds that

    φKφ weakly -  in L(0,T;H1)L2(0,T;H3)W1,85(0,T;(H1)),σKσ weakly -  in L2(0,T;H1)L(0,T;L2)W1,54(0,T;(W1,5)),μKμ weakly  in L2(0,T;H1),pKp weakly  in L85(0,T;H1)L87(0,T;H2),vK0 strongly  in L2(0,T;L2)L87(0,T;H1),

    where (φ,μ,σ,p) satisfies

    tφ,ζH1,(H1)=Ωm(φ)μζ+Γφ(φ,μ,σ)ζ dx, (2.16a)
    Ωμζ dx=ΩAΨ(φ)ζ+Bφζχσζ dx, (2.16b)
    tσ,ϕW1,5,(W1,5)=Ωn(φ)(Dσχφ)ϕS(φ,μ,σ)ϕ dx, (2.16c)
    Ωpζ dx=Ω(μ+χσ)φζ dx, (2.16d)

    for all ζH1, ϕW1,5 and a.e. t(0,T). A posteriori, it holds that

    tφ, tσL2(0,T;(H1)),

    and thus φ(0)=φ0 and σ(0)=σ0.

    Theorem 2.4 (Limit of vanishing chemotaxis). For b,χ(0,1], we denote a weak solution to (1.1)-(1.2) with corresponding initial conditions (φ0,σ0) by (φχ,μχ,σχ,vχ,pχ). Then, as b0 and χ0, it holds that

    φχφ weakly -  in L(0,T;H1)L2(0,T;H3)W1,85(0,T;(H1)),σχσ weakly -  in L2(0,T;H1)L(0,T;L2)W1,54(0,T;(W1,5)),μχμ weakly  in L2(0,T;H1),pχp weakly  in L85(0,T;H1)L87(0,T;H2),vχv weakly  in L2(0,T;L2)L87(0,T;H1),

    and

    div(φχvχ)div(φv) weakly  in L85(0,T;(H1)),div(σχvχ)div(σv) weakly  in L54(0,T;(W1,5)),

    where (φ,μ,σ,v,p) satisfies

    tφ,ζH1,(H1)=Ωm(φ)μζ+Γφ(φ,μ,σ)ζ+φvζ dx, (2.19a)
    Ωμζ dx=ΩAΨ(φ)ζ+Bφζ dx, (2.19b)
    tσ,ϕW1,5,(W1,5)=Ωn(φ)DσϕS(φ,μ,σ)ϕ+σ vϕ dx, (2.19c)
    Ωpζ dx=Ω1KΓvζ+μφζ dx, (2.19d)
    Ωv ζ dx=ΩK(pμφ)ζ dx, (2.19e)

    for all ζH1, ϕW1,5, ζL2 and a.e. t(0,T), with the attainment of initial conditions as in Definition 2.1.


    3. Galerkin approximation

    We will employ a Galerkin approximation similar to the one used in [22]. For the approximation, we use the eigenfunctions of the Neumann--Laplacian operator {wi}iN. Recall that the inverse Neumann--Laplacian operator L:=(ΔN)1|L20:L20L20 is compact, positive and symmetric. Indeed, let f,gL20 with z=Lf, y=Lg. Then,

    (Lf,f)L2=Ωzf dx=Ω|z|2 dx0,(Lf,g)L2=Ωzy dx=(f,Lg)L2.

    Furthermore, let {fn}nNL20 denote a sequence with corresponding solution sequence {zn=Lfn}nNH1L20. By elliptic regularity theory, we have that znH2N for all nN. Then, by reflexive weak compactness theorem and Rellich--Kondrachov theorem, there exists a subsequence such that znjzH1L20 as j.

    Thus, by the spectral theorem, the operator L admits a countable set of eigenfunctions {vn}nN that forms a complete orthonormal system in L20. The eigenfunctions of the Neumann--Laplacian operator is then given by w1=1, wi=vi1 for i2, and {wi}iN is a basis of L2.

    Elliptic regularity theory gives that wiH2N and for every gH2N, we obtain for gk:=ki=1(g,wi)L2wi that

    Δgk=ki=1(g,wi)L2Δwi=ki=1(g,λiwi)L2wi=ki=1(g, Δwi)L2wi=ki=1(Δg,wi)L2wi,

    where λi is the corresponding eigenvalue to wi. This shows that Δgk converges strongly to Δg in L2. Making use of elliptic regularity theory again gives that gk converges strongly to g in H2N. Thus the eigenfunction {wi}iN of the Neumann--Laplace operator forms an orthonormal basis of L2 and is also a basis of H2N.

    Later in Section 5, we will need to use the property that H2N is dense in H1 and W1,5. We now sketch the argument for the denseness of H2N in W1,5 and the argument for H1 follows in a similar fashion.

    Lemma 3.1. H2N is dense in W1,5.

    Proof. Take gW1,5, as Ω has a C3-boundary, by standard results [12, Thm. 3, x5.3.3] there exists a sequence gnC(¯Ω) such that gng strongly in W1,5. Let ε>0 be fixed, and define Dε:={xΩ: dist(x,Ω)ε}. Let ζεCc(Ω) be a smooth cut-off function such that ζε=1 in Ω¯Dε and ζε=0 in ¯Dε2.

    As gnC(¯Ω), its trace on Ω is well-defined. Choosing ε sufficiently small allows us to use a classical result from differential geometry about tubular neighbourhoods, i.e., for any zTubε(Ω):={xRd: |dist(z,Ω)|ε} there exists a unique yΩ such that

    z=y+dist(z,Ω)n(y),

    where n is the outer unit normal of Ω. We consider a bounded smooth function fn,ε:RdR such that

    fn,ε(z)=gn(y) for all zTubε(Ω) satisfying z=y+ dist(z,Ω)n(y).

    We now define the smooth function Gn,ε as

    Gn,ε(x):=ζε(x)gn(x)+(1ζε(x))fn,ε(x).

    By construction, the values of the function fn,ε in DεTubε(Ω) are constant in the normal direction, so Gn,εn=0 on Ω and thus Gn,εH2N. Furthermore, we compute that

    Gn,εgnL5=(1ζε)(fn,εgn)L5(Dε),(Gn,εgn)L5=(gnfn,ε)ζε+(1ζε)gn+(1ζε)fn,εL5.

    Using that gn,fn,ε are smooth functions on ¯Ω and that the Lebesgue measure of Dε tends to zero as ε0 we have the strong convergence of Gn,ε to gn in L5. For the difference in the gradients, we use that ζε1 a.e. in Ω, Lebesgue's dominated convergence theorem and the boundedness of gn and fn,ε to deduce that

    (1ζε)gnL5+(1ζε)fn,εL50 as ε0.

    For the remaining term (gnfn,ε)ζεL5 we use that the support of ζε lies in Dε¯Dε2 and for any zDε¯Dε2,

    |fn,ε(z)gn(z)|=|gn(y)gn(y+dist(z,Ω)n(y))|dist(z,Ω)0|gn(y+ξn(y))|dξgnLdist(z,Ω)Cε.

    That is, fn,ε converges uniformly to gn in Dε¯Dε2. Furthermore, using ζεLCε in Dε¯Dε2 and |Dε¯Dε2|Cε we obtain (gnfε,n)ζεL5Cε150 as ε0. This shows that Gn,ε converges strongly to gn in W1,5.

    We denote

    Wk:=span{w1,,wk}

    as the finite dimensional space spanned by the first k basis functions and consider

    φk(t,x)=ki=1αki(t)wi(x),μk(t,x)=ki=1βki(t)wi(x),σk(t,x)=ki=1γki(t)wi(x), (3.1a)

    and the following Galerkin ansatz: For 1jk,

    Ωtφkwj dx=Ωm(φk)μkwj+Γφ,kwj+φkvkwjdx, (3.2a)
    Ωμkwj dx=ΩAΨ(φk)wj+Bφkwjχσkwj dx, (3.2b)
    Ωtσkwjdx=Ωn(φk)(Dσkχφk)wjSkwj+σkvkwjdx+Ωb(σσk)wj dHd1, (3.2c)

    where we define the Galerkin ansatz for the pressure pk and the velocity field vk by

    pk=(ΔN)1(1KΓvdiv((μk+χσk)φk)), (3.3)
    vk=K(pk(μk+χσk)φk), (3.4)

    and we set

    Γφ,k:=Γφ(φk,μk,σk),Sk:=S(φk,μk,σk).

    Note that in (3.3), the properties ΓvL20 and φkn=0 on Ω show that the term inside the bracket belongs to L20 and hence pk is well-defined. Let M and S denote the following mass and stiffness matrices, respectively: For 1i,jk,

    Mij=Ωwiwj dx, Sij:=Ωwiwj dx.

    Thanks to the orthonormality of {wi}iN in L2, we see that M is the identity matrix. It is convenient to define the following matrices with components

    (Ck)ji:=Ωwivkwjdx,(MΩ)ji:=ΩwiwjdH,(Skm)ji:=Ωm(φk)wiwjdx,(Skn)ji:=Ωn(φk)wiwjdx,

    for 1i,jk. Furthermore, we introduce the notation

    Rkφ,j:=ΩΓφ,kwj dx,RkS,j:=ΩSkwj dx,ψkj:=ΩΨ(φk)wj dx,Σkj:=Ωσwj dH,

    for 1i,jk, and denote

    Rkφ:=(Rkφ,1,,Rkφ,k), RkS:=(RkS,1,,RkS,k), ψk:=(ψk1,,ψkk), Σk:=(Σk1,,Σkk),

    as the corresponding vectors, so that we obtain an initial value problem for a system of equations for αk:=(αk1,αkk),βk:=(βk1,,βkk), and γk:=(γk1,,γkk) as follows,

    ddtαk=Skmβk+Rkφ+Ckαk, (3.5a)
    βk=Aψk+BSαkχγk, (3.5b)
    ddtγk=Skn(Dγkχαk)RkS+CkγkbMΩγk+bΣk, (3.5c)
    pk=(ΔN)1(1KΓvdiv((μk+χσk)φk)) (3.5d)
    vk=K(pk(μk+χσk)φk), (3.5e)

    and we supplement (3.5) with the initial conditions

    (αk)j(0)=Ωφ0wj dx,(γk)j(0)=Ωσ0wj dx, (3.6)

    for 1jk, which satisfy

    ki=1(αk)i(0)wiH1Cφ0H1, kj=1(γk)i(0)wiL2σ0L2kN, (3.7)

    for some constant C not depending on k.

    We can substitute (3.5b), (3.5d) and (3.5e) into (3.5a) and (3.5c), and obtain a coupled system of ordinary differential equations for αk and γk, where Skm, Ck and Skn depend on the solutions αk and γk in a non-linear manner. Continuity of m(), n(), Ψ() and the source terms, and the stability of (ΔN)1 under perturbations imply that the right-hand sides of (3.5) depend continuously on (αk,γk). Thus, we can appeal to the theory of ODEs (via the Cauchy--Peano theorem [4, Chap. 1, Thm. 1.2]) to infer that the initial value problem (3.5)-(3.6) has at least one local solution pair (αk,γk) defined on [0,tk] for each kN.

    We may define βk via the relation (3.5b) and hence the Galerkin ansatz φk,μk and σk can be constructed from (3.1). Then, we can define pk and vk via (3.3) and (3.4), respectively. Furthermore, as the basis function wj belongs to H2 for each jN, by the Sobolev embedding H2L, we obtain that div(wiwj)L2 for i,jN and hence the function div((μk+χσk)φk) belongs to L2. Then, by elliptic regularity theory, we find that pk(t)H2NL20 for all t[0,tk]. This in turn implies that

    vk(t){f H1:divf=Γv, fn=0 on Ω} for all t[0,tk]. (3.8)

    Next, we show that the Galerkin ansatz can be extended to the interval [0, T] using a priori estimates.


    4. A priori estimates

    In this section, the positive constants C are independent of k, Γv, K, b and χ, and may change from line to line. We will denote positive constants that are uniformly bounded for b,χ(0,1] and are also uniformly bounded for K(0,1] when Γv=0 by the symbol E.

    We first state the energy identity satisfied by the Galerkin ansatz. Let δij denote the Kronecker delta. Multiplying (3.2a) with βkj, (3.2b) with ddtαkj, (3.2c) with Nk,σ:=Dγkj+χ(δ1jαkj), and then summing the product from j=1 to k lead to

    Ωtφkμkdx=Ωm(φk)|μk|2+Γφ,kμk+φkvkμkdx,Ωμktφkdx=ddtΩAΨ(φk)+B2|φk|2dxΩχσktφkdx,ΩtσkNk,σdx=Ωn(φk)|Nk,σ|2SkNk,σ+σkvkNk,σdx +Ωb(σσk)Nk,σdHd1.

    Here, we used that w1=1 and w1=0. Then, summing the three equations leads to

    ddtΩAΨ(φk)+B2|φk|2+N(φk,σk)dx+Ωm(φk)|μk|2+n(φk)|Nk,σ|2dx+ΩDb|σk|2dHd1=ΩΓφ,kμkSkNk,σ+(φkvkμk+σkvkNk,σ)dx+Ωb(σNk,σσkχ(1φk))dHd1. (4.2)

    Next, multiplying (3.4) with 1Kvk, integrating over Ω and integrating by parts gives

    Ω1K|vk|2dx=Ωpkvk+(μk+χσk)φkvkdx=ΩΓvpk+(μk+χσk)φkvk dx,

    where we used that divvk=Γv and vkn=0 on Ω. Similarly, we see that

    Ω(φkμk+σkNk,σ)vkdx=Ωφkvkμk+σkvk(Dσk+χ(1φk))dx=ΩφkΓvμk+(μk+χσk)vkφkD2vk|σk|2dx=ΩΓv(φkμk+D2|σk|2)+(μk+χσk)φkvk dx.

    In particular, we have

    Ω1K|vk|2 dx=ΩΓv(pkμkφkD2|σk|2)(φkμk+σkNk,σ)vk dx. (4.4)

    Adding (4.3) to (4.2) leads to

    ddtΩAΨ(φk)+B2|φk|2+N(φk,σk)dx+Ωm(φk)|μk|2+n(φk)|Nk,σ|2+1K|vk|2dx+ΩDb|σk|2dHd1=ΩΓφ,kμkSkNk,σ+Γv(pkμkφkD2|σk|2)dx+Ωb(σ(Dσk+χ(1φk))σkχ(1φk))dHd1. (4.3)

    To derive the first a priori estimate for the Galerkin ansatz, it suffices to bring (4.4) into a form where we can apply Gronwall's inequality. We start with estimating the boundary term on the right-hand side of (4.4). By Hölder's inequality and Young's inequality,

    |Ωb(σ(Dσk+χ(1φk))σkχ(1φk))dHd1|b(σL2(Ω)Dσk+χ(1φk)L2(Ω)+χσkL2(Ω)(|Ω|12+φkL2(Ω)))Db2σk2L2(Ω)+b(1+χ2D)φk2L2(Ω)+bC(χ+(1+χ2)σ2L2(Ω)).

    By the trace theorem and the growth condition (2.4), we have

    φ2L2(Ω)C2tr(φ2L2+φ2L2)C2tr(1R1Ψ(φ)L1+φ2L2)+C(R2,|Ω|,Ctr), (4.5)

    where the positive constant Ctr from the trace theorem only depends on Ω, and so

    Ωb(σ(Dσk+χ(1φk))σkχ(1φk))dHd1Db2σk2L2(Ω)+Cb(1+χ2)(Ψ(φk)L1+φk2L2)+Cb(1+χ2)+bC(χ+(1+χ2)σ2L2(Ω)). (4.6)

    4.1. Estimation of the source terms

    For the source term

    ΩΓφ,kμkSkNk,σ+Γv(pkμkφkD2|σk|2) dx

    that appears on the right-hand side of (4.4) we will divide its analysis into two parts. We first analyse the part involving Γv, which will involve a closer look at the Darcy subsystem to deduce an estimate on pkL2. For the remainder Γφ,kμkSkNk,σ term we will estimate it differently based on the assumptions on Θφ.


    4.1.1 Pressure estimates

    Before we estimate the source terms involving Γv, we look at the Darcy subsystem, which can be expressed as an elliptic equation for the pressure (we will drop the subscript k for clarity)

    Δp=1KΓvdiv((μ+χσ)φ) in Ω, with ˉp=0, (4.7a)
    p=0 on Ω. (4.7b)

    The following lemma is similar to [22, Lem. 3.1], and the hypothesis is fulfilled by the Galerkin ansatz.

    Lemma 4.1. Let ΩR3 be a bounded domain with C3-boundary. Given φH2N, μ,σH1, the source term ΓvL20, and the function p satisfying the above elliptic equation (4.7). Then, the following estimate hold

    pL2CKΓvL2+C(μL2+χσL6)φL32+CˉμφL2, (4.8)

    for some positive constant C depending only on Ω.

    Proof. We first recall some properties of the inverse Neumann-Laplacian operator. Suppose for gL20, f=(ΔN)1gH1L20 solves

    Δf=g in Ω, f=0 on Ω. (4.9)

    Then, testing with f and integrating over Ω, applying integration by parts and the Poincaré inequality (1.9) leads to

    (ΔN)1gH1=fH1cfL2CgL2, (4.10)

    for positive constants c and C depending only on Cp. Elliptic regularity theory then gives that fH2N with

    fH2C(fH1+gL2)CgL2, (4.11)

    with a positive constant C depending only on Ω. Returning to the pressure system, we observe from (2.10) and the above that

    pL21K(ΔN)1ΓvL2+(ΔN)1(div((μ+χσ)φ))L2CKΓvL2+(ΔN)1(div((μˉμ+χσ)φ))L2+(ΔN)1(div(ˉμφ))L2, (4.12)

    for some positive constant C depending only on Cp. Note that the third term on the right-hand side can be estimated as

    ˉμ(ΔN)1div(φˉφ)L2=ˉμφˉφL2CpˉμφL2. (4.13)

    We now consider estimating the second term on the right-hand side of (4.12). By assumption μ,σH1 and φH2N, we have that

    (μˉμ+χσ)φL2μˉμ+χσL6φL3, (4.14)

    and so if we consider the function h:=(ΔN)1(div((μˉμ+χσ)φ)), then we obtain that

    Ωhζ dx=Ω(μˉμ+χσ)φζ dxζH1 (4.15)

    must hold, and by (4.14) and the Poincaré inequality (1.9) with zero mean it holds that hH1L20. We now define f:=(ΔN)1(h)H2N, and consider testing with ζ=f in (4.15), leading to

    Ω|h|2 dx=Ωhf dx=Ω(μˉμ+χσ)φf dx.

    Since fH2N, elliptic regularity theory and Hölder's inequality gives

    h2L2(μˉμ+χσ)φL65fL6C(μˉμ+χσ)φL65fH2C(μˉμ+χσ)φL65hL2,

    where the constant C depends on Ω and the constant in (4.11). Thus we obtain

    (ΔN)1(div((μˉμ+χσ)φ))L2C(μˉμ+χσ)φL65C(μˉμL6+χσL6)φL32 (4.16)

    for some constant C depending only on Ω. By the Sobolev embedding H1L6 (with constant CSob that depends only on Ω) and the Poincaré inequality, we find that

    μˉμL6CSobμˉμH1c(CSob,Cp)μL2. (4.17)

    Substituting the above elements into (4.12) yields (4.8).

    Remark 4.1. We choose not to use the estimate

    chL2hL2(μˉμ+χσ)φL2 (4.18)

    obtained from substituting ζ=h in (4.15), where c is a positive constant depending only on Cp, since by (4.14) we require control of φ in the L3(Ω)-norm and this is not available when deriving the first a priori estimate. Thus, we make use of the auxiliary problem f=(ΔN)1(h) to derive another estimate on hL2 that involves controlling φ in the weaker L32(Ω)-norm.

    Next, we state regularity estimates for the pressure and the velocity field. The hypothesis will be fulfilled for the Galerkin ansatz once we derived the a priori estimates in Section 4. Note that in Lemma 4.2 below, we consider a source term ΓvL2(0,T;L20), so that our new regularity results for the pressure and the velocity is also applicable to the setting considered in [22].

    Lemma 4.2. Let φL(0,T;H1)L2(0,T;H2NH3), σL2(0,T;H1), μL2(0,T;H1), the source term ΓvL2(0,T;L20), and the function p satisfying (4.7). Then,

    p85L85(H1)C1φ65L(H1)μ+χσ85L2(H1)φ25L2(H3)+C1KT15Γv85L2(L2), (4.19)

    for some positive constant C1 depending only on Ω, and

    p87L87(H2)C2T37K87Γv87L2(L2)+C2T27p87L85(H1)+C2φ27L(H1)μ+χσ87L2(H1)φ67L2(H3), (4.20)

    for some positive constant C2 depending only on Ω. Moreover, if we have the relation

    v=K(p(μ+χσ)φ),

    then

    Dv87L87(L2)C3Kp87L87(H2)+C3Kμ+χσ87L2(H1)φ67L2(H3)φ27L(H1), (4.21)

    for some positive constant C3 depending only on Ω.

    Proof. From (4.7) we see that p satisfies ˉp = 0 and

    Ωpζ dx=Ω(μ+χσ)φζ+1KΓvζ dxζH1(Ω).

    Testing with ζ=p and applying the Hölder's inequality and the Poincaré inequality (1.9) gives

    pL2(μ+χσ)φL2+CpKΓvL2. (4.22)

    Applying Hölder's inequality and the Sobolev embedding H1L6 yields that

    (μ+χσ)φL2μ+χσL6φL3CSobμ+χσH1φL3. (4.23)

    By the Gagliardo--Nirenberg inequality (1.10) with parameters j=0, p=3, r=2, m=2, d=3 and q=2,

    φL3Cφ14H2φ34L2Cφ14H3φ34H1, (4.24)

    where C>0 is a constant depending only on Ω. Then, the boundedness of μ,σ in L2(0,T;H1) and φ in L2(0,T;H3)L(0,T;H1) leads to

    T0(μ+χσ)φ85L2dtCT0μ+χσ85H1φ25H3φ65H1dtCφ65L(H1)μ+χσ85L2(H1)φ25L2(H3).

    By (4.22) we find that

    T0p85L2dtT0(μ+χσ)φ85L2+CKΓv85L2dtC[φ]65L(H1)μ+χσ85L2(H1)φ25L2(H3)+CKT15Γv85L2(L2),

    where the positive constant C depends only on Ω. As ˉp = 0, by the Poincarö inequality (1.9), we see that

    p85L85(H1)Cφ65L(H1)μ+χσ85L2(H1)φ25L2(H3)+CKT15Γv85L2(L2),

    for some positive constant C depending only on Ω. Next, we see that

    div((μ+χσ)φ)L2(μ+χσ)ΔφL2+(μ+χσ)φL2μ+χσL6ΔφL3+(μ+χσ)L2φL.

    By the Gagliardo-Nirenberg inequality (1.10), we find that

    D2φL3Cφ34H3φ14L6Cφ34H3φ14H1,φLCφ34H3φ14L6Cφ34H3φ14H1, (4.25)

    and so, we have

    div((μ+χσ)φ)L2Cμ+χσH1φ34H3φ14H1. (4.26)

    That is, div((μ+χσ)φ)L2. Since by assumption ΓvL20, using elliptic regularity theory, we find that p(t,)H2 for a.e. t and there exists a constant C depending only on Ω, such that

    pH2C(pH1+div((μ+χσ)φ)L2+K1ΓvL2). (4.27)

    Furthermore, from (4.26), we see that

    T0div((μ+χσ)φ)87L2dtCφ27L(H1)T0μ+χσ87H1φ67H3dtCφ27L(H1)μ+χσ87L2(H1)φ67L2(H3),

    and so for some positive constant C depending only on Ω, it holds that

    T0p87H2dtCT37K87Γv87L2(L2)+CT27p87L85(H1)+Cφ27L(H1)μ+χσ87L2(H1)φ67L2(H3). (4.28)

    For the velocity field v we estimate as follows. Let 1i,j3 be fixed, we obtain from (4.25),

    DivjL2=KDiDjpDi(μ+χσ)Djφ(μ+χσ)DiDjφL2K(pH2+(μ+χσ)L2φL+μ+χσL6D2φL3)K(pH2+Cμ+χσH1φ34H3φ14H1). (4.29)

    Applying the same calculation as in (4.28) yields

    T0Dv87L2dtCKT0p87H2+μ+χσ87H1φ67H3φ27H1dtCK(p87L87(H2)+μ+χσ87L2(H1)φ67L2(H3)φ27L(H1)),

    for some positive constant C depending only on Ω.


    4.1.2 Source term from the Darcy system

    To estimate the third source term

    ΩΓv(pkμkφkD2|σk|2) dx=ΩΓv(pkˉμkφk+(μkˉμk)φkD2|σk|2) dx

    of the energy equality we use Hölder’s inequality to obtain

    |ΩΓvD2|σk|2+Γvφk(μkˉμk)dx|D2ΓvL2σk2L4+ΓvL32μkˉμkL6φkL6.

    By the Gagliardo--Nirenberg inequality (1.10) with j=0, r=2, m=1, p=4, q=2 and α=34, we have

    σk2L4Cσk32H1σk12L2=C(σk2L2+σk12L2σk32L2).

    By Young's inequality with Hölder exponents (i.e., abεpap+εq/pqbq for 1p+1q=1 and ε>0), we find that

    D2ΓvL2σk2L4C(ΓvL2σk2L2+Γv4L2σk2L2)+n0D24σk2L2,

    for some positive constant C depending only in n0, D and Ω. Then, by (4.17) we have

    |ΩΓvD2|σk|2+Γvφk(μk¯μk)dx|n0D24σk2L2+C(1+Γv4L2)σk2L2+ΓvL32c(Cp,CSob)μkL2φkH1n0D24σk2L2+C(1+Γv4L2)σk2L2+m08μk2L2+CΓv2L2φk2H1,

    where the positive constant C depends only on Ω, m0, n0 and D. Here we point out that the assumption ΓvL4(0,T;L20) is needed. For the remainder term Γv(pk¯μkφk), we find that

    pk¯μkφk=((ΔN)1(1KΓvdiv((μk¯μk+χσk)φk)¯μkdiv(φk¯φk)))¯μkφk=((ΔN)1(1KΓvdiv((μk¯μk+χσk)φk)))¯μk¯φk,

    where we used

    (ΔN)1(¯μkdiv(φk¯φk))=¯μk(φk¯φk).

    Then, by

    ΩΓv¯φk¯μk dx=¯μk¯φkΩΓv dx=0,

    it holds that

    ΩΓv(pk¯μkφk) dx=ΩΓv((ΔN)1(1KΓvdiv((μk¯μk+χσk)φk))).

    Applying the calculations in the proof of Lemma 4.1 (specifically (4.10), (4.16) and (4.17)), H¨older’s inequality and Young’s inequality, we find that

    |ΩΓv(pk¯μkφk)dx|CKΓv2L2+CΓvL2(μkL2+χσkH1)φkL32CKΓv2L2+m08μk2L2+n0D24σk2L2+σk2L2+C(1+χ2)Γv2L2φk2L2,

    where C is a positive constant depending only on |Ω|, Cp, CSob, D, n0 and m0. Here we point out that if we applied (4.18) instead of (4.8) then we obtain a term containing φL3 on the right-hand side and this cannot be controlled by the left-hand side of (4.4). Using (2.4) we have

    φk2L21R1Ψ(φk)L1+R2R1|Ω|. (4.30)

    Then, we obtain the following estimate

    |ΩΓv(pkμkφkD2|σk|2)dx|C(1KΓv2L2+1)+n0D22σk2L2+m04μk2L2+C(1+Γv4L2)σk2L2+CΨ(φk)L1+C(1+χ2)Γv2L2φk2L2, (4.31)

    for some positive constant C depending only on R1, R2, Ω, m0, n0 and D. Here we point out that it is crucial for the source term Γv to be prescribed and is not a function of φ, μ and σ, otherwise the product term Γv4L2σk2L2 and Γv2L2φk2L2 cannot be controlled in the absence of any a priori estimates. For the remaining source term

    ΩΓφ,kμkSkNk,σ dx

    we split the analysis into two cases and combine with (4.31) to derive an energy inequality.


    4.1.3 Energy inequality for non-negative Θφ

    Suppose Θφ is non-negative and bounded, and Ψ is a potential that satisfies (2.5). We will estimate the mean of μk by setting j=1 in (3.2b), and using the growth condition (2.5) to obtain

    |Ωμkdx|2=|ΩAΨ(φk)χσkdx|22A2Ψ(φk)2L1+2χ2σk2L12A2R24(|Ω|+|Ω|12φkL2)2+2χ2|Ω|σk2L2C(A,R4,|Ω|)+4A2R24|Ω|σk2L2+2χ2|Ω|σk2L2.

    Then, by the Poincarö inequality (1.9) and the growth condition (2.4), we find that

    μk2L22C2Pμk2L2+2|Ω||¯μk|22C2pμk2L2+8A2R24φk2L2+4χ2σk2L2+C(A,R4,|Ω|)2C2pμk2L2+8A2R24R1Ψ(φk)L1+4χ2σk2L2+C(A,R4,R1,R2,|Ω|). (4.32)

    Note that by the specific form (2.1) for Γφ we have that

    Γφ,kμk=Λφ(φk,σk)μkΘφ(φk,σk) |μk|2.

    Moving the non-negative term Θφ(φk,σk)|μk|2 to the left-hand side of (4.4) and subsequently neglecting it, we estimate the remainder using the growth condition (2.3) and Hölder's inequality as follows (here we use the notation Λφ,k:=Λφ(φk,σk)),

    |ΩΛφ,kμkSk(Dσk+χ(1φk))dx|Λφ,kL2μkL2+(ΛS,kL2+R0μkL2)Dσk+χ(1φk)L2C(1+χ+(1+χ)φkL2+(1+D)σkL2)μkL2+C(1+φkL2+σL2)(χ|Ω|12+DσL2+χφkL2) (4.33)

    where C is a positive constant depending only on R0 and |Ω|. By Young's inequality, (4.32) and (4.30), we have

    |ΩΛφ,kμkSk(Dσk+χ(1φk))dx|m08C2pμk2L2+C(1+χ+D+χ2)φk2L2+C(1+χ+D)2σk2L2+C(1+χ+χ2)m04μk2L2+C(1+χ2)σk2L2+C(1+χ2)Ψ(φk)L1+C(1+χ2), (4.34)

    for some positive constant C depending only on |Ω|, R0, R1, R2, R4, A, D, Cp and m0. Using the fact that

    DσL2(Dσ+χ(1φ))L2+χφL2,

    we now estimate the right-hand side of (4.4) using (4.6), (4.31) and (4.34), which leads to

    ddtΩAΨ(φk)+B2|φk|2+D2|σk|2+χσk(1φk)dx+m02μk2L2+n0D22σk2L2+1Kvk2L2+Db2σk2L2(Ω)C(1+b)(1+χ2)Ψ(φk)L1+C(Γv2L2+b)(1+χ2)φk2L2+C(1+χ2+Γv4L2)σk2L2+C(1+b)(1+χ2)+CKΓv2L2+bC(1+χ2)σ2L2(Ω), (4.35)

    for some positive constant C not depending on Γv, K, b and χ. Integrating (4.35) with respect to t from 0 to s(0,T] leads to

    AΨ(φk(s))L1+B2φk(s)2L2+D2σk(s)2L2+Ωχσk(s)(1φk(s))dx+s0m02μk2L2+n0D22σk2L2+1Kvk2L2+Db2σk2L2(Ω)dts0C(1+b)(1+χ2)(1+Γv4L2)(Ψ(φk)L1+φk2L2+σk2L2)dt+C(1+b)(1+χ2)T+CKΓv2L2(0,T;L2)+Cb(1+χ2)σ2L2(0,T;L2(Ω))+C(Ψ(φ0)L1+φ02H1+σ02L2), (4.36)

    for some positive constant C independent of Γv, K, χ and b. Here we used σ0L2 and φ0H1, which implies by the growth condition (2.5) that Ψ(φ0)L1. Next, by Hölder's inequality and Young's inequality we have

    |Ωχσk(x,s)(1φk(x,s))dx|2D8σk(s)2L2+2χ2|Ω|D+2χ2Dφk(s)2L2D4σk(s)2L2+2χ2DR1Ψ(φk(s))L1+2χ2|Ω|D(1+R2). (4.37)

    Substituting (4.37) into (4.36) then yields

    min(A2χ2DR1,B2,D4)(Ψ(φk(s))L1(Ω)+φk(s)2L2(Ω)+σk(s)2L2(Ω))+s0μk2L2+σk2L2+1Kvk2L2+b2σk2L2(Ω)dts0C(1+b)(1+χ2)(1+Γv4L2)(Ψ(φk)L1+φk2L2+σk2L2)dt+C(1+b)(1+χ2)(1+T)+CKΓv2L2(0,T;L2), (4.38)

    for some positive constant C independent of Γv, K, b and χ. Setting

    α:=C(1+b)(1+χ2)(1+T)+CKΓv2L2(0,T;L2),β:=C(1+b)(1+χ2)(1+Γv4L2)L1(0,T), (4.39)

    and noting that

    α(1+s0β(t)exp(t0β(r) dr) dt)α(1+βL1(0,T)exp(βL1(0,T)))<,

    we find after applying the Gronwall inequality (1.11) to (4.38) leads to

    sups(0,T](Ψ(φk(s))L1+φk(s)2L2+σk(s)2L2)+T0μk2L2+σk2L2+1Kvk2L2+b2σk2L2(Ω)dtE, (4.40)

    where we recall that E denotes a constant that is uniformly bounded for b,χ(0,1] and is also uniformly bounded for K(0,1] when Γv=0.


    4.1.4 Energy inequality for positive Θφ

    Suppose Θφ satisfies (2.6) and Ψ is a potential satisfying the growth condition (2.7). Similar to the previous case, we see that the specific form for Γφ leads to

    Γφ,kμk=Λφ(φk,σk)μkΘφ(φk,σk) |μk|2.

    We move the term Θφ(φk,σk)|μk|2 to the left-hand side of (4.4) and estimate the remainder as in (4.33). Using Young's inequality differently and also (4.30), we have

    |ΩΛφ,kμkSk(Dσk+χ(1φk))dx|R52μk2L2+C(1+χ+D+χ2)φk2L2+C(1+χ+D)2σk2L2+C(1+χ+χ2)R52μk2L2+C(1+χ2)σk2L2+C(1+χ2)Ψ(φk)L1+C(1+χ2), (4.41)

    for some positive constant C depending only on |Ω|, R5, R1, R2, A, D, and Cp. Using (4.6), (4.31), (4.41) and the lower bound ΘφR5, instead of (4.35) we obtain from (4.4)

    ddtΩAΨ(φk)+B2|φk|2+D2|σ|k2+χσk(1φk)dx+R52μk2L2+m02μk2L2(Ω)+n0D22σk2L2+1Kvk2L2+Db2σk2L2(Ω)C(1+b)(1+χ2)Ψ(φk)L1+C(Γv2L2+b)(1+χ2)φk2L2+C(1+χ2+Γv4L2)σk2L2+C(1+b)(1+χ2)+CKΓv2L2+Cb(1+χ2)σ2L2(Ω), (4.42)

    for some positive constant C independent of Γv, K, b and χ. We point out the main difference between (4.35) and the above is the appearance of the term R52μk2L2 on the left-hand side. The positivity of Θφ allows us to absorb the μk2L2 term on the right-hand side of (4.41) and thus we do not need to use (4.32), which was the main reason why Ψ has to be a quadratic potential for a non-negative Θφ. Then, applying a similar argument as in Section 4.1.3, we arrive at an analogous energy inequality to (4.40),

    sups(0,T](Ψ(φk(s))L1+φk(s)2L2+σk(s)2L2)+T0μk2H1+σk2L2+1Kvk2L2+b2σk2L2(Ω)dtE. (4.43)

    Using (4.32) and (4.30) applied to (4.40), and similarly using (4.30) applied to (4.43) we obtain

    sups(0,T](Ψ(φk(s))L1+φk(s)2H1+σk(s)2L2)+T0μk2H1+σk2L2+1Kvk2L2+b2σk2L2(Ω)dtE. (4.44)

    This a priori estimate implies that the Galerkin ansatz φk, μk, σk and vk can be extended to the interval [0,T]. To determine if pk can also be extended to the interval [0,T] we require some higher order estimates for φk in order to use (4.19).


    4.2. Higher order estimates

    Let Πk denote the orthogonal projection onto the finite-dimensional subspace Wk. From (3.2b) we may view φk as the solution to the following elliptic equation

    BΔu+u=μkAΠk(Ψ(u))+χσk+u in Ω, (4.45a)
    nu=0 on Ω. (4.45b)

    For the case where Ψ satisfies (2.5), as {φk}kN is bounded in L(0,T;H1), we have that {Ψ(φk)}kN is also bounded in L(0,T;H1). Using the fact that our basis functions {wi}iN are the eigenfunctions of the inverse Neumann-Laplacian operator and is therefore orthogonal in H1, and the Sobolev embedding H1Lr for r[1,6], there exists a positive constant C independent of φk such that

    Πk(Ψ(φk))XCΨ(φk)X for X=H1 or Lr,1r6. (4.46)

    Then, this implies that {Πk(Ψ(φk))}kN is also bounded in L(0,T;H1). As the right-hand side of (4.45a) belongs to H1 for a.e. t(0,T), and the boundary Ω is C3, by elliptic regularity theory, we have

    φkL2(H3)C(1+φkL2(H1)+μk+χσkL2(H1))E, (4.47)

    for some positive constant C depending only on Ω and R4. For the case where Ψ satisfies (2.7), we employ a bootstrap argument from [18, x3.3]. The growth assumption (2.7) implies that

    |Ψ(y)|C(1+|y|m), |Ψ(y)|C(1+|y|m1) for m[1,5). (4.48)

    For fixed m[1,5), we define a sequence of positive numbers {lj}jN by

    l1m6,lj+1=6lj6(5m)lj.

    It can be shown that {lj}jN is a strictly increasing sequence such that lj as j. The Gagliardo--Nirenberg inequality (1.10) then yields the following continuous embedding

    L2(0,T;W2,lj)L(0,T;L6)L2m(0,T;Lmlj+1). (4.49)

    At the first step, the boundedness of {φk}kN in L(0,T;H1) yields

    Πk(Ψ(φk))2Ll1C(1+φk2mL6),

    which implies that {Πk(Ψ(φk))}kN is bounded in L2(0,T;Ll1). As the other terms on the right-hand side of (4.45) are bounded in L2(0,T;H1), elliptic regularity then yields that {φk}kN is bounded in L2(0,T;W2,l1), and thus in L2m(0,T;Lml2) by (4.49).

    At the j-th step, we have {φk}kN is bounded in L2(0,T,W2,lj)L2m(0,T;Lmlj+1). Then, it holds that

    Πk(Ψ(φk))2L2(Llj+1)C(1+φk2mL2m(Lmlj+1)),

    and so {Πk(Ψ(φk))}kN is bounded in L2(0,T;Llj+1). Elliptic regularity then implies that {φk}kN is bounded in L2(0,T;W2,lj+1).

    We terminate the bootstrapping procedure once lj6 for some jN. This occurs after a finite number of steps as limjlj=. Altogether, we obtain that {φk}kN is bounded in L2(0,T;W2,6). From (4.48) it holds that

    Ψ(φk)φk2C(1+|φk|2m2)|φk|2 for m[1,5),

    and by the following continuous embeddings obtain from the Gagliardo-Nirenberg inequality (1.10),

    L2(0,T;W2,6)L(0,T;H1)L2m(0,T;L2m)L2m2(0,T;L) for m[1,5),

    we find that {Πk(Ψ(φk))}kN is bounded in L2(0,T;H1). Applying elliptic regularity once more leads to the boundedness of {φk}kN in L2(0,T;H3). Consequently, the hypotheses of Lemma 4.2 are satisfied and we obtain that

    pkL85(H1)E,

    which implies that the Galerkin ansatz pk can be extended to the interval [0,T].


    4.3. Estimates for the convection terms and the time derivatives

    By the Gagliardo-Nirenberg inequality (1.10) with j=0, p=, m=3, r=2, q=2 and d=3, we have

    φkLCφk14H3φk34L6Cφk14H3φk34H1.

    For any ζL83(0,T;H1) with coefficients {ζkj}1jkRk such that Πkζ=kj=1ζkjwj, we can estimate

    |T0ΩφkvkΠkζdx dt|T0vkL2φkLΠkζL2dtCφk34L(H1)vkL2(L2)φk14L2(H3)ζL83(H1). (4.50)

    Using (4.44) and (4.47), we find that

    div(φkvk)L85((H1))K12E. (4.51)

    Next, multiplying (3.2a) by ζkj, summing from j=1 to k and then integrating in time from 0 to T leads to

    |T0Ωtφkζdx dt|T0m1μkL2ΠkζL2dt+T0Γφ,kL2ΠkζL2+div(φkvk)(H1)ΠkζH1dt.

    By (2.1), (2.2) and (4.44), we find that

    Γφ,kL2(L2)C(R0,|Ω|,T)(1+φkL2(L2)+μkL2(L2)+σkL2(L2))E,

    and so, by Hölder’s inequality, we find that

    |T0Ωtφkζ dx dt|(ET18+div(φkvk)L85((H1)))ζL83(H1).

    Taking the supremum over ζL83(0,T;H1) and using (4.44) and (4.51) yields that

    tφkL85((H1))E(1+K12), (4.52)

    Similarly, by Hölder's inequality and the following Gagliardo--Nirenberg inequality (1.10) with j=0, r=2, m=1, p=103, q=2 and d=3,

    fL103Cf35H1f25L2,

    which in turn implies that {σk}kN is bounded uniformly in L103(Q). Then, we find that for any ζL5(0,T;W1,5),

    |T0ΩσkvkΠkζdx dt|T0σkL103vkL2ζL5dtσkL103(Q)vkL2(L2)ζL5(L5), (4.53)

    and

    div(σkvk)L54((W1,5))K12E. (4.54)

    A similar calculation to (4.52) yields that

    tσkL54((W1,5))E(1+K12). (4.55)

    Remark 4.2. We may also use the Gagliardo-Nirenberg inequality to deduce that

    fLrCf3(r2)2rH1f6r2rL2 for any r(2,6).

    Then, the computation (4.53) becomes

    |T0ΩσkvkΠkζ dx dt|CvkL2(L2)σk6r2rL(L2)σk3(r2)2rL2(H1)ζL4r6r(L2rr2),

    which implies that {div(σkvk)}kN and {tσk}kN are bounded uniformly in

    L4r5r6(0,T;(W1,2rr2)) for r(2,6).

    Note that the temporal exponent decreases while the spatial exponent increases as r increases, and they intersect at the point r=103

    Here we point out that even with the improved regularity vkL87(0,T;H1), we are unable to show div(σkvk) is bounded in the dual space (H1). Indeed, let q, r>1 be constants yet to be determined such that 1q+1r=12. Then, from Hölder's inequality we have

    |T0ΩσkvkΠkζ dx dt|T0σkLqvkLrζL2 dt.

    By the Gagliardo--Nirenberg inequality we have for α=323q1, β=323r1,

    |T0ΩσkvkΠkζdx dt|CT0σk1αL2σkαH1vkβH1vk1βL2ζL2dtCσk1αL(L2)T0σkαH1vkβH1vk1βL2ζH1dtCσk1αL(L2)σkαLαx1(H1)vkβLβx2(H1)vk1βL(1β)x3(L2)ζLx4(H1),

    where

    1x1+1x2+1x3+1x4=1,αx12,βx287,(1β)x32. (4.56)

    Note that α=323q=3r, and then substituting into the constraints (4.56) we find that

    1x1+1x2+1x3α2+78β+1β2=32r+2116218r+32r14=1716+38r>1. (4.57)

    Hence, we cannot find x1, x2, x3 and x4 satisfying (4.56) and we are unable to deduce that div(σkvk) lies in the dual space (H1) even with the improved regularity vkL87(0,T;H1).


    5. Passing to the limit

    From (4.44), (4.47), (4.19), (4.20), (4.21), (4.51), (4.52), (4.54), (4.55), we find that

    {φk}kNbounded in L(0,T;H1)L2(0,T;H3),{tφk}kN,{div(φkvk)}kN bounded in L85(0,T;(H1)),{σk}kN bounded in L(0,T;L2)L2(0,T;H1)L2(Σ),{tσk}kN,{div(σkvk)}kN bounded in L54(0,T;(W1,5)),{μk}kN bounded in L2(0,T;H1),{pk}kN bounded in L85(0,T;H1)L87(0,T;H2),{vk}kN bounded in L2(0,T;L2)L87(0,T;H1).

    By standard compactness results (Banach-Alaoglu theorem and reflexive weak compactness theorem), and [28, x8, Cor. 4], and the compact embeddings in dimension 3 (see [1, Thm. 6.3] and [15, Thm. 11.2, p. 31])

    Hj+1(Ω)=Wj+1,2(Ω)⊂⊂Wj,q(Ω)j0,jZ,1q<6,

    and the compact embedding L2⊂⊂(H1), we obtain, for a relabelled subsequence, the following weak/weak-* convergences:

    φkφ weakly -  in L(0,T;H1)L2(0,T;H3)W1,85(0,T;(H1)),σkσ weakly -  in L2(0,T;H1)L(0,T;L2)L2(Σ),tσktσ weakly  in L54(0,T;(W1,5)),μkμ weakly  in L2(0,T;H1),pkp weakly  in L85(0,T;H1)L87(0,T;H2),vkv weakly  in L2(0,T;L2)L87(0,T;H1),div(φkvk)ξ weakly  in L85(0,T;(H1)),div(σkvk)θ weakly  in L54(0,T;(W1,5)),

    and the following strong convergences:

    φkφ strongly  in C0([0,T];Lr)L2(0,T;W2,r) and a.e. in Q,σkσ strongly  in C0([0,T];(H1))L2(0,T;Lr) and a.e. in Q,

    for any r[1,6) and some functions ξL85(0,T;(H1)), θL54(0,T;(W1,5)).

    For the rest of this section, we fix 1jk and δCc(0,T). Then, we have δ(t)wjC(0,T;H2). By continuity of m(), we see that m(φk)m(φ) a.e. in Q. Thanks to the boundedness of m(), applying Lebesgue's dominated convergence theorem to (m(φk)m(φ))2|δwj|2 yields

    m(φk)δwjm(φ)δwjL2(Q)0 as k.

    Together with the weak convergence μkμ in L2(0,T;H1), we obtain by the product of weak-strong convergence

    T0Ωm(φk)δwjμk dx dtT0Ωm(φ)δwjμ dx dt as k.

    Terms involving n() can be dealt with in a similar fashion. For the source term Γφ,k=Λφ(φk,σk)Θφ(φk,σk)μk, by the continuity and boundedness of Θφ, the a.e. convergence of φkφ and σkσ in Q, we may apply Lebesgue's dominated convergence theorem to deduce that

    T0Ω|δwj(Θφ(φk,σk)Θφ(φ,σ))|2 dx dt0 as k,

    that is, we obtain the strong convergence δwjΘφ(φk,σk)δwjΘφ(φ,σ) in L2(Q). Hence, the weak convergence μkμ in L2(0,T;H1) yields

    T0ΩδwjΘφ(φk,σk)μk dx dtT0ΩδwjΘφ(φ,σ)μ dx dt as k.

    Meanwhile, by the triangle inequality ||a||b|||ab|, and Hölder's inequality, we obtain

    T0Ω|(|φk||φ|)(δwj)| dx dtφkφL2(0,T;L2)δwjL2(0,T;L2)0

    and

    T0Ω|(|σk||σ|)(δwj)| dx dtσkσL2(0,T;L2)δwjL2(0,T;L2)0

    as k. In particular, we have

    (1+|φk|+|σk|)|δwj|(1+|φ|+|σ|)|δwj| strongly in L1(Q) as k.

    By the continuity of Λφ we have

    Λφ(φk,σk)Λφ(φ,σ) a.e. as k,|Λφ(φk,σk)δwj|R0(1+|φk|+|σk|)|δwj|.

    which leads to

    Then, the generalised Lebesgue dominated convergence theorem (see [27, Thm. 1.9, p. 89], or [2, Thm. 3.25, p. 60]) yields

    Λφ(φk,σk)δwjΛφ(φ,σ)δwj strongly in L1(Q) as k,

    which leads to

    T0ΩΓφ(φk,μk,σk)δwj dx dtT0ΩΓφ(φ,μ,σ)δwj dx dt as k. (5.1)

    The same arguments can be applied for the source term S and for the derivative Ψ(φ) satisfying the linear growth condition (2.5). For potentials satisfying the growth condition (2.7), we refer to the argument in [18, x3.1.2].

    To identify the limits ξ and θ of the convection terms div(φkvk) and div(σkvk), respectively, we argue as follows. Since δwjC(0,T;H2)L83(0,T;H1), by the weak convergence div(φkvk)ξ in L85(0,T;(H1)), we have

    T0Ωdiv(φkvk)δwj dx dtT0ξ,wjH1,(H1)δ dt as k.

    Next, applying integrating by parts and by the boundary conditions vkn=0 on Ω (see (3.8)), we see that

    T0Ωdiv(φkvk)δwj dx dt=T0Ωδφkvkwj dx dt. (5.2)

    Moreover, we claim that δφkwj converges strongly to δφwj in L2(0,T;L2). Indeed, we compute

    T0Ω|δ|2|wj|2|φkφ|2dx dtT0|δ|2wj2L6φkφ2L3dtwj2H2δ2L(0,T)φkφ2L2(L3)0

    as k by the strong convergence φkφ in L2(0,T;Lr) for r[1,6). Together with the weak convergence vkv in L2(0,T;L2), when passing to the limit k in (5.2) we find that

    T0ξ,wjH1,(H1)δ dt=T0Ωδφvwj dx dt.

    Applying integration by parts on the right-hand side shows that ξ=div(φv) in the sense of distributions.

    Now considering δ(t)wj as an element in L5(0,T;W1,5), a similar argument can be used to show θ=div(σv) in the sense of distributions using the strong convergence σkσ in L2(0,T;Lr) for r[1,6), the weak convergence vkv in L2(0,T;L2), and the weak convergence div(σkvk)ϕ in L54(0,T;(W1,5)).

    For the pressure and the velocity, we apply ΔN to both sides of (3.3) and test with wj, then integrating by parts leads to

    Ωpkwj dx=Ω1KΓvwj+(μk+χσk)φkwj dx.

    Multiplying by δ(t), integrating in time and passing to the limit k, keeping in mind the weak convergences pkp in L85(0,T;H1), μkμ in L2(0,T;H1), σkσ in L2(0,T;H1), and the strong convergence φkφ in L2(0,T;W2,r) for r[1,6) leads to

    T0Ωδ(t)pwj dx dt=T0Ωδ(t)(1KΓvwj+(μ+χσ)φwj)dx dt. (5.3)

    Here we used that wjH2, and

    T0Ω|δ|2|φkφ|2|wj|2dx dtδ2L(0,T)wj2W1,6φkφ2L2(W1,3)0 as k, (5.4)

    to deduce that δφkwjδφwj in L2(0,T;L2). Fix 1j1,j2,j3k, and define ζ=(wj1,wj2,wj3). Then, we can consider δ(t)ζ as an element in L83(0,T;L2)L2(0,T;L2). Taking the scalar product of (3.4) with δζ, integrating over Ω and in time from 0 to T leads to

    T0Ωδ(t)(vk+Kpk)ζ dx dt=T0ΩδK(μk+χσk)φkζ dx dt. (5.5)

    By the weak convergences vkv in L2(0,T;L2), μkμ in L2(0,T;H1), σkσ in L2(0,T;H1), pkp in L85(0,T;L2), and the strong convergence δφkζδφζ in L2(0,T;L2) (which is proved in a similar manner as (5.4)), we find that passing to the limit in (5.5) yields

    T0Ωδ(t)(v+Kp)ζ dx dt=T0Ωδ(t)K(μ+χσ)φζ dx dt. (5.6)

    Then, multiplying (3.2) with δCc(0,T), integrating with respect to time from 0 to T, and passing to the limit k, we obtain

    T0δ(t)tφ,wjH1,(H1)dt=T0Ωδ(t)(m(φ)μwj+Γφwj+φvwj)dx dt,T0Ωδ(t)μwjdx dt=T0Ωδ(t)(AΨ(φ)wj+Bφwjχσwj)dx dt,T0δ(t)tσ,wjW1,5,(W1,5)dt=T0Ωδ(t)(n(φ)(Dσχφ)wjSwj)dx dt+T0δ(t)(Ωσvwj dx+Γb(σσ)wj dHd1)dt.

    Since the above, (5.3) and (5.6) hold for all δCc(0,T), we infer that {φ,μ,σ,p,v} satisfies (2.12) with ζ=ϕ=wj for a.e. t(0,T) and for all j1. As {wj}jN is a basis for H2N, and H2N is dense in both H1 and W1,5 (see Section 3), we see that {φ,μ,σ,p,v} satisfy (2.12a), (2.12b), (2.12d) for all ζH1, (2.12c) for all ϕW1,5, and (2.12e) for all ζL2.

    Attainment of initial conditions. It remains to show that φ and σ attain their corresponding initial conditions. Strong convergence of φk to φ in C0([0,T];L2), and the fact that φk(0)φ0 in L2 imply that φ(0)=φ0. Meanwhile, as the limit function σ belongs to the function space C0([0,T];(H1)), we see that σ(0):=σ(,0) makes sense as an element of (H1). Let ζH1 be arbitrary, then by the strong convergence σkσ in C0([0,T];(H1)) we see that

    σk(0),ζH1,(H1)σ(0),ζH1,(H1) as k.

    On the other hand, by (3.7), we have σk(0)σ0 in L2. This yields

    σ0,ζH1,(H1)=limkσk(0),ζH1,(H1)=σ(0),ζH1,(H1).

    Energy inequality. For the energy inequality (2.14) we employ the weak/weak-* lower semicontinuity of the norms and dual norms to (4.44), (4.47), (4.19), (4.20), (4.21), (4.51), (4.52), (4.54), and (4.55).


    6. Asymptotic limits


    6.1. Limit of vanishing permeability

    For K,b(0,1] let (φK,μK,σK,vK,pK) denote a weak solution to (1.1)-(1.2) with Γv=0, obtain from Theorem 2.1. By we deduce that, for a relabelled subsequence as b0 and K0, the following weak/weak-* convergences:

    φKφ weakly -  in L(0,T;H1)L2(0,T;H3)W1,85(0,T;(H1)),σKσ weakly -  in L2(0,T;H1)L(0,T;L2)W1,54(0,T;(W1,5)),μKμ weakly  in L2(0,T;H1),pKp weakly  in L85(0,T;H1)L87(0,T;H2),

    and the following strong convergences:

    φKφ strongly  in C0([0,T];Lr)L2(0,T;W2,r) and a.e. in Q,σKσ strongly  in C0([0,T];(H1))L2(0,T;Lr) and a.e. in Q,vK0 strongly  in L2(0,T;L2)L87(0,T;H1),div(φKvK)0 strongly  in L85(0,T;(H1)),div(σKvK)0 strongly  in L54(0,T;(W1,5)),

    for any r[1,6). The strong convergence of the velocity and the convection terms to zero follows from (2.14). Upon multiplying (2.12) by δCc(0,T) and passing to the limit b,K0, we obtain that the limit functions (φ,μ,σ,p) satisfy

    tφ,ζH1,(H1)=Ωm(φ)μζ+Γφ(φ,μ,σ)ζ dx, (6.1a)
    Ωμζ dx=ΩAΨ(φ)ζ+Bφζχσζ dx, (6.1b)
    tσ,ϕW1,5,(W1,5)=Ωn(φ)(Dσχφ)ϕS(φ,μ,σ)ϕ dx (6.1c)
    Ωpζ dx=Ω(μ+χσ)φζ dx, (6.1d)

    for all ζH1 and ϕW1,5 and a.e. t(0,T).

    Note that substituting any ζL2(0,T;H1) into (6.1a), integrating in time from 0 to T, using Hölder's inequality and the linear growth condition for Γφ leads to the deduction that tφL2(0,T;(H1)). To show that tσL2(0,T;(H1)) we argue as follows. For any ξL2(0,T;H1), we can define

    F(ξ):=T0Ωn(φ)(Dσχφ)ξS(φ,μ,σ)ξ dx dt.

    By Hölder's inequality and the growth condition on S, we see that FL2(0,T;(H1)). It is known that the set of functions that are finite linear combinations of C1c(0,T)H2N(Ω):={δ(t)ϕ(x):δC1c(0,T),ϕH2N(Ω)} is dense in C1c(0,T;H1) (see for instance [26, p. 384], and in fact this is what we use in Section 5). Let ζC1c(0,T;H1) and let {ζn}nN denote a sequence of functions of the above form such that ζnζ in C1c(0,T;H1) as n. Then, substituting ϕ=ζn in (6.1c), integrating over t from 0 to T, and passing to the limit n yields

    limnT0tσ,ζnW1,5,(W1,5) dt=limnF(ζn)=F(ζ).

    Moreover, by the definition of the weak time derivative, we have

    T0tσ,ζnW1,5,(W1,5) dt=T0Ωσtζn dx dtT0Ωσtζ dx dt as n.

    Hence, we obtain

    T0Ωσtζ dx dt=F(ζ)=T0Ωn(φ)(Dσχφ)ζS(φ,μ,σ)ζ dx dt

    for all ζC1c(0,T;H1). This implies that the weak time derivative tσ satisfies

    T0tσζH1,(H1) dt=F(ζ)ζC1c(0,T;H1),

    and as F belongs to L2(0,T;(H1)), we find that tσ also belongs to L2(0,T;(H1)). Furthermore, due to the improved regularity tσL2(0,T;(H1)), we use the continuous embedding

    L2(0,T;H1)H1(0,T;(H1))C0([0,T];L2)

    to deduce that σ(0)=σ0.


    6.2. Limit of vanishing chemotaxis

    For χ,b(0,1], let (φχ,μχ,σχ,vχ,pχ) denote a weak solution to (1.1)-(1.2) obtain from Theorem 2.1. By (2.14) we deduce that, for a relabelled subsequence as b0 and χ0, the following weak/weak-* convergences:

    φχφ weakly -  in L(0,T;H1)L2(0,T;H3)W1,85(0,T;(H1)),σχσ weakly -  in L2(0,T;H1)L(0,T;L2)W1,54(0,T;(W1,5)),μχμ weakly  in L2(0,T;H1),pχp weakly  in L85(0,T;H1)L87(0,T;H2),vχv weakly  in L2(0,T;L2)L87(0,T;H1),div(φχvχ)div(φv) weakly  in L85(0,T;(H1)),div(σχvχ)div(σv) weakly  in L54(0,T;(W1,5)),

    and the following strong convergences:

    φχφ strongly  in C0([0,T];Lr)L2(0,T;W2,r) and a.e. in Q,σχσ strongly  in C0([0,T];(H1))L2(0,T;Lr) and a.e. in Q,

    for any r[1,6). For any δCc(0,T) and ζH1, we have

    |T0Ωδχσχζdx dt|χσχL2(L2)ζL2δL2(0,T)0,|T0Ωδn(φχ)χφχζdx dt|n1χφχL2(L2)ζL2δL2(0,T)0,|T0Ωδχσχφχζ dx dt|χζL2σχL2(L4)φχL2(L4)δL(0,T)0,

    as χ0. Thus, multiplying (2.12) with δCc(0,T), and then passing to the limit b,χ0, we see that (φ,μ,σ,v,p) satisfies

    tφ,ζH1,(H1)=Ωm(φ)μζ+Γφ(φ,μ,σ)ζ+φvζ dx, (6.2a)
    Ωμζ dx=ΩAΨ(φ)ζ+Bφζ dx, (6.2b)
    tσ,ϕW1,5,(W1,5)=Ωn(φ)DσϕS(φ,μ,σ)ϕ+σvϕ dx, (6.2c)
    Ωpζ dx=Ω1KΓvζ+μφζ dx, (6.2d)
    Ωvζ dx=ΩK(pμφ)ζ dx, (6.2e)

    for all ζH1, ϕW1,5, ζL2 and a.e. t(0,T).


    7. Existence in two dimensions

    We first derive an analogous result to Lemma 4.2 for two dimensions.

    Lemma 7.1. Let ΩR2 be a bounded domain with C3-boundary. Let φL(0,T;H1)L2(0,T;H2NH3), σL2(0,T;H1), μL2(0,T;H1), the source term ΓvL2(0,T;L20), and the function p satisfying (4.7). Then,

    pLk(0,T;H1)Lq(0,T;H2), vLq(0,T;H1),

    for any

    1k<2,1q<43.

    Proof. We estimate (4.23) differently than in the proof of Lemma 4.2. By Hölder’s inequality for any 1s< we have

    (μ+χσ)φL2μ+χσL2sφL2ss1.

    Then, by the Gagliardo--Nirenberg inequality (1.10) with p=2ss1, j=0, r=2, m=2, d=2, q=2, and α=12s12s=12s, we find that

    φL2ss1Cφ12sH2φ112sL2Cφ12sH3φ112sH1.

    Then, by Hölder's inequality and the Sobolev embedding H1Lr for 1r< in two dimensions, we have for w,y1,

    T0(μ+χσ)φwL2dtCT0μ+χσwL2sφwL2ss1dtCφw2s12sL(H1)μ+χσwLwy(H1)φw2sLw2syy1(H3).

    As μ,σ belong to L2(0,T;H1) and φ belongs to L2(0,T;H3)L(0,T;H1), we need

    wy=2,w2syy1=2y=2s+12s,w=4s1+2s.

    Since w=4s1+2s<2 for all s[1,), and ΓvL2(0,T;L20), the computations in the proof of Lemma 4.2 yields that

    pLk(0,T;H1) for 1k<2

    Next, we see that

    div((μ+χσ)φ)L2(μ+χσ)ΔφL2+(μ+χσ)φL2μ+χσL2sΔφL2ss1+(μ+χσ)L2φL.

    By the Gagliardo--Nirenberg inequality (1.10) with p=, j=0, r=2, m=2, d=2, q=2 and α=12, we have

    φLCφ12H2φ12L2Cφ12H3φ12H1, (7.1)

    and with p=2ss1, j=1, r=2, m=2, d=2, q=2 and α=s+12s(12,1] for s[1,), we have

    ΔφL2ss1Cφs+12sH2φs12sL2Cφs+12sH3φs12sH1. (7.2)

    Hence, for w,y,z1, we find that

    T0div((μ+χσ)φ)wL2dtCφw2L(H1)μ+χσwLwz(H1)φw2Lw2zz1(H3)+Cφws12sL(H1)μ+χσwLwy(H1)φws+12sLws+12syy1(H3).

    Since s+12s1 for all s[1,), we require

    wy=2,wy(s+1)2s(y1)=2y=3s+12s,w=4s1+3s.

    We choose z=3s+12s(32,2] so that

    wz=2,w2zz1=2s1+3s3s+1s+1=2ss+1[1,2),

    and thus we obtain

    T0div((μ+χσ)φ)4s1+3sL2dtCφ2s1+3sL(H1)μ+χσ4s1+3sL2(H1)φ2s1+3sL2ss+1(H3)+Cφ2s21+3sL(H1)μ+χσ4s1+3sL2(H1)φ2s+21+3sL2(H3).

    From (4.27) and using the fact that 4s1+3s<4s1+2s for all s[1,), we see that

    pLq(0,T;H2) for 1q<43.

    Similarly, from (4.29), (7.1) and (7.2), we obtain for fixed 1i,j2, and any s[1,),

    DivjL2=KDiDjp(Di(μ+χσ)Djφ(μ+χσ)DiDjφL2K(pH2+(μ+χσ)L2φL+μ+χσL2sD2φL2ss1)K(pH2+Cμ+χσH1(φ12H3φ12H1+φs+12sH3φs12sH1)). (7.3)

    Then, a similar calculation shows that the right-hand side is bounded in L4s1+3s(0,T), which in turn implies that

    vLq(0,T;H1) for 1q<43.

    By the above new estimates we can show that div(φv) and tφ have improved temporal regularity, and that div(σv) and tσ belong to the dual space (H1).

    Lemma 7.2. For dimension d=2, let (φk,μk,σk,pk,vk) denote the Galerkin ansatz from Section 3 satisfying (4.44). Then, it holds that for 43w<2 and 1<r<87,

    div(φkvk)Lw((H1))+div(σkvk)Lr((H1))K12E,tφkLw((H1))+tσkLr((H1))E(1+K12),

    where E denotes positive constants that are uniformly bounded for b,χ(0,1] and are also uniformly bounded for K(0,1] when Γv=0.

    Proof. The assertions for tφk and tσk will follow via similar arguments in Section 4.3 once we establish the assertion for the convection terms. In dimension d=2, we have the embedding L2(0,T;H1)L(0,T;L2)L4(Q), and by the Gagliardo--Nirenberg inequality (1.10) with p=4, j=0, r=2, d=2, m=1, q=2 and α=12,

    fL4Cf12H1f12L2.

    Consider an arbitrary ζLs(0,T;H1) for some s1 yet to be determined. Then, we compute that

    |T0ΩσkvkΠkζdx dt|T0σkL4vkL4ζL2dtCσkL4(Q)(T0vk23H1vk23L2ζ43H1dt)34CσkL4(Q)vk12L23x1(H1)vk12L23x2(L2)ζL43x3(H1),

    where x1,x2,x31 satisfy

    1x1+1x2+1x3=1,23x1<43,23x22x3>6.

    Then, from (4.44) and (7.3), it holds that

    |T0ΩσkvkΠkζ dx dt|EK12ζLs(H1) for s=43x3>8,

    that is, {div(σkvk)}kN is uniformly bounded in the dual space of Ls(0,T;H1) for s>8. Similarly, by the Gagliardo--Nirenberg inequality (1.10) with p=, j=0, r=2, d=2, m=3, q[1,) and α=1q+1,

    φkLCφk1q+1H3φkqq+1LqCφk1q+1H3φkqq+1H1.

    Proceeding as in (4.50), we find that for an arbitrary ζLs(0,T;H1), where s1 is yet to be determined,

    |T0ΩφkvkΠkζdx dt|T0vkL2φkLΠkζL2dtCφkqq+1L(H1)vkL2(L2)φk1q+1L2(H3)ζL2(q+1)q(H1)EK12ζL2(q+1)q(H1),

    and so {div(φkvk)}kN is uniformly bounded in the dual space of Ls(0,T;H1) for s=2+2q(2,4].

    Remark 7.1. We point out that in the absence of the regularity result vkLq(0,T;H1) from Lemma 7.1, and if we only have vkL2(0,T;L2), then we obtain

    |T0ΩσkvkΠkζ dx dt|σkL4(Q)vkL2(L2)ζL4(L4),

    and this implies that both {div(σkvk)}kN and {tσk}kN are bounded uniformly only in L43(0,T;(W1,4)).


    8. Discussion

    Reformulations of Darcy’s law and the pressure. Associated to Darcy's law (1.1b) is the term λv:=pμφD2|σ|2 which will contribute the source term Γvλv in the energy identity (4.4). In [20, Rmk. 2.1] three other reformulations of Darcy's law (1.1b) and the pressure are considered:

    (R1) Let q:=pAΨ(φ)B2|φ|2 so that

    λv=q+AΨ(φ)+B2|φ|2D2|σ|2+χσ(1φ)μφ, (8.1a)
    v=K((qB2|φ|2)BΔφφ)=K(q+Bdiv(φφ)). (8.1b)

    (R2) Let ˆp:=p+D2|σ|2+χσ(1φ) so that

    λv=ˆpμφD|σ|2χσ(1φ), (8.2a)
    v=K(ˆpμφ(Dσ+χ(1φ))σ). (8.2b)

    (R3) Let ˜p:=pD2|σ|2μφ so that

    λv=˜p, (8.3a)
    v=K(˜p+φμ+σ(Dσ+χ(1φ))). (8.3b)

    From the viewpoint of estimating the source term Γvλv, we see that (8.3a) has the advantage of being the simplest. Meanwhile, for (8.2a) the analysis for Γvλv is similar to that performed in Section 4.1.2, but for (8.1a) the main difficulty will be to estimate the terms (AΨ(φ)+B2|φ|2)Γv and (D2|σ|2+χσφ)Γv, which at first glance would require the assumption that ΓvL(Q), and obtaining an L2-estimate for the pressure q from the Darcy law (8.1b) would be difficult due to the term div(φφ).

    Other boundary conditions for the pressure and velocity. In [20, x2.4.4] the authors have discussed possible boundary conditions for the velocity and for the pressure. As discussed in Section 2 following Assumption 2.1, we require the source term Γv to have zero mean due to the no-flux boundary condition vn=0 on Ω. The general energy identity (with homogeneous Neumann boundary conditions for φ and μ) from [20, Equ. (2.27)] reads as

    ddtΩAΨ(φ)+B2|φ|2+D2|σ|2+χσ(1φ)dx+Ωm(φ)|μ|2+n(φ)|(Dσ+χ(1φ))|2+1K|v|2dx=ΩΓφμS(Dσ+χ(1φ))+Γvλvdx+Ω(Dσ+χ(1φ))n(φ)(Dnσ)(vn)(D2|σ|2+χσ(1φ)+p) dHd1,

    and we see the appearance of an extra boundary source term involving the normal component of the velocity and the pressure. Here it would be advantageous to use the rescaled pressure ˆp and the Darcy law (8.2b), as the extra boundary source term will become

    Ω(vn)ˆp dHd1,

    which motivates the consideration of a Robin-type boundary condition for ˆp

    g=aˆpvn=aˆp+KnˆpK(Dσ+χ(1φ))nσ on Ω,

    for some given datum g and positive constant a. On one hand, this would allow us to consider source terms Γv that need not have zero mean, but on the other hand, the analysis of the Darcy system becomes more complicated. In particular, the weak formulation of the pressure system now reads as

    ΩKˆpζdx+ΩaˆpζdHd1=ΩΓvζ+K(μφ+(Dσ+χ(1φ))σ)ζdx+ΩgζdHd1,

    and we observe that the term Dσσ on the right-hand side belongs to L1 as σ has at most H1-spatial regularity from the energy identity. Thus, it is not clear if the pressure system can be solved with the regularities stated in Lemma 4.1. A deeper study into the theory of linear elliptic equations with right-hand sides of the form divf where fL1 is required.


    Conflict of Interest

    All authors declare no conflicts of interest in this paper.


    [1] R.A. Adams and J.J.F. Fournier, Sobolev spaces. Pure and applied mathematics, Vol. 140. 2 Ed., Elsevier/Academic Press, Amsterdam, 2003.
    [2] H.W. Alt, Linear Functional Analysis. An Application-Oriented Introduction. Translated from the German edition by Robert Nürnberg. Universitext. Springer Berlin London, 2016.
    [3] S. Bosia, M. Conti, and M. Grasselli, On the Cahn-Hilliard-Brinkman system, Commun. Math. Sci., 13 (2015), no. 6, 1541-1567.
    [4] E.A. Coddington and N. Levinson, Theory of Ordinary Di erential Equations. International series in pure and applied mathematics. Tata McGraw-Hill, New York, 1955.
    [5] P. Colli, G. Gilardi, and D. Hilhorst, On a Cahn–Hilliard type phase field model related to tumor growth. Discrete Contin. Dyn. Syst., 35 (2015), no. 6, 2423-2442.
    [6] P. Colli, G. Gilardi, E. Rocca, and J. Sprekels, Vanishing viscosities and error estimate for a Cahn-Hilliard type phase field system related to tumor growth, Nonlinear Anal. Real World Appl., 26 (2015), 93-108.
    [7] P. Colli, G. Gilardi, E. Rocca, and J. Sprekels, Asymptotic analyses and error estimates for a Cahn-Hilliard type phase field system modelling tumor growth, Discrete Contin. Dyn. Syst. Ser. S., in press (2016).
    [8] V. Cristini and J. Lowengru, Multiscaled modeling of cancer. An Integrated Experiemental and Mathematical Modeling Approach. Cambridge University Press, 2010.
    [9] M. Dai, E. Feireisl, E. Rocca, G. Schimperna, and M. Schonbek, Analysis of a di use interface model for multispecies tumor growth, Preprint arXiv:1507.07683 (2015).
    [10] F. Della Porta and M. Grasselli, On the nonlocal Cahn-Hilliard-Brinkman and Cahn-Hilliard-Hele-Shaw systems, Commun. Pure Appl. Anal., 15 (2016), 299-317.
    [11] E. DiBenedetto, Degenerate Parabolic Equations. Universitext. Springer–Verlag New York, 1993.
    [12] L.C. Evans, Partial Differential Equations. Graduate Studies in Mathematics, Volume 19. AMS,Providence, Rhode Island, 2002.
    [13] A. Fasano, A. Bertuzzi, and A. Gandolfi, Mathematical modeling of tumour growth and treatment.Complex Systems in Biomedicine. Springer Milan, 2006.
    [14] X. Feng and S.Wise, Analysis of a Darcy-Cahn-Hilliard diffuse interface model for the Hele-Shawflow and its fully discrete finite element approximation, SIAM J. Numer. Anal., 50 (2012), no. 3,1320-1343.
    [15] A. Friedman, Partial Differential Equations. Holt, Rinehart and Winston, New York, 1969.
    [16] S. Frigeri, M. Grasselli, and E. Rocca, On a diffuse interface model of tumor growth, European J.Appl. Math., 26 (2015), 215-243.
    [17] H. Garcke and K.F. Lam, Well-posedness of a Cahn-Hilliard system modelling tumour growth withchemotaxis and active transport, European J. Appl. Math., in press (2016). DOI http://dx.doi.org/10.1017/S0956792516000292.
    [18] H. Garcke and K.F. Lam, Analysis of a Cahn-Hilliard system with non zero Dirichlet conditionsmodelling tumour growth with chemotaxis, Preprint arXiv:1604.00287 (2016).
    [19] H. Garcke, K.F. Lam, and E. Rocca, Optimal control of treatment time in a diffuse interface modelfor tumor growth, Preprint arXiv:1608.00488 (2016).
    [20] H. Garcke, K.F. Lam, E. Sitka, and V. Styles, A Cahn-Hilliard-Darcy model for tumour growth withchemotaxis and active transport, Math. Models Methods Appl. Sci., 26 (2016), no. 6, 1095-1148.
    [21] A. Hawkins-Daarud, K.G. van der Zee, and J.T. Oden, Numerical simulation of a thermodynamicallyconsistent four-species tumor growth model, Int. J. Numer. Methods Biomed. Eng., 28 (2012),3-24.
    [22] J. Jiang, H.Wu, and S. Zheng, Well-posedness and long-time behavior of a non-autonomous Cahn-Hilliard-Darcy system with mass source modeling tumor growth, J. Differential Equ., 259 (2015),no. 7, 3032-3077.
    [23] H. Lee, J. Lowengrub, and J. Goodman, Modeling pinchoff and reconnection in a Hele–Shaw cell.I. The models and their calibration, Phys. Fluids, 14 (2002), no. 2, 492-513.
    [24] H. Lee, J. Lowengrub, and J. Goodman, Modeling pinchoff and reconnection in a Hele-Shaw cell.II. Analysis and simulation in the nonlinear regime, Phys. Fluids, 14 (2002), no. 2, 514-545.
    [25] J.S. Lowengrub, E. Titi, and K. Zhao, Analysis of a mixture model of tumor growth, European J.Appl. Math., 24 (2013), 691-734.
    [26] M. Renardy and R.C. Rogers, An Introduction to Partial Differential Equations. Texts in AppliedMathematics. 2 Eds., Springer-Verlag New York, 2004.
    [27] H.L. Royden and P. Fitzpatrick, Real Analysis. Featured Titles for Real Analysis Series. 4 Eds.,Pearson Prentice Hall, Boston, 2010.
    [28] J. Simon, Compact sets in space Lp(0; T; B), Ann. Mat. Pura Appl., 146 (1986), no. 1, 65-96.
    [29] X. Wang and H. Wu, Long-time behavior for the Hele-Shaw-Cahn-Hilliard system, Asymptot.Anal., 78 (2012), no. 4, 217-245.
    [30] X. Wang and Z. Zhang, Well-posedness of the Hele-Shaw-Cahn-Hilliard system, Ann. Inst. H. Poincar´e Anal. Non Lin´eaire, 30 (2013), no. 3, 367-384.
  • This article has been cited by:

    1. Harald Garcke, Kei Fong Lam, Elisabetta Rocca, Optimal Control of Treatment Time in a Diffuse Interface Model of Tumor Growth, 2018, 78, 0095-4616, 495, 10.1007/s00245-017-9414-4
    2. Andrea Giorgini, Maurizio Grasselli, Hao Wu, The Cahn–Hilliard–Hele–Shaw system with singular potential, 2018, 35, 02941449, 1079, 10.1016/j.anihpc.2017.10.002
    3. Harald Garcke, Kei Fong Lam, Analysis of a Cahn--Hilliard system with non-zero Dirichlet conditions modeling tumor growth with chemotaxis, 2017, 37, 1553-5231, 4277, 10.3934/dcds.2017183
    4. Alain Miranville, Elisabetta Rocca, Giulio Schimperna, On the long time behavior of a tumor growth model, 2019, 267, 00220396, 2616, 10.1016/j.jde.2019.03.028
    5. Andrea Signori, Penalisation of long treatment time and optimal control of a tumour growth model of Cahn–Hilliard type with singular potential, 2021, 41, 1553-5231, 2519, 10.3934/dcds.2020373
    6. Harald Garcke, Kei Fong Lam, 2018, Chapter 12, 978-3-319-75939-5, 243, 10.1007/978-3-319-75940-1_12
    7. KEI FONG LAM, HAO WU, Thermodynamically consistent Navier–Stokes–Cahn–Hilliard models with mass transfer and chemotaxis, 2018, 29, 0956-7925, 595, 10.1017/S0956792517000298
    8. Matthias Ebenbeck, Harald Garcke, On a Cahn--Hilliard--Brinkman Model for Tumor Growth and Its Singular Limits, 2019, 51, 0036-1410, 1868, 10.1137/18M1228104
    9. Luca Dedè, Alfio Quarteroni, Isogeometric Analysis of a Phase Field Model for Darcy Flows with Discontinuous Data, 2018, 39, 0252-9599, 487, 10.1007/s11401-018-0079-3
    10. Andrea Signori, Optimal Distributed Control of an Extended Model of Tumor Growth with Logarithmic Potential, 2020, 82, 0095-4616, 517, 10.1007/s00245-018-9538-1
    11. Marvin Fritz, Ernesto A. B. F. Lima, Vanja Nikolić, J. Tinsley Oden, Barbara Wohlmuth, Local and nonlocal phase-field models of tumor growth and invasion due to ECM degradation, 2019, 29, 0218-2025, 2433, 10.1142/S0218202519500519
    12. Matthias Ebenbeck, Harald Garcke, Analysis of a Cahn–Hilliard–Brinkman model for tumour growth with chemotaxis, 2019, 266, 00220396, 5998, 10.1016/j.jde.2018.10.045
    13. Cecilia Cavaterra, Elisabetta Rocca, Hao Wu, Long-Time Dynamics and Optimal Control of a Diffuse Interface Model for Tumor Growth, 2019, 0095-4616, 10.1007/s00245-019-09562-5
    14. Michele Colturato, Sliding mode control for a diffuse interface tumor growth model coupling a Cahn–Hilliard equation with a reaction–diffusion equation, 2020, 43, 0170-4214, 6598, 10.1002/mma.6403
    15. Harald Garcke, Kei Fong Lam, Andrea Signori, On a phase field model of Cahn–Hilliard type for tumour growth with mechanical effects, 2021, 57, 14681218, 103192, 10.1016/j.nonrwa.2020.103192
    16. Pierluigi Colli, Gianni Gilardi, Jürgen Sprekels, A Distributed Control Problem for a Fractional Tumor Growth Model, 2019, 7, 2227-7390, 792, 10.3390/math7090792
    17. Andrea Signori, Vanishing parameter for an optimal control problem modeling tumor growth, 2020, 117, 18758576, 43, 10.3233/ASY-191546
    18. Sergio Frigeri, Kei Fong Lam, Elisabetta Rocca, 2017, Chapter 9, 978-3-319-64488-2, 217, 10.1007/978-3-319-64489-9_9
    19. Pierluigi Colli, Gianni Gilardi, Jürgen Sprekels, Asymptotic analysis of a tumor growth model with fractional operators, 2020, 120, 18758576, 41, 10.3233/ASY-191578
    20. Matthias Ebenbeck, Kei Fong Lam, Weak and stationary solutions to a Cahn–Hilliard–Brinkman model with singular potentials and source terms, 2020, 10, 2191-950X, 24, 10.1515/anona-2020-0100
    21. Jürgen Sprekels, Hao Wu, Optimal Distributed Control of a Cahn–Hilliard–Darcy System with Mass Sources, 2021, 83, 0095-4616, 489, 10.1007/s00245-019-09555-4
    22. Matthias Ebenbeck, Patrik Knopf, Optimal medication for tumors modeled by a Cahn–Hilliard–Brinkman equation, 2019, 58, 0944-2669, 10.1007/s00526-019-1579-z
    23. Marvin Fritz, Ernesto A. B. F. Lima, J. Tinsley Oden, Barbara Wohlmuth, On the unsteady Darcy–Forchheimer–Brinkman equation in local and nonlocal tumor growth models, 2019, 29, 0218-2025, 1691, 10.1142/S0218202519500325
    24. Pierluigi Colli, Andrea Signori, Jürgen Sprekels, Optimal Control of a Phase Field System Modelling Tumor Growth with Chemotaxis and Singular Potentials, 2019, 0095-4616, 10.1007/s00245-019-09618-6
    25. Luca Dedè, Harald Garcke, Kei Fong Lam, A Hele–Shaw–Cahn–Hilliard Model for Incompressible Two-Phase Flows with Different Densities, 2018, 20, 1422-6928, 531, 10.1007/s00021-017-0334-5
    26. Harald Garcke, Sema Yayla, Long-time dynamics for a Cahn–Hilliard tumor growth model with chemotaxis, 2020, 71, 0044-2275, 10.1007/s00033-020-01351-3
    27. Pierluigi Colli, Gianni Gilardi, Gabriela Marinoschi, Elisabetta Rocca, Sliding Mode Control for a Phase Field System Related to Tumor Growth, 2019, 79, 0095-4616, 647, 10.1007/s00245-017-9451-z
    28. Harald Garcke, Kei Fong Lam, Robert Nürnberg, Emanuel Sitka, A multiphase Cahn–Hilliard–Darcy model for tumour growth with necrosis, 2018, 28, 0218-2025, 525, 10.1142/S0218202518500148
    29. Matthias Ebenbeck, Harald Garcke, Robert Nürnberg, Cahn–Hilliard–Brinkman systems for tumour growth, 2021, 0, 1937-1179, 0, 10.3934/dcdss.2021034
    30. Marvin Fritz, Prashant K. Jha, Tobias Köppl, J. Tinsley Oden, Barbara Wohlmuth, Analysis of a new multispecies tumor growth model coupling 3D phase-fields with a 1D vascular network, 2021, 61, 14681218, 103331, 10.1016/j.nonrwa.2021.103331
    31. Cecilia Cavaterra, Sergio Frigeri, Maurizio Grasselli, Nonlocal Cahn–Hilliard–Hele–Shaw Systems with Singular Potential and Degenerate Mobility, 2022, 24, 1422-6928, 10.1007/s00021-021-00648-1
    32. Luca Scarpa, Andrea Signori, On a class of non-local phase-field models for tumor growth with possibly singular potentials, chemotaxis, and active transport, 2021, 34, 0951-7715, 3199, 10.1088/1361-6544/abe75d
    33. Patrik Knopf, Andrea Signori, Existence of weak solutions to multiphase Cahn–Hilliard–Darcy and Cahn–Hilliard–Brinkman models for stratified tumor growth with chemotaxis and general source terms, 2022, 47, 0360-5302, 233, 10.1080/03605302.2021.1966803
    34. Pierluigi Colli, Andrea Signori, Jürgen Sprekels, Optimal Control Problems with Sparsity for Tumor Growth Models Involving Variational Inequalities, 2022, 194, 0022-3239, 25, 10.1007/s10957-022-02000-7
    35. Pierluigi Colli, Gianni Gilardi, Elisabetta Rocca, Jürgen Sprekels, Well-posedness and optimal control for a Cahn–Hilliard–Oono system with control in the mass term, 2022, 15, 1937-1632, 2135, 10.3934/dcdss.2022001
    36. Andrea Giorgini, Kei Fong Lam, Elisabetta Rocca, Giulio Schimperna, On the Existence of Strong Solutions to the Cahn--Hilliard--Darcy System with Mass Source, 2022, 54, 0036-1410, 737, 10.1137/20M1376443
    37. Pierluigi Colli, Andrea Signori, Jürgen Sprekels, Second-order analysis of an optimal control problem in a phase field tumor growth model with singular potentials and chemotaxis, 2021, 27, 1292-8119, 73, 10.1051/cocv/2021072
    38. Mostafa Bendahmane, Kenneth H. Karlsen, Martingale solutions of stochastic nonlocal cross-diffusion systems, 2022, 17, 1556-1801, 719, 10.3934/nhm.2022024
    39. Tania Biswas, Elisabetta Rocca, Long time dynamics of a phase-field model of prostate cancer growth with chemotherapy and antiangiogenic therapy effects, 2022, 27, 1531-3492, 2455, 10.3934/dcdsb.2021140
    40. Elisabetta Rocca, Luca Scarpa, Andrea Signori, Parameter identification for nonlocal phase field models for tumor growth via optimal control and asymptotic analysis, 2021, 31, 0218-2025, 2643, 10.1142/S0218202521500585
    41. Giulio Schimperna, On the Cahn-Hilliard-Darcy system with mass source and strongly separating potential, 2022, 15, 1937-1632, 2305, 10.3934/dcdss.2022008
    42. Xiaopeng Zhao, Optimal Distributed Control of Two-Dimensional Navier–Stokes–Cahn–Hilliard System with Chemotaxis and Singular Potential, 2023, 88, 0095-4616, 10.1007/s00245-023-09976-2
    43. Marvin Fritz, Tumor Evolution Models of Phase-Field Type with Nonlocal Effects and Angiogenesis, 2023, 85, 0092-8240, 10.1007/s11538-023-01151-6
    44. Christoph Helmer, Ansgar Jüngel, Existence analysis for a reaction-diffusion Cahn–Hilliard-type system with degenerate mobility and singular potential modeling biofilm growth, 2023, 0, 1078-0947, 0, 10.3934/dcds.2023069
    45. Jürgen Sprekels, Fredi Tröltzsch, Second-Order Sufficient Conditions in the Sparse Optimal Control of a Phase Field Tumor Growth Model with Logarithmic Potential, 2024, 30, 1292-8119, 13, 10.1051/cocv/2023084
    46. Helmut Abels, Harald Garcke, Jonas Haselböck, Existence of weak solutions to a Cahn–Hilliard–Biot system, 2025, 81, 14681218, 104194, 10.1016/j.nonrwa.2024.104194
    47. Cecilia Cavaterra, Sergio Frigeri, Maurizio Grasselli, A Nonlocal Cahn–Hilliard–Darcy System with Singular Potential, Degenerate Mobility, and Sources, 2025, 91, 0095-4616, 10.1007/s00245-025-10239-5
    48. Cedric Riethmüller, Erlend Storvik, Jakub Wiktor Both, Florin Adrian Radu, Well-posedness analysis of the Cahn–Hilliard–Biot model, 2025, 84, 14681218, 104271, 10.1016/j.nonrwa.2024.104271
  • Reader Comments
  • © 2016 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(7506) PDF downloads(1364) Cited by(47)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog