Skip to main content
Springer logoLink to Springer
. 2022 Sep 20;29(6):72. doi: 10.1007/s00030-022-00805-z

Phase transitions in porous media

Chiara Gavioli 1,, Pavel Krejčí 2
PMCID: PMC9519144  PMID: 36187358

Abstract

The full quasistatic thermomechanical system of PDEs, describing water diffusion with the possibility of freezing and melting in a visco-elasto-plastic porous solid, is studied in detail under the hypothesis that the pressure-saturation hysteresis relation is given in terms of the Preisach hysteresis operator. The resulting system of balance equations for mass, momentum, and energy coupled with the phase dynamics equation is shown to admit a global solution under general assumptions on the data.

Keywords: Porous media, Phase transitions, Hysteresis

Introduction

A model for fluid flow in partially saturated porous media with thermomechanical interaction was proposed and analyzed in [2, 4]. The model was subsequently extended in [15] by including the effects of freezing and melting of the water in the pores. Typical examples in which such situations arise are related to groundwater flows and to the freezing-melting cycles of water sucked into the pores of concrete. Due to the specific volume difference between water and ice, this process produces important pressure changes and represents one of the main reasons for the degradation of construction materials in buildings, bridges, and roads. The model of [15] still neglects the influence of changes of microstructure, as for example the breaking of pores, but the main thermomechanical interactions between the state variables are taken into account.

The model is based on the assumption that slow diffusion of the fluid through the porous solid is a dominant effect, so that the Lagrangian description is considered to be appropriate. It is assumed that volume changes of the solid matrix material are negligible with respect to the pore volume evolution during the process. The pores are filled with a mixture of H2O and gas, and H2O itself is a mixture of the liquid (water) and the solid phase (ice). That is, in addition to the standard state variables like capillary pressure, displacement, and absolute temperature, we need to consider the evolution of a phase parameter χ representing the relative proportion of water in the H2O part and its influence on pressure changes due to the different mass densities of water and ice.

The resulting system consists of mechanical balance equation for the deformations of the solid body, mass balance equation based on the Darcy law for the fluid diffusion with interaction terms similar to the Biot system studied, e. g., in [20], a differential inclusion for the phase fraction χ of relaxed Stefan type as in [22] and governing the water-ice phase transition, and the energy balance equation derived from the first and the second principles of thermodynamics with heat sources due to viscosity, plasticity, diffusion, and phase transition.

The present paper develops the ideas of [15] in the sense that the effects of capillary hysteresis, which is assumed to be of Preisach type in agreement with the results of [6], shear stresses, and elastoplasticity are considered in full generality. This represents an enormous increase of mathematical complexity. While the momentum balance in the shear-stress-free case in [15] can be reduced to an ODE, here, we need to exploit deeper results from the theory of PDEs to control the interactions between individual components of the system, as well as a generalized Moser iteration scheme from [8]. Additional difficulties are due to the effects of the three heat sources produced by mechanical hysteresis dissipation (plasticity, capillarity, phase transitions).

The paper is divided into five sections. In the next Sect. 1, we briefly recall the principles of the model introduced in [15], taking into account capillary hysteresis and elastoplastic hysteresis effects as in [4]. Section 2 is devoted to a short survey of the Preisach hysteresis model. In Sect. 3, we state the mathematical problem, the main assumptions on the data, and the main Theorem 3.3, the proof of which is split into Sects. 4 and 5. The steps of the proof are as follows. In Sect. 4, we first cut off some of the pressure and temperature dependent terms in the system by means of a cut-off parameter R, regularize the mass balance equation with a fourth order term depending on an additional small regularizing parameter η, solve the related problem employing a Galerkin approximation scheme, and pass to the singular limit η0 in the regularizing term. Then, in Sect. 5, we derive a series of R-independent estimates like the energy estimate, the so-called Dafermos estimate (with negative small powers of the temperature), Moser-type and then higher-order estimates for the capillary pressure and for the temperature which allow us in Sect. 6 to pass to the limit in the cut-off system as R, which will conclude the proof of the existence result.

The model

Consider a bounded domain ΩR3 of class C1,1 filled with an elastoplastic solid matrix material with pores containing a mixture of H2O and gas, where we assume that H2O may appear in one of the two phases: water or ice. We state the balance laws in referential (Lagrangian) coordinates. We have in mind construction materials where large deformations are not expected to occur. This hypothesis enables us to reduce the complexity of the problem and assume that the deformations are small in order to avoid higher degree nonlinearities. We denote for xΩ and time t[0,T]

  • p(xt) ... capillary pressure;

  • u(xt) ... displacement vector in the solid;

  • ε(x,t)=su(x,t) ... linear strain tensor, (su)ij:=12uixj+ujxi;

  • θ(x,t) ... absolute temperature;

  • χ(x,t)[0,1] ... relative amount of water in the H2O part.

The model derived in [15] aims at coupling the effects of capillarity, interaction between a deformable solid matrix material and H2O in the pores which may undergo water-ice phase transitions, and energy exchange between the individual components of the system. Hysteresis is included following the modeling section of [4]. The full system consists of equations describing mass balance (1.1), mechanical equilibrium (1.2), energy balance (1.3) and phase evolution (1.4) in the form

((χ+ρ(1-χ))(G[p]+divu))t=div(μ(p)p), 1.1
-div(Bsut+P[su])+(p(χ+ρ(1-χ))+β(θ-θc))=g, 1.2
CV(θ)t-div(κ(θ)θ)=Bsut:sut+μ(p)|p|2+DP[su]t+(χ+ρ(1-χ))|DG[p]t|+γ(θ,divu)χt2-Lθcθχt-βθdivut, 1.3
-γ(θ,divu)χt+(1-ρ)pG[p]-UG[p]+pdivu+Lθθc-1I[0,1](χ). 1.4

We refer to [15] for the details of the physical arguments. Let us just mention that the mass balance (1.1) is derived from Darcy’s law, and μ(p)p is the specific liquid mass flux. The constant ρ(0,1) is the ratio between ice and water mass densities, whereas the symbol G describes the pressure–saturation curve and, following [2, 4], is of the form

G[p]=f(p)+G0[p].

Here f is a bounded monotone function satisfying Hypothesis 3.1 (vi) below, whereas G0 is the Preisach hysteresis operator from Sect. 2. The momentum balance (1.2) is derived by assuming that the process is quasi-static, so that the inertia term is negligible. Here B is a positive definite viscosity matrix, P the constitutive operator of elastoplasticity defined below in (3.6), β is the thermal expansion coefficient, θc is the melting temperature at standard pressure and g is a given volume force (gravity, e. g.). The term p(χ+ρ(1-χ)) represents the pressure component due to the phase transition. Finally, the energy balance (1.3) and the inclusion (1.4) for the evolution of the phase parameter are derived from the principles of thermodynamics with the aid of the energy balance for both the plasticity and the pressure–saturation operator

P[ε]:εt=UP[ε]t+DP[ε]t,G[p]tp=UG[p]t+|DG[p]t|, 1.5

where UP,UG and DP,DG are the potential and dissipation operators, and · is a seminorm in the space Rsym3×3 of symmetric 3×3 tensors. In (1.3), CV(θ) is the caloric component of the internal energy, κ(θ) is the heat conductivity coefficient, L is the latent heat, I[0,1] is the indicator function of the interval [0, 1] and I[0,1] is its subdifferential, γ(θ,divu) is the phase relaxation time (which we assume to explicitly depend on both θ and divu for technical reasons).

Note that the values of G have to be naturally confined between 0 and 1, so that the system is degenerate in the sense that we do not control a priori the time derivatives of p in (1.1). Another difficulty is related to the lack of spatial regularity of χ. The temperature field is problematic as well: Eq.  (1.3) contains high order heat source terms, which are difficult to handle and prevent the temperature from being regular.

We complement the system (1.1)–(1.4) with initial conditions

p(x,0)=p0(x)u(x,0)=u0(x)θ(x,0)=θ0(x)χ(x,0)=χ0(x)inΩ, 1.6

and boundary conditions

u=0μ(p)p·n=α(x)(p-p)κ(θ)θ·n=ω(x)(θ-θ)onΩ, 1.7

where p is a given outer pressure, θ is a given outer temperature, α:Ω[0,) is the permeability of the boundary and ω:Ω[0,) is the heat conductivity of the boundary.

The solution to (1.1)–(1.4) was only constructed in [15] under the assumption that shear stresses in the momentum balance Eq. (1.2) as well as all hysteresis effects are neglected. Then (1.2) turns into an ODE for the relative volume change divu, which considerably simplifies the analysis. Here we prove existence of a global solution for the full problem under suitable hypotheses.

Hysteresis in capillarity phenomena

The operator G is considered as a sum

G[p]=f(p)+G0[p], 2.1

where f is a monotone function satisfying Hypothesis 3.1 (vi) in Sect. 3 below, and G0 is a Preisach operator that we briefly describe here.

For a given input function pW1,1(0,T) and a memory parameter r>0 we define the scalar function ξr(t) as the solution of the variational inequality

|p(t)-ξr(t)|rt[0,T],(ξr)t(p(t)-ξr(t)-z)0a.\,e.z[-r,r], 2.2

with prescribed initial condition

ξr(0)=max{p(0)-r,min{0,p(0)+r}}. 2.3

We have indeed for all r>0 the initial bound

ξr(0)|p(0)|. 2.4

The mapping fr:W1,1(0,T)W1,1(0,T) which with each pW1,1(0,T) associates the solution ξr=fr[p]W1,1(0,T) of (2.2)–(2.3) is called the play. This concept goes back to [10], and the proof of the following statements can be found e.g. in [12].

Proposition 2.1

For each r>0, the mapping fr:W1,1(0,T)W1,1(0,T) is Lipschitz continuous and admits a Lipschitz continuous extension to fr:C[0,T]C[0,T] in the sense that for every p1,p2C[0,T] and every t[0,T] we have

|fr[p1](t)-fr[p2](t)|maxτ[0,t]|p1(τ)-p2(τ)|. 2.5

Moreover, for each pW1,1(0,T), the energy balance equation

fr[p]tp-12fr2[p]t=|rfr[p]t| 2.6

and the identity

fr[p]tpt=fr[p]t2 2.7

hold a. e. in (0, T).

Given a nonnegative function ψL1((0,)×R), called the Preisach density, we define the Preisach operator G0 as a mapping that with each pC[0,T] associates the integral

G0[p](t)=00fr[p](t)ψ(r,v)dvdr. 2.8

Hence,

G0[p](t)00ψ(r,v)dvdr=:Cψ+,-G0[p](t)0-0ψ(r,v)dvdr=:Cψ- 2.9

for all pC[0,T] and all t[0,T], and we assume

0<Cψ±<12. 2.10

From (2.6)–(2.8) we immediately deduce the Preisach energy identity

G0[p]tp-U0[p]t=|D0[p]t|a.\,e., 2.11

provided we define the Preisach potential U0 and the dissipation operator D0 by the integrals

U0[p](t)=00fr[p](t)vψ(r,v)dvdr,D0[p](t)=00fr[p](t)rψ(r,v)dvdr. 2.12

The energy identity in (1.5) then holds with the choice

UG[p]=pf(p)-0pf(z)dz+U0[p],DG[p]=D0[p]. 2.13

With the notation

Φ(p)=0pf(z)dz,V(p)=pf(p)-Φ(p)=0pf(z)zdz 2.14

we can also write

UG[p]=V(p)+U0[p], 2.15

thus separating the hysteretic from the non-hysteretic part.

For our purposes, we adopt the following hypothesis on the Preisach density.

Hypothesis 2.2

There exists a function ψL1(0,) such that for a. e. (r,v)(0,)×R we have 0ψ(r,v)ψ(r) and

Cψ:=0(1+r2)ψ(r)dr<.

A straightforward computation shows that G0 (and, consequently, G) are Lipschitz continuous in C[0, T]. Indeed, by (2.5) and Hypothesis 2.2 we obtain for p1,p2C[0,T] and t[0,T] that

|G0[p2](t)-G0[p1](t)|=|0fr[p1](t)fr[p2](t)ψ(r,v)dvdt|Cψmaxτ[0,t]|p2(τ)-p1(τ)|. 2.16

Moreover, the Preisach potential is continuous from L2(Ω;C[0,T]) to L1(Ω;C[0,T]), as it holds

|U0[p2](t)-U0[p1](t)|Cψ2maxτ[0,t]|p2-p1|(τ)|p2(t)|+|p1(t)|+2. 2.17

From Hypothesis 2.2 and identity (2.7) for the play we also obtain

0<U0[p]Cψ(1+|p|)2,|D0[p]t|C|pt|. 2.18

The Preisach operator admits also a family of “nonlinear” energies. As a consequence of (2.7), we have for a. e. t the inequality

fr[p]t(p-fr[p])0,

hence

fr[p]t(h(p)-h(fr[p]))0

for every nondecreasing function h:RR. Hence, for every absolutely continuous input p, a counterpart of (2.11) in the form

G0[p]th(p)-Uh[p]t0a.\,e. 2.19

holds with a modified potential

Uh[p](t)=00fr[p](t)h(v)ψ(r,v)dvdr. 2.20

This is related to the fact that for every absolutely continuous nondecreasing function h^:RR, the mapping Gh^:=G0h^ is also a Preisach operator, see [13].

Statement of the problem

We introduce the spaces

X=W1,2(Ω),X0={ψW1,2(Ω;R3):ψ|Ω=0},Xq=W1,q(Ω) 3.1

for some q>2 that will be specified below in Theorem 3.3. Taking into account the boundary conditions (1.7), we consider (1.1)–(1.4) in variational form

Ω((χ+ρ(1-χ))(f(p)+G0[p]+divu))tϕdx+Ωμ(p)p·ϕdx=Ωα(x)(p-p)ϕds(x), 3.2
Ω(P[su]+Bsut):sψdx-Ω(p(χ+ρ(1-χ))+β(θ-θc))divψdx=Ωg·ψdx, 3.3
Ω(CV(θ)t-Bsut:sut-DP[su]t-μ(p)|p|2-(χ+ρ(1-χ))|D0[p]t|-γ(θ,divu)χt2+Lθcχt+βdivutθ)ζdx+Ωκ(θ)θ·ζdx=Ωω(x)(θ-θ)ζdx, 3.4
γ(θ,divu)χt+I[0,1](χ)(1-ρ)Φ(p)+pG0[p]-U0[p]+pdivu+Lθθc-1a.\,e. 3.5

for a. e. t(0,T) and all test functions ϕX, ψX0 and ζXq. Note that we split the capillary hysteresis terms in hysteretic and non-hysteretic part according to (2.1), (2.13)–(2.15). This is done in view of the regularization performed in Sect. 4, where only the non-hysteretic part will be affected by the cut-off.

We assume the following hypotheses to hold.

Hypothesis 3.1

There exist constants A>0, B>0, θ¯>0 such that

  • (i)

    Ae, Ah, B are constant symmetric positive definite fourth order tensors such that Aeξ:ξA|ξ|2, Ahξ:ξA|ξ|2, Bξ:ξB|ξ|2 for all ξR3×3;

  • (ii)

    gL(0,T;L2(Ω;R3))W1,2(0,T;L2(Ω;R3)) is a given function and there exists a function G^L4(Ω×(0,T)) such that g=-G^;

  • (iii)

    αW1,(Ω), α(x)0 a. e. and Ωα(x)ds(x)>0; ωL(Ω), ω(x)0 a. e. and Ωω(x)ds(x)>0;

  • (iv)

    pL(Ω×(0,T)) and ptL2(Ω×(0,T)), θL(Ω×(0,T)), θtL2(Ω×(0,T)), θ(x,t)θ¯ a. e.;

  • (v)

    p0L(Ω)W2,2(Ω), u0X0W1,4(Ω;R3), θ0L(Ω)W1,2(Ω), θ0(x)θ¯ a. e., χ0L(Ω), χ0(x)[0,1] a. e.

We also assume that there exist constants f>f>0, ν(0,1/2], μ>0, c>c>0, 1/2b<b^<1, κ>κ>0, 0<a<1-b, a<a^<(8+3a+2b)(1+b)7-2b, γ>γ>0 such that the nonlinearities satisfy the following conditions:

  • (vi)

    G[p]=f(p)+G0[p] where f:R(Cψ-,1-Cψ+) with Cψ± from (2.10) is a continuously differentiable function, f(1+|p|)-1-νf(p)f for all pR, and G0 is the Preisach operator from Sect. 2 with density function satisfying Hypothesis 2.2 and with potential U0 and dissipation D0 as in (2.12);

  • (vii)

    μ:RR is a continuous function, μ(p)μ for all pR;

  • (viii)

    CV:[0,)[0,) is a continuously differentiable function, CV(θ)=:cV(θ) is such that c(1+θb)cV(θ)c(1+θb^) for all θ0;

  • (ix)

    κ:[0,)[0,) is a continuous function, κ(1+θ1+a)κ(θ)κ(1+θ1+a^) for all θ0;

  • (x)

    γ:[0,)×[0,)[0,) is a continuous function, γ(1+θ+|divu|2)γ(θ,divu)γ(1+θ+|divu|2) for all θ0,uR3;

  • (xi)

    P:C([0,T];Rsym3×3)C([0,T];Rsym3×3) is the constitutive operator of elastoplasticity with dissipation operator DP defined below in (3.6)–(3.9).

Remark 3.2

In this remark we comment on the more technical hypotheses.

  • (vi)

    The growth condition for f is in agreement with the physical requirement that f has to degenerate when p±. The specific form of the lower bound will play a substantial role in the Moser iteration argument.

  • (viii)

    The growth condition for cV will be of fundamental importance in Sect. 5.6 where, in order to estimate divut in Lq(0,T;L2(Ω)) with an exponent q>4, we will need a higher integrability (in space) for the temperature than simply L(0,T;L1(Ω)).

  • (ix)
    The tangled bound
    a^<(8+3a+2b)(1+b)7-2b
    for the growth exponent of the function κ is required in Sect. 5.9, where we apply an iterative method in order to derive higher order estimates for the temperature.
  • (x)

    The dependence of the relaxation coefficient γ on both θ and divu is uncommon but crucial for obtaining estimates (4.34) and (5.45).

We model the elastoplasticity following [17]. We assume that a convex subset 0ZRsym3×3 with nonempty interior representing the admissible plastic stress domain is given in the space Rsym3×3 of symmetric tensors, and that the constitutive relation between the strain tensor ε and the stress tensor σ involves two fourth order tensors Ah (the kinematic hardening tensor) and Ae (the elasticity tensor). We define the constitutive operator P by the formula

P[ε]=Ahε+σp, 3.6

where σp is the solution of the variational inequality

σpZ,(εt-Ae-1σtp):(σp-z)0a.e.zZ,σp(0)=QZ(ε(0)) 3.7

for a given εW1,1(0,T;Rsym3×3), where QZ:Rsym3×3Z is the orthogonal projection onto Z. The variational inequality (3.7) has a unique solution σpW1,1(0,T;Rsym3×3) and the solution mapping

P:W1,1(0,T;Rsym3×3)W1,1(0,T;Rsym3×3):εσp

is strongly continuous, see [12]. It holds

P[ε]t|εt|. 3.8

The energy potential UP and the dissipation operator DP associated with P are defined by the formula

UP[ε]=12Ahε:ε+12Ae-1σp:σp,DP[ε]=ε-Ae-1σp. 3.9

Let MZ denote the Minkowski functional of the polar set Z to Z. The energy identity

P[ε]:εt-UP[ε]t=DP[ε]ta.e., 3.10

where ·=MZ(·) is a seminorm in Rsym3×3, and the inequalities

UP[ε]A2|ε|2,DP[ε]tC|εt| 3.11

hold for all inputs εW1,1(0,T;Rsym3×3).

The operator P can be extended to a continuous operator in the space C([0,T];Rsym3×3) in the sense that if {εm;mN} is a sequence in C([0,T];Rsym3×3), then

limmmaxt[0,T]|εm(t)-ε(t)|=0limmmaxt[0,T]|P[εm](t)-P[ε](t)|=0. 3.12

For two inputs ε1,ε2W1,1(0,T;Rsym3×3) we denote σi=P[εi], i=1,2. Then

(σ1-σ2):(ε1)t-(ε2)t12ddt(Ah(ε1-ε2):(ε1-ε2)+Ae-1(σ1p-σ2p):(σ1p-σ2p))a.e., 3.13
|σ1(t)-σ2(t)|C|ε1(0)-ε2(0)|+0t|(ε1)t-(ε2)t|(τ)dτt[0,T] 3.14

with a constant C depending only on Ah and Ae.

For inputs εL2(Ω;W1,1(0,T;Rsym3×3)) we obtain from (3.14) similarly as in [4, Formula (6.25)] the inequality

|σ(x,t)|C|ε(x,0)|+0t|εt(x,τ)|dτa.e. 3.15

The main result of the paper reads as follows.

Theorem 3.3

Let Hypothesis 3.1 hold. Then there exists a solution (p,u,θ,χ) to the system (3.2)–(3.5) with initial conditions (1.6) with the regularity

  • pL(Ω×(0,T)), ptL2(Ω×(0,T)), M(p)L2(0,T;W2,2(Ω)) with M(p) given by (5.43);

  • utLq(0,T;X0W1,q(Ω;R3)) for all q<(8+3a+2b)(4+b)7-2b, suL2(Ω;C([0,T];Rsym3×3));

  • θLq(Ω×(0,T)) for all q<(8+3a+2b)(4+b)7-2b, θL2(Ω×(0,T);R3), θtL2(0,T;W-1,q(Ω)) with q>2 given by (5.66);

  • χLq(Ω;C[0,T]), χtLq(Ω×(0,T)) for all q[1,), χ(x,t)[0,1] a. e.

The reason why we do not specify the precise value of q here is that it relies on a certain number of intermediate computations that cannot be detailed at this stage. The proof of Theorem 3.3 will be divided into several steps. In order to eliminate possible degeneracy of the functions f and μ, we start by regularizing the problem by means of a large parameter R. Then we prove that this regularized problem admits a solution by the standard Faedo–Galerkin method: here the parameter R will be of fundamental importance in order to gain some regularity. Once we have derived suitable estimates, we pass to the limit in the Faedo–Galerkin scheme. The second part of the proof will consist in the derivation of a priori estimates independent of R, which will allow us to pass to the limit in the regularized system and infer the existence of a solution with the desired regularity.

In what follows, we denote by C any positive constant depending only on the data, by CR any constant depending on the data and on R and by CR,η any constant depending on the data, on R and on η, all independent of the dimension n of the Galerkin approximation. Furthermore, we denote by |v|r the Lr(Ω)-norm of a function vLr(Ω) or vLr(Ω;R3) for r[1,], and the norm of a function vW1,r(Ω) will be denoted by |v|1;r. We systematically use the Korn’s inequality (see [19])

Ω|sw|2(x)dxcwW1,2(Ω;R3)2 3.16

for every wX0 with a constant c>0 independent of w. We will also often use the Poincaré inequality (see [9, 16]) in the form

|v|1;22CΩ|v|2(x)dx+Ωγ(x)|v|2(x)ds(x) 3.17

for functions vW1,2(Ω) provided γL(Ω) is such that γ0 a. e. and Ωγ(x)ds(x)>0. Finally, let us recall the Gagliardo–Nirenberg inequality (see [3, 9]) for vW1,r(Ω) on a bounded Lipschitzian domain ΩRN in the form

|v|qC|v|s1-δ|v|1;rδ 3.18

with r<N, s<q<(rN)/(N-r) and with a constant C depending only on qrs, where

δ=1s-1q1s-1r+1N.

Cut-off system

We choose a regularizing parameter R>1, and first solve a cut-off system with the intention to let R.

For zR we denote by

QR(z)=max{-R,min{z,R}} 4.1

the projection of R onto [-R,R]. Then we cut-off some nonlinearities by setting

fR(p)=f(p)for|p|Rf(R)+f(R)(p-R)forp>Rf(-R)+f(-R)(p+R)forp<-R, 4.2
ΦR(p)=0pfR(z)dz,VR(p)=pfR(p)-ΦR(p)=0pfR(z)zdz, 4.3
μR(p)=μ(QR(p))=μ(p)for|p|Rμ(R)forp>Rμ(-R)forp<-R, 4.4
γR(p,θ,divu)=γ(QR(θ+)+(p2-R2)+,divu) 4.5

for p,θ,divuR. Note that by Hypothesis 3.1 (vi) we deduce that |fR(p)||f(0)|+f|p|, from which

|fR(p)|C1+|p|,|ΦR(p)|C1+p2,C|p|1-ν-1VR(p)Cp2, 4.6

and also, from Hypothesis 3.1 (x),

γ1+QR(θ+)+(p2-R2)++|divu|2γR(p,θ,divu)γ1+QR(θ+)+(p2-R2)++|divu|2. 4.7

We replace (3.2)–(3.5) by the cut-off system

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tϕdx+ΩμR(p)p·ϕdx=Ωα(x)(p-p)ϕds(x), 4.8
Ω((P[su]+Bsut):sψ)dx-Ω(p(χ+ρ(1-χ))+β(QR(θ+)-θc))divψdx=Ωg·ψdx, 4.9
Ω(CV(θ)t-Bsut:sut-DP[su]t-μR(p)QR(|p|2)-(χ+ρ(1-χ))|D0[p]t|-γR(p,θ,divu)χt2+Lθcχt+βdivutQR(θ+))ζdx+Ωκ(QR(θ+))θ·ζdx=Ωω(x)(θ-θ)ζds(x), 4.10
γR(p,θ,divu)χt+I[0,1](χ)(1-ρ)ΦR(p)+pG0[p]-U0[p]+pdivu+LQR(θ+)θc-1a.\,e. 4.11

for all test functions ϕ,ζX and ψX0. For the system (4.8)–(4.11) the following result holds true.

Proposition 4.1

Let Hypothesis 3.1 hold and let R>1 be given. Then there exists a solution (p,u,θ,χ) to (4.8)–(4.11), (1.6) with the regularity

  • pLq(Ω;C[0,T]) for all q[1,6), pL2(Ω×(0,T);R3), ptL2(Ω×(0,T));

  • utL2(0,T;X0), sutL4(Ω×(0,T);Rsym3×3);

  • θL2(Ω×(0,T)), θL(0,T;L2(Ω;R3)), θtL2(Ω×(0,T));

  • χLq(Ω;C[0,T]), χtLq(Ω×(0,T)) for all q[1,).

We split the proof of Proposition 4.1 in two steps. First, in Sect. 4.1, we further regularize the system by means of a small parameter η>0 in order to obtain some extra-regularity for the gradient of the capillary pressure. Then, in Sect. 4.2, we solve this new problem by Galerkin approximations. Here the extra-regularization will be of fundamental importance in order to pass to the limit in the nonlinearity QR(|p(n)|2), where n is the dimension of the Galerkin scheme. As a last step, we let η0.

W2,2-regularization of the capillary pressure

We define the functions

MR(p):=0pμR(z)dz,KR(θ):=0θκ(QR(z+))dz 4.12

for p,θR, and introduce the new variables v=MR(p), z=KR(θ). We then choose another regularizing parameter η(0,1) and consider the following system in the unknowns v,u,z,χ:

Ω((χ+ρ(1-χ))(fR(MR-1(v))+G0[MR-1(v)]+divu))tϕdx+Ωv·ϕ+ηΔvΔϕdx=Ωα(x)(p-MR-1(v))ϕds(x), 4.13
Ω((P[su]+Bsut):sψ)dx-Ω(MR-1(v)(χ+ρ(1-χ))+β(QR((KR-1(z))+)-θc))divψdx=Ωg·ψdx, 4.14
Ω(CV(KR-1(z))t-Bsut:sut-DP[su]t-μR(MR-1(v))QR(|(MR-1(v))|2)-(χ+ρ(1-χ))|D0[MR-1(v)]t|-γR(MR-1(v),KR-1(z),divu)χt2+Lθcχt+βdivutQR((KR-1(z))+))ζdx+Ωz·ζdx=Ωω(x)(θ-KR-1(z))ζds(x), 4.15
γR(MR-1(v),KR-1(z),divu)χt+I[0,1](χ)(1-ρ)ΦR(MR-1(v))+MR-1(v)G0[MR-1(v)]-U0[MR-1(v)]+MR-1(v)divu+LQR((KR-1(z))+)θc-1a.\,e. 4.16

with test functions ϕW2,2(Ω), ψX0 and ζX.

Galerkin approximations

For each fixed R>1 and η(0,1), system (4.13)–(4.16) will be solved by Faedo–Galerkin approximations. To this end, let W=ei:i=0,1,2,L2(Ω) be the complete orthonormal systems of eigenfunctions defined by

-Δei=λieiinΩ,ei·n|Ω=0

with λ0=0, λi>0 for i1. Given nN, we approximate v and z by the finite sums

v(n)(x,t)=i=0nvi(t)ei(x),z(n)(x,t)=k=0nzk(t)ek(x)

where the coefficients vi,zk:[0,T]R and u(n),χ(n) will be determined as the solution of the system

Ω((χ(n)+ρ(1-χ(n)))(fR(p(n))+G0[p(n)]+divu(n)))teidx+λi+ηλi2vi=Ωα(x)(p-p(n))eids(x), 4.17
Ω((P[su(n)]+Bsut(n)):sψdx-Ω(p(n)(χ(n)+ρ(1-χ(n)))+β(QR((θ(n))+)-θc))divψdx=Ωg·ψdx, 4.18
Ω(CV(θ(n))t-Bsut(n):sut(n)-DP[su(n)]t-μR(p(n))QR(|p(n)|2)-(χ(n)+ρ(1-χ(n)))|D0[p(n)]t|-γR(p(n),θ(n),divu(n))|χt(n)|2+Lθcχt(n)+βdivut(n)QR((θ(n))+))ekdx+λkzk=Ωω(x)(θ-θ(n))ekds(x), 4.19
γR(p(n),θ(n),divu(n))χt(n)+I[0,1](χ(n))(1-ρ)ΦR(p(n))+p(n)G0[p(n)]-U0[p(n)]+p(n)divu(n)+LQR((θ(n))+)θc-1a.\,e. 4.20

for i,k=0,1,,n and for all ψX0, and with p(n):=MR-1(v(n)), θ(n):=KR-1(z(n)). We prescribe the initial conditions

vi(0)=ΩMR(p0(x))ei(x)dx,u(n)(x,0)=u0(x),zk(0)=ΩKR(θ0(x))ek(x)dx,χ(n)(x,0)=χ0(x). 4.21

This is an ODE system coupled with a nonlinear PDE (4.18). It is nontrivial to prove that such a system admits a unique strong solution. We proceed as follows. For a given function wLr(Ω×(0,T)) consider the equation

ΩBsut(x,t):sψ(x)dx+ΩP[su](x,t):sψ(x)dx=Ωw(x,t)divψ(x)dx, 4.22

which is to be satisfied for every ψX0 a. e. in (0, T) together with an initial condition u(x,0)=u0(x), u0X0W1,r(Ω;Rn) and boundary condition u=0 on Ω.

Step 1. As an application of the Lr-regularity theory for elliptic systems in divergence form (see e. g. [21]) we can conclude that for every wLr(Ω×(0,T)) with some r[2,) the problem

ΩBsut(x,t):sψ(x)dx=Ωw(x,t)divψ(x)dx

has for a. e. t(0,T) a unique solution such that sut(·,t)Lr(Ω;Rsym3×3), and it holds

Ω|sut|r(x,t)dxCΩ|w|r(x,t)dxfor a.\,e.t(0,T).

Integrating over t(0,T) we obtain that sutLr(Ω×(0,T);Rsym3×3).

Step 2. Let us define the convex and closed subset Ur:={wLr(0,T;X0):swtLr(Ω×(0,T);Rsym3×3),w(x,0)=u0(x)a.\,e.}Lr(0,T;X0W1,r(Ω;R3)). Note that the trace operator is well defined on this space, so that the initial condition makes sense. Let u^Ur, and let u be the solution of the equation

ΩP[su^](x,t):sψ(x)dx+ΩBsut(x,t):sψ(x)dx=Ωw(x,t)divψ(x)dx,

the existence of which follows from Step 1. We prove that the mapping u^tut is a contraction with respect to a suitable norm.

Indeed, let u^1,u^2 be given, and let u1,u2 be the corresponding solutions. The difference u¯=u1-u2 is the solution of the equation

ΩBsu¯t(x,t):sψ(x)dx=-Ω(P[su^1]-P[su^2])(x,t):sψ(x)dx.

According to Step 1, we have

Ω|su¯t|r(x,t)dxCΩ|P[su^1]-P[su^2]|r(x,t)dxa.\,e. 4.23

By inequality (3.14) we have for a. e. (xt)

|P[su^1]-P[su^2]|(x,t)C0t|s(u^1-u^2)t|(x,τ)dτ

with a constant C>0. Hence, by Hölder’s inequality,

Ω|P[su^1]-P[su^2]|r(x,t)dxCΩ0t|s(u^1-u^2)t|(x,τ)dτrdxCtr-10tΩ|s(u^1-u^2)t|r(x,τ)dxdτ. 4.24

Now, set

W(t)=Ω|s(u1-u2)t|r(x,t)dx,W^(t)=Ω|s(u^1-u^2)t|r(x,t)dx.

It follows from (4.23) and (4.24) that

W(t)Ctr-10tW^(τ)dτ.

We now multiply both sides of the above inequality by e-Ctr, and after an integration over t[0,T] we obtain from the Fubini Theorem

0Te-CtrW(t)dt1p0Te-Cτr-e-CTrW^(τ)dτ1p0Te-CtrW^(t)dt.

This means that the mapping u^tut is a contraction in Lr(0,T;X0W1,r(Ω;R3)) with respect to the weighted norm

|||ut|||=0Te-CtrΩ|sut|r(x,t)dxdt1/r,

hence it has a unique fixed point which is a solution of (4.22).

Step 3. The mapping which with a right-hand side wLr(Ω×(0,T)) associates the solution utLr(0,T;X0W1,r(Ω;R3)) of (4.22) is Lipschitz continuous. Indeed, consider w1,w2 and the corresponding solutions u1,u2, and set as before w¯=w1-w2, u¯=u1-u2. As a counterpart of (4.23) we get

Ω|su¯t|r(x,t)dxCΩ|P[su1]-P[su2]|r+|w¯|r(x,t)dxa.\,e.,

and the computations as in (4.24) yield

Ω|su¯t|r(x,t)dxCtr-10tΩ|su¯t|r(x,τ)dxdτ+CΩ|w¯|r(x,t)dxa.\,e.

We obtain the Lipschitz continuity result when we test by e-Cptr and integrate over t[0,T], similarly as in Step 2.

Now, coming back to our Eq. (4.18), we see that it is of the form (4.22) with w(x,t)=w(n)(x,t):=p(n)(χ(n)+ρ(1-χ(n)))+β(QR((θ(n))+)-θc)(x,t)+G^(x,t), where G^ is from Hypothesis 3.1 (ii). Therefore, denoting by S:w(n)ut(n) its associated solution operator, we conclude that (4.17)–(4.20) give rise to a system of ODEs with a locally Lipschitz continuous right-hand side containing the operator S.

Thus system (4.17)–(4.20) has a unique strong solution in a maximal interval of existence [0,Tn][0,T]. This interval coincides with the whole [0, T], provided we prove that the solution remains bounded in [0,Tn).

We now derive a series of estimates. Note that we decompose the auxiliary variables v and z instead of p and θ into a Fourier series with respect to the basis W because we are going to test Eqs. (4.17) and (4.19) by nonlinear expressions of p and θ, namely, by their Kirchhoff transforms (4.12). Indeed, the Galerkin method allows only to test by linear functions and their derivatives.

Moreover, we do not discretize the momentum balance equation because considering the full PDE is the only way to deduce compactness of the sequence {sut(n)}, which is needed in order to pass to the limit in some nonlinear terms. Indeed, we will not be able to control higher derivatives of u(n), and this will prevent us from applying the usual embedding theorems.

Estimates independent of n

Estimate 1. We test (4.17) by vi and sum up over i=0,1,,n, and (4.18) by ψ=ut(n). Then we sum up the two equations to obtain

Ω(χ(n)+ρ(1-χ(n)))fR(p(n))tMR(p(n))dx+Ω(χ(n)+ρ(1-χ(n)))G0[p(n)]tMR(p(n))dx+ΩBsut(n):sut(n)+UP[su(n)]t+DP[su(n)]tdx+Ω|v(n)|2+η|Δv(n)|2dx+Ωα(x)(p(n)-p)MR(p(n))ds(x)=-Ω(1-ρ)χt(n)fR(p(n))+G0[p(n)]+divu(n)MR(p(n))dx+Ω(χ(n)+ρ(1-χ(n)))divut(n)p(n)-MR(p(n))dx+Ωβ(QR((θ(n))+)-θc)divut(n)dx+Ωg·ut(n)dx. 4.25

where we exploited also the energy identity (3.10). We now define

VM,R(p):=0pfR(z)MR(z)dz

so that

Ω(χ(n)+ρ(1-χ(n)))fR(p(n))tMR(p(n))dx=ddtΩ(χ(n)+ρ(1-χ(n)))VM,R(p(n))dx-Ω(1-ρ)χt(n)VM,R(p(n))dx,

and introduce the modified Preisach potential as a counterpart to (2.20)

UM,R[p]:=00fr[p]MR(v)ψ(r,v)dvdr>0

which satisfies

G0[p]tMR(p)-UM,R[p]t0a.\,e.

according to (2.19). Note that (4.2) and (4.34.4) together with Hypothesis 3.1 (vi) and (vii) yield

cRp2VM,R(p)CRp2 4.26

for all pR, with some positive constants cR,CR depending only on R. Moreover, the estimate

UM,R[p]CR1+|p|2 4.27

holds as a counterpart of (2.18). By the definition of v(n) and Hypothesis 3.1 (vii) we deduce

Ω|v(n)|2dx=Ω|μR(p(n))|2|p(n)|2dx(μ)2Ω|p(n)|2dx. 4.28

Moreover, thanks again to Hypothesis 3.1 (vii), the boundary term is such that

Ωα(x)(p(n)-p)MR(p(n))ds(x)=Ωα(x)|p(n)|2MR(p(n))ds(x)-Ωα(x)pp(n)MR(p(n))ds(x)μΩα(x)|p(n)|2ds(x)-CRΩα(x)|pp(n)|ds(x),

where for pR we set

MR(p):=MR(p)/pforp0,MR(0)forp=0.

Young’s inequality and Hypothesis 3.1 (iii), (iv) give

Ωα(x)(p(n)-p)MR(p(n))ds(x)μ2Ωα(x)|p(n)|2ds(x)-CR. 4.29

Moreover, by Hölder’s inequality and Hypothesis 3.1 (ii),

Ω(g·ut(n))(x,t)dxCΩ|ut(n)|2(x,t)dx1/22CcBB2Ω|sut(n)|2dx1/2C+B8Ω|sut(n)|2dx 4.30

where in the last line we used first Korn’s inequality (3.16) and then Young’s inequality. Neglecting some lower order positive terms on the left-hand side, exploiting estimates (4.6), (4.26) and (4.27), and the fact that ρ(χ+ρ(1-χ))1 for all χ[0,1], from (4.25) and the subsequent computations we obtain

ddtΩ((χ(n)+ρ(1-χ(n)))(VM,R(p(n))+UM,R[p(n)])+UP[su(n)])dx+Ω(|p(n)|2+η|Δv(n)|2+B2|sut(n)|2)dx+Ωα(x)|p(n)|2ds(x)CR1+Ω|χt(n)||p(n)|2+|χt(n)||divu(n)||p(n)|+|p(n)|2dx, 4.31

where we used also Hypothesis 3.1 (i), Young’s inequality and the pointwise inequality

|divu|2(x,t)3|su|2(x,t)for all(x,t)Ω×(0,T) 4.32

to absorb |divut(n)| on the left-hand side together with the term coming from (4.30).

We need to control |χt(n)|. To this aim note that (4.20) is of standard form, namely,

χt(n)+I[0,1](χ(n))F(n)

with

F(n)=(1-ρ)ΦR(p(n))+p(n)G0[p(n)]-U0[p(n)]+p(n)divu(n)+LQR((θ(n))+)/θc-1γR(p(n),θ(n),divu(n)), 4.33

or, equivalently,

χ(n)[0,1],(F(n)-χt(n))(χ(n)-χ~)0a.\,e.χ~[0,1].

This yields, thanks to (2.18), (4.6) and (4.7),

|χt(n)(x,t)|C(1+|p(n)|2)+|divu(n)|2/2+LQR((θ(n))+)/θc-1γ1+(QR((θ(n))+)+(|p(n)|2-R2)++|divu(n)|2CR 4.34

for a. e. (x,t)Ω×(0,Tn). We now come back to (4.31) and integrate in time 0τdt for some τ[0,Tn]. The initial conditions are kept under control thanks to (3.7), (3.9), (4.26), (4.27) and Hypothesis 3.1 (v). Hence Young’s inequality yields

Ω|p(n)|2+|su(n)|2(x,τ)dx+0τΩ|p(n)|2+η|Δv(n)|2+|sut(n)|2(x,t)dxdt+0τΩα(x)|p(n)|2(x,t)ds(x)dtCR1+0τΩ|p(n)|2+|divu(n)|2(x,t)dxdt.

Using (4.32) and Grönwall’s lemma, we thus obtain

supessτ(0,Tn)Ω|p(n)|2+|su(n)|2(x,τ)dxCR, 4.35
0TnΩ(|p(n)|2+|sut(n)|2)(x,t)dx+Ωα(x)|p(n)|2(x,t)ds(x)dtCR, 4.36

and also

0TnΩ|Δv(n)|2(x,t)dxdtCRη. 4.37

Now, in terms of the variable v(n)=MR(p(n)), the boundary condition is nonlinear. By the spatial W2,2-regularity result for parabolic equations with nonlinear boundary conditions on C1,1 domains stated and proved in [14, Theorem 4.1], we see that

MR(p(n))L2(0,Tn;W2,2(Ω))2CRη. 4.38

This, together with the Sobolev embedding W1,2(Ω)L6(Ω), yields

0TΩ|p(n)|6(x,t)dx1/3dtCRη,

and since W1,6(Ω)C(Ω¯) we also get

0TsupessxΩ|p(n)|2(x,t)dtCRη. 4.39

Estimate 2. We test (4.17) by v˙i and sum up over i=0,1,,n. We get

Ω((χ(n)+ρ(1-χ(n)))(fR(p(n))+G0[p(n)]+divu(n)))tMR(p(n))tdx+Ωv(n)·vt(n)+ηΔv(n)Δvt(n)dx=Ωα(x)(p-p(n))MR(p(n))tds(x). 4.40

Defining

μ^R(p):=0pMR(z)zdz=0pμR(z)zdz

for pR, we can rewrite

Ωα(x)(p(n)-p)MR(p(n))tds(x)=Ωα(x)(μ^R(p(n))-pMR(p(n)))tds(x)+Ωα(x)ptMR(p(n))ds(x).

Hence, computing the time derivative in the first summand and rearranging the terms, we can rewrite (4.40) as

ddtΩ12|v(n)|2+η|Δv(n)|2dx+Ωα(x)(μ^R(p(n))-pMR(p(n)))ds(x)+Ω(χ(n)+ρ(1-χ(n)))fR(p(n))t+G[p(n)]tMR(p(n))tdx=-Ω(1-ρ)χt(n)fR(p(n))+G0[p(n)]+divu(n)MR(p(n))tdx-Ω(χ(n)+ρ(1-χ(n)))divut(n)MR(p(n))tdx-Ωα(x)ptMR(p(n))ds(x). 4.41

Combining (2.8) with the identity (2.7) for the play, we see that it holds

G0[p(n)]tMR(p(n))t=G0[p(n)]tpt(n)μR(p(n))0.

Hence by (4.2), (4.34.4) and Hypothesis 3.1 (vi), (vii) we obtain the pointwise lower bound

(χ(n)+ρ(1-χ(n)))fR(p(n))t+G0[p(n)]tMR(p(n))tρfR(p(n))μR(p(n))|pt(n)|2CR|pt(n)|2.

We now integrate (4.41) in time 0τdt for some τ[0,Tn). Note that μ^R(p)μp2/2 for all pR. Hence, arguing as for estimate (4.29) with p(n)MR(p(n)) replaced by μ^R(p(n)), we obtain that the boundary term on the left-hand side is such that

Ωα(x)(μ^R(p(n))-pMR(p(n)))(x,τ)ds(x)μ4Ωα(x)|p(n)|2(x,τ)ds(x)-CR. 4.42

Concerning the initial conditions, we employ Hypothesis 3.1 (iv) and (v). Thus, exploiting also (4.28) and (4.34), we get

Ω|p(n)|2+η|Δv(n)|2(x,τ)dx+Ωα(x)|p(n)|2(x,τ)ds(x)+0τΩ|pt(n)|2(x,t)dxdtCR(1+0τΩ|divu(n)||pt(n)|+|p(n)||pt(n)|+|divut(n)||pt(n)|dxdt+0τΩα(x)|pt||p(n)|ds(x)dt).

Young’s inequality, Hypothesis 3.1 (iii) and (iv), and estimates (4.32), (4.35), (4.36) give

supessτ(0,Tn)Ω|p(n)|2(x,τ)dx+Ωα(x)|p(n)|2(x,τ)ds(x)CR, 4.43
supessτ(0,Tn)Ω|Δv(n)|2(x,τ)dxCRη, 4.44
0TnΩ|pt(n)|2(x,t)dxdtCR. 4.45

Estimate 3. We test (4.19) by z˙k and sum over k=0,1,,n. We obtain

Ω(CV(θ(n))t-Bsut(n):sut(n)-DP[su(n)]t-μR(p(n))QR(|p(n)|2)-(χ(n)+ρ(1-χ(n)))|D0[p(n)]t|-γR(p(n),θ(n),divu(n))|χt(n)|2+(Lθcχt(n)+βdivut(n))QR((θ(n))+))KR(θ(n))tdx+Ωz(n)·zt(n)dx=Ωω(x)(θ-θ(n))KR(θ(n))tds(x). 4.46

Defining

κ^R(θ):=0θKR(z)zdz=0θκ(QR(z+))zdz

for θR, we can rewrite

Ωω(x)(θ(n)-θ)KR(θ(n))tds(x)=Ωω(x)(κ^R(θ(n))-θKR(θ(n)))tds(x)+Ωω(x)θtKR(θ(n))ds(x).

Hence, rearranging the terms in (4.46), we obtain

ddt12Ω|z(n)|2dx+Ωω(x)(κ^R(θ(n))-θKR(θ(n)))ds(x)+ΩCV(θ(n))κ(QR((θ(n))+))|θt(n)|2dx=Ω(Bsut(n):sut(n)+DP[su(n)]t+μR(p(n))QR(|p(n)|2)+(χ(n)+ρ(1-χ(n)))|D0[p(n)]t|+γR(p(n),θ(n),divu(n))|χt(n)|2-(Lθcχt(n)+βdivut(n))QR((θ(n))+))κ(QR((θ(n))+))θt(n)dx-Ωω(x)θtKR(θ(n))ds(x).

We now integrate in time 0τdt for some τ[0,Tn). By Hypothesis 3.1 (viii) and (ix) it holds

Ω|z(n)|2dx(κ)2Ω|θ(n)|2dx,0τΩCV(θ(n))κ(QR((θ(n))+))|θt(n)|2dxdtcκ0τΩ|θt(n)|2dxdt.

Note also that κ^R(θ)κθ2/2 for all θR. Hence, using Young’s inequality as in (4.29) we obtain

Ωω(x)(κ^R(θ(n))-θKR(θ(n)))ds(x)κ2Ωω(x)|θ(n)|2ds(x)-CRΩω(x)|θθ(n)|2ds(x)κ4Ωω(x)|θ(n)|2ds(x)-CR.

Concerning the initial conditions, we employ Hypothesis 3.1 (iv) and (v). Thus, exploiting also (2.18), (3.11), (4.7) and (4.34), we get

Ω|θ(n)|2(x,τ)dx+Ωω(x)|θ(n)|2(x,τ)ds(x)+0τΩ|θt(n)|2(x,t)dtCR(1+0τΩ|sut(n)|2+|pt(n)|+|p(n)|2+|divu(n)|2|θt(n)|dxdt+0τΩω(x)|θt||θ(n)|ds(x)dt). 4.47

We see that the approximate solution remains bounded in the maximal interval of existence [0,Tn]. Hence the solution exists globally, and for every nN we have Tn=T.

We further need to estimate the terms sut(n), p(n), divu(n) in the norm of L4(Ω×(0,T)). Note that (4.43), (4.45) entail p(n)L(0,T;L2(Ω;R3)), pt(n)L2(Ω×(0,T)) independently of n. Thus by the anisotropic embedding formulas ([3, Theorem 10.2] on p. 143 of the Russian version, see also [1, 5]) we deduce

0TΩ|p(n)|4(x,t)dxdtCR. 4.48

Now, let us consider (4.18) rewritten in the form

ΩP[su(n)](x,t):sψ(x)dx+ΩBsut(n)(x,t):sψ(x)dx=Ωw(n)(x,t)divψ(x)dx, 4.49

where

w(n)(x,t):=p(n)(χ(n)+ρ(1-χ(n)))+β(QR((θ(n))+)-θc)(x,t)+G^(x,t)

according to Hypothesis 3.1 (ii). By the already mentioned Lr-regularity (with some r[2,)) for elliptic systems in divergence form and by (3.8) we deduce, arguing as for (4.24),

Ω|sut(n)|r(x,t)dxCΩ|su(n)|r(x,0)dx+Ctr-10tΩ|sut(n)|r(x,τ)dxdτ+CΩ|w(n)|r(x,t)dxa.\,e. 4.50

By (4.48) and Hypothesis 3.1 (ii) we see that

0τΩ|w(n)|4(x,t)dxdtCR1+0τΩ|p(n)|4(x,t)dxdtCR.

Therefore, choosing r=4 in (4.50) and using also Hypothesis 3.1 (v), by Grönwall’s lemma we obtain

0τΩ|sut(n)|4(x,t)dxdtCR,

from which we deduce a bound also for the term divu(n) since sut(n) is dominant. Thus, coming back to (4.47) and using Hypothesis 3.1 (iii), (iv) and estimate (4.45), we finally obtain

Ω|θ(n)|2(x,τ)dx+Ωω(x)|θ(n)|2(x,τ)ds(x)+0τΩ|θt(n)|2(x,t)dxdtCR(1+0τΩω(x)|θ(n)|2(x,t)ds(x)dt).

Applying Grönwall’s lemma and Poincaré’s inequality (3.17) we finally obtain the estimates

supessτ(0,T)Ω|θ(n)|2+|θ(n)|2(x,τ)dx+Ωω(x)|θ(n)|2(x,τ)ds(x)CR, 4.51
0TΩ|θt(n)|2(x,t)dxdtCR. 4.52

Limit as n

For the moment we keep the regularization parameters η and R fixed, and let n in (4.17)–(4.20). From estimates (4.35), (4.36), (4.38), (4.43), (4.44), (4.45), (4.51), (4.52), we see that there exists a subsequence of {(p(n),θ(n)):nN}, which is again indexed by n, and functions p,θ such that

pt(n)pt,θt(n)θtweakly inL2(Ω×(0,T)),θ(n)θweakly-star inL(0,T;L2(Ω;R3)),p(n)pstrongly inLq(Ω;C[0,T])forq[1,6)and inL2(Ω×(0,T)),p(n)pstrongly inL2(Ω×(0,T);R3),θ(n)θstrongly inL2(Ω×(0,T))and inL2(Ω×(0,T)),

where the strong convergences are obtained by compact embedding, see [3]. We also need strong convergence of the sequences {su(n)} and {sut(n)} in order to pass to the limit in some nonlinear terms. Taking the difference of (4.49) for indices n and m, and testing by ψ=ut(n)-ut(m) we obtain, arguing as for Step 3 of the existence part,

Ω|sut(n)-sut(m)|2(x,τ)dxCΩ|w(n)-w(m)|2(x,τ)dxCΩ(1+|p(n)-p(m)|2+|p(n)|2|χ(n)-χ(m)|2+|θ(n)-θ(m)|2)(x,τ)dx 4.53

a. e. in (0, T). The L1-Lipschitz continuity result for variational inequalities (see [11, Theorem 1.12]) tells us that

|χ(n)-χ(m)|(x,τ)0τ|χt(n)-χt(m)|(x,t)dt20τ|F(n)-F(m)|(x,t)dt 4.54

with the notation of (4.33), where we have by virtue of (2.16) for a. e. xΩ that

0τ|F(n)-F(m)|(x,t)dtCR(1+maxs[0,τ]|p(n)-p(m)|(x,s)+0τ(|divu(n)-divu(m)|+|θ(n)-θ(m)|)(x,t)dt).

For t[0,T] put

U(t)=Ω|su(n)-su(m)|2(x,t)dx,W(t)=Ω|sut(n)-sut(m)|2(x,t)dx,π(t)=1+supessxΩ|p(n)|2(x,t),Π(t)=0tπ(s)ds,y(t)=Ω1+maxs[0,t]|p(n)-p(m)|2(x,s)+|θ(n)-θ(m)|2(x,t)dx.

Note that by (4.39) Π is bounded above independently of n. We further have

U(t)C1+0tW(s)ds,

and (4.53) is of the form

W(τ)CRy(τ)+π(τ)0τy(t)+0tW(s)dsdt.

Thus, from Fubini’s theorem and Grönwall’s lemma we obtain

0TW(τ)dτCR0TeC(Π(T)-Π(τ))y(τ)+π(τ)0τy(t)dtdτCR(0TeC(Π(T)-Π(τ))y(τ)dτ+0TtTeC(Π(T)-Π(τ))π(τ)dτy(t)dt)CR,η0Ty(τ)dτ,

that is,

Ωsupessτ(0,T)|su(n)-su(m)|2(x,τ)dx0TΩ|sut(n)-sut(m)|2(x,τ)dxdτCR,η1+Ωmaxs[0,T]|p(n)-p(m)|2(x,s)+0T|θ(n)-θ(m)|2(x,τ)dτdx. 4.55

The sequences {p(n)} and {θ(n)} are Cauchy in L2(Ω;C[0,T]) and L2(Ω×(0,T)), respectively, hence {su(n)} and {sut(n)} are also Cauchy sequences in L2(Ω;C([0,T];Rsym3×3)) and in L2(Ω×(0,T);Rsym3×3), respectively. Thus we conclude

su(n)sustrongly inL2(Ω;C([0,T];Rsym3×3)),sut(n)sutstrongly inL2(Ω×(0,T);Rsym3×3).

This is enough to pass to the limit in the nonlinearities by virtue of [7, Theorem 12.10], and we obtain

fR(p(n))fR(p),ΦR(p(n))ΦR(p)strongly inL2(Ω×(0,T)),G0[p(n)]G0[p]strongly inL2(Ω;C[0,T]),G0[p(n)]tG0[p]tweakly inL2(Ω×(0,T)),|D0[p(n)]t||D0[p]t|weakly inL2(Ω×(0,T)),U0[p(n)]U0[p]strongly inL1(Ω;C[0,T]),P[su(n)]P[su]strongly inL2(Ω;C([0,T];Rsym3×3)),DP[su(n)]tDP[su]tweakly inL2(Ω×(0,T)),Bsut(n):sut(n)Bsut:sutstrongly inL1(Ω×(0,T)),CV(θ(n))CV(θ)strongly inLq(Ω×(0,T))for allq1,21+b^,QR(|p(n)|2)QR(|p|2)μR(p(n))μR(p)QR((θ(n))+)QR(θ+)1γR(p(n),θ(n),divu(n))1γR(p,θ,divu)strongly inLq(Ω×(0,T))for allq[1,),

from which

fR(p(n))tfR(p)t,CV(θ(n))tCV(θ)tweaklyinL2(Ω×(0,T)),γR(p(n),θ(n),divu(n))γR(p,θ,divu)stronglyinLq(Ω×(0,T))forallq[1,).

The convergence of the Preisach hysteresis terms G0[p(n)], U0[p(n)] and |D0[p(n)]t| follow from (2.16), (2.17) and (2.11), respectively. The convergence of the plasticity terms P[su(n)] and DP[su(n)]t follows from (3.12) and (3.10), respectively. We now prove that the sequences {χ(n)}, {χt(n)} converge strongly in appropriate function spaces. Inequality (4.54) yields

Ωsupτ[0,T]|χ(n)-χ(m)|(x,τ)dx0TΩ|χt(n)-χt(m)|(x,t)dxdt20TΩ|F(n)-F(m)|(x,t)dxdt

where, due to the above convergences, F(n)F strongly in L1(Ω×(0,T)). Hence we conclude that {χ(n)(x,t)} and {χt(n)(x,t)} are Cauchy sequences in L1(Ω;C[0,T]) and in L1(Ω×(0,T)), respectively. Moreover, since both |χ(n)| and |χt(n)| admit a uniform pointwise upper bound (see (4.34)), we can use the Lebesgue dominated convergence theorem to conclude that

χ(n)χstrongly inLq(Ω;C[0,T])for allq[1,),χt(n)χtstrongly inLq(Ω×(0,T))for allq[1,).

Passing to the limit as n in (4.17)–(4.20) we see that (p,u,θ,χ) is a solution to (4.13)–(4.16), (1.6) with the regularity stated in Proposition 4.1.

Limit as η0

Let us denote by (v(η),u(η),z(η),χ(η)) the solution to (4.13)–(4.16). Note that estimate (4.44) is preserved in the limit as n, hence the limit function v(η) satisfies the inequality

supessτ(0,T)Ω|Δv(η)|2(x,τ)dxCRη.

We now choose ϕW2,2(Ω) such that ϕ|Ω=0, and integrate by parts in equation (4.13) to obtain

Ω((χ(η)+ρ(1-χ(η)))(fR(MR-1(v(η)))+G0[MR-1(v(η))]+divu(η)))tϕdx+Ω-Δv(η)ϕ+ηΔv(η)Δϕdx=0. 4.56

Introducing the new variable v^(η)=Δv(η), we rewrite (4.56) in the form

Ωv^(η)ϕ-ηΔϕdx=Ωh(η)ϕdx 4.57

where

h(η)=((χ(η)+ρ(1-χ(η)))(fR(MR-1(v(η)))+G0[MR-1(v(η))]+divu(η)))t.

Note that the term G0[p(η)]t is of order pt(η) by (2.7) and (2.8). Hence by estimates (4.34), (4.35), (4.36), (4.45) we see that h(η)L2(Ω×(0,T)), and its L2-norm is bounded independently of η.

Consider now the system {e^k:kN} of eigenfunctions of the negative Laplace operator with zero Dirichlet boundary conditions

-Δe^k=νke^k,e^k|Ω=0,Ω|e^k(x)|2dx=1.

They form a complete orthonormal system in L2(Ω) with 0<ν1ν2ν3. The functions h(η),v^(η) admit the expansions

h(η)(x,t)=k=1hk(η)(t)e^k(x),v^(η)(x,t)=k=1v^k(η)(t)e^k(x),

with coefficients hk(η):[0,T]R, v^k(η):[0,T]R. Choosing ϕ=e^k in (4.57) we obtain

v^k(η)(t)=hk(η)(t)1+ηνk,

hence

0TΩ|v^(η)(x,t)|2dxdt=0Tk=1|v^k(η)(t)|2dt0Tk=1|hk(η)(t)|2dt=0TΩ|h(η)(x,t)|2dxdtCR 4.58

for some positive constant CR independent of η. The estimate (4.45) is preserved in the limit n. Thus we get for a subsequence η0 that

v(η)vstrongly inL2(Ω×(0,T);R3),ηΔv(η)0strongly inL2(Ω×(0,T)),

from which, by definition of MR in (4.12) and Hypothesis 3.1 (vii),

MR-1(v(η))MR-1(v)stronglyinL2(Ω×(0,T);R3),QR(|MR-1(v(η))|2)QR(|MR-1(v)|2)stronglyinLq(Ω×(0,T))forallq[1,).

Also the estimates (4.35), (4.36), (4.43), (4.51), (4.52) are preserved when n. Since they are independent of η, by letting η0 we obtain for v(η) and θ(η) the same convergences as before.

Note that as a side product, from (4.58) we get that the estimate

0TΩ|ΔMR(p)|2(x,t)dxdtCR

holds also in the limit as η0. Hence, arguing as for (4.38) we get

MR(p)L2(0,T;W2,2(Ω))2CR.

This, by Sobolev embedding, yields

0TΩ|p|6(x,t)dx1/3dtCR, 4.59

as well as

0TsupessxΩ|p(n)|2(x,t)dtCR.

But then we can argue as in Sect. 4.4 and obtain an inequality similar to (4.55), but with a constant independent of η. This entails the strong convergence of the sequences su(η) and sut(η). The rest of the convergence argument follows exactly as at the end of Sect. 4.4, and this concludes the proof of Proposition 4.1.

Estimates independent of R

We now come back to our cut-off system (4.8)–(4.11). We are going to derive a series of estimates independent of R. More precisely, after proving that the temperature stays away from zero, we will perform the energy estimate and the Dafermos estimate in order to gain some regularity for the temperature. Subsequently, a key-step will be the derivation of a bound for p in an anisotropic Lebesgue space. Then an analogous estimate based on the particular structure of Eq. (4.9) is obtained for sut. We finally show that this is sufficient for starting the Moser iteration and obtain an L bound for p. After deriving some higher order estimates for the capillary pressure and for the temperature, we will be ready to let R tend to in (4.8)–(4.11).

Positivity of the temperature

For every nonnegative test function ζX we have, by virtue of (4.10),

ΩCV(θ)tζ+κ(QR(θ+))θ·ζdx+Ωω(x)(θ-θ)ζds(x)=Ω(Bsut:sut+DP[su]t+μR(p)QR(|p|2)+(χ+ρ(1-χ))|D0[p]t|+γR(p,θ,divu)χt2-Lθcχt+βdivutQR(θ+))ζdxΩB3|divut|2+γχt2-(Lθcχt+βdivut)QR(θ+)ζdx,

where in the last line we used Hypothesis 3.1 (i) together with inequality (4.32), and also estimate (4.7). Then, by Young’s inequality,

ΩCV(θ)tζ+κ(QR(θ+))θ·ζdx+Ωω(x)(θ-θ)ζds(x)-CΩ(QR(θ+))2ζdx

with a constant C depending on L,θc,β,B,γ. Let now φ(t) be the solution of the ODE

ddtCV(φ(t))+Cφ2(t)=0,φ(0)=θ¯

with θ¯ from Hypothesis 3.1. Then φ is nondecreasing and positive. Taking into account the fact that CV(φ)t=-Cφ2 and φ=0, for every nonnegative test function ζX we have in particular

Ω(CV(φ)-CV(θ))tζ+κ(QR(θ+))(φ-θ)·ζdx+Ωω(x)(θ-θ)ζds(x)CΩ(QR(θ+))2-φ2ζdx.

Consider now the following regularization of the Heaviside function

Hε(z)=0forz0,zεfor0<zε,1forz>ε,

for ε>0, and set ζ(x,t)=Hε(φ(t)-θ(x,t)) which is an admissible test function. This yields

Ω(CV(φ)-CV(θ))tHε(φ-θ)dx0.

By the Lebesgue Dominated Convergence Theorem we can pass to the limit in the above inequality for ε0, getting

Ω(CV(φ)-CV(θ))tH(φ-θ)dx0,

that is, by the monotonicity of CV,

ddtΩCV(φ)-CV(θ)+dx0,(CV(φ)-CV(θ))+(x,0)=0

which implies (CV(φ)-CV(θ))+0. Owing again to the monotonicity of CV and φ, we conclude that, independently of R,

θ(x,t)φ(t)φ(T)=:θT>0for allxandt. 5.1

We now pass to a series of estimates independent of R.

Energy estimate

Since we proved that the temperature stays positive, from now on we will write QR(θ+)=QR(θ).

We test (4.8) by ϕ=p, (4.9) by ψ=ut and (4.10) by ζ=1. Summing up the three resulting equations we obtain

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpdx+ΩμR(p)|p|2dx+Ω(P[su]+Bsut):sutdx-Ω(p(χ+ρ(1-χ))+β(QR(θ)-θc))divutdx+Ω(CV(θ)t-Bsut:sut-DP[su]t-μR(p)QR(|p|2)-(χ+ρ(1-χ))|D0[p]t|-γR(p,θ,divu)χt2)dx+Ω(Lθcχt+βdivut)QR(θ)dx=Ωα(x)(p-p)pds(x)+Ωg·utdx+Ωω(x)(θ-θ)ds(x).

Note that some of the terms cancel out. Moreover, recalling the notation introduced in () and the energy balance (2.11), the identities

Ω((χ+ρ(1-χ))fR(p))tpdx=ddtΩ(χ+ρ(1-χ))VR(p)dx+Ω(1-ρ)χtΦR(p)dx, 5.2
Ω((χ+ρ(1-χ))G0[p])tpdx-Ω(χ+ρ(1-χ))|D0[p]t|dx=ddtΩ(χ+ρ(1-χ))U0(p)dx+Ω(1-ρ)χtpG0[p]-U0[p]dx 5.3

hold true. Hence we obtain, using also (3.10) and (4.11),

ddtΩ(CV(θ)+Lχ+βθcdivu+(χ+ρ(1-χ))(VR(p)+U0[p])+UP[su])dx+ΩμR(p)(|p|2-QR(|p|2))dx+Ω(α(x)(p-p)p+ω(x)(θ-θ))ds(x)=ddtΩg·udx-Ωgt·udx. 5.4

We now integrate in time 0τdt. On the left-hand side Young’s inequality, (3.11) and (4.32) entail

UP[su]+βθcdivuA8|su|2-C. 5.5

By the definition of QR in (4.1), it holds |p|2QR(|p|2). The boundary term is such that

0τΩ(α(x)(p-p)p+ω(x)(θ-θ))ds(x)dt0τΩα(x)p22+ω(x)θds(x)dt-C 5.6

thanks to Young’s inequality and Hypothesis 3.1 (iii) and (iv). Concerning the right-hand side of (5.4), the time integration gives

Ω(g·u)(x,τ)dx-Ωg(x,0)·u0(x)dx-0τΩ(gt·u)(x,t)dxdt,

where the term containing the initial conditions is controlled by using Hölder’s inequality and observing that

Ω|g|2(x,0)dx=Ω|g|2(x,τ)dx-20τΩ(g·gt)(x,t)dxdt.

Hence by Young’s inequality and Hypothesis 3.1 (ii), (v) we deduce

Ω(g·u)(x,τ)dx-Ωg(x,0)·u0(x)dx-0τΩ(gt·u)(x,t)dxdtA16Ω|su|2(x,τ)dx+C1+0τΩ|su|2(x,t)dxdt,

where we used also Korn’s inequality (3.16). The first term in the last line is absorbed by (5.5). Finally, the initial conditions are kept under control thanks to (2.18), (3.9), (4.6) and Hypothesis 3.1 (v). Hence what we eventually get is

ΩCV(θ)+VR(p)+|su|2(x,τ)dx+0τΩ(α(x)p2+ω(x)θ)(x,t)ds(x)dtC1+0τΩ|su|2(x,t)dxdt,

and applying Grönwall’s lemma we finally obtain the estimates

supessτ(0,T)ΩCV(θ)+VR(p)+|su|2(x,τ)dxC, 5.7
0TΩ(α(x)p2+ω(x)θ)(x,t)ds(x)dtC. 5.8

Estimate (5.7) also gives

supessτ(0,T)Ω|θ|1+b(x,τ)dxC, 5.9

where b is from Hypothesis 3.1 (viii).

Dafermos estimate

We set θ^:=QR(θ) and test (4.10) by ζ=-θ^-a, with a from Hypothesis 3.1. This yields the identity

Ω(CV(θ)t-Bsut:sut-DP[su]t-μR(p)QR(|p|2)-(χ+ρ(1-χ))|D0[p]t|-γR(p,θ,divu)χt2+Lθcχt+βdivutθ^)(-θ^-a)dx+Ωκ(θ^)θ·(-θ^-a)dx=Ωω(x)(θ-θ)(-θ^-a)ds(x). 5.10

It holds

ΩCV(θ)t(-θ^-a)dx=-ddtΩFa(θ)dx

where

Fa(θ):=0θcV(s)(QR(s))ads,

and by Hypothesis 3.1 (ix) also

Ωκ(θ^)θ·(-θ^-a)dx=Ωκ(θ^)aθ^-a-1θ·θ^dxaκΩ|θ^|2dx.

Hence from (5.10) we get, using also Hypothesis 3.1 (i) and inequalities (4.32), (4.7),

Ω(B3|divut|2+γχt2)θ^-adx+aκΩ|θ^|2dxΩLθcχt+βdivutθ^1-adx+Ωω(x)(θ-θ)θ^-ads(x)+ddtΩFa(θ)dx, 5.11

where we neglected some positive terms on the left-hand side. Young’s inequality yields

Lθcχt+βdivutθ^1-aγ2χt2+B4|divut|2θ^-a+Cθ^2-a,

with a constant C depending only on L,θc,β,B,γ, whereas the boundary term is such that

Ωω(x)(θ-θ)θ^-ads(x)C1+Ωω(x)θds(x)

by (5.1) and Hypothesis 3.1 (iii), (iv). Note also that Fa(θ)CV(θ) for all θ0. Thus, integrating (5.11) in time 0τdt for some τ[0,T] and neglecting some other positive terms on the left-hand side we obtain

0τΩ|θ^|2(x,t)dxdtC1+0τΩθ^2-a(x,t)dxdt 5.12

thanks to estimates (5.7), (5.8). Now, owing to estimate (5.9), we can apply the Gagliardo–Nirenberg inequality (3.18) with the choices s=1+b, r=2 and N=3 obtaining, for t(0,T),

|θ^(t)|qC1+|θ^(t)|2δ

with δ=6(q-1-b)(5-b)q and for every 1+b<q<6. In particular, since δ·(5-b)q3(q-1-b)=2, this and (5.12) yield

0T|θ^(t)|qq(5-b)/3(q-1-b)dtC1+0T|θ^(t)|22dtC1+0T|θ^(t)|2-a2-adt. 5.13

Let us now choose q=2-a, which is admissible in the sense that 1+b<2-a thanks to Hypothesis 3.1. Since 5-b3(1-a-b)>1, we can apply Young’s inequality on the right-hand side getting

0T|θ^(t)|2-a2-adtC.

Substituting in (5.13) entails

0TΩ|θ^|2(x,t)dxdtC. 5.14

Coming back to (5.13) again and choosing q=8/3+2b/3, we also get

0TΩθ^8/3+2b/3(x,t)dxdtC. 5.15

Mechanical energy estimate

In order to estimate the capillary pressure in a suitable anisotropic Lebesgue space, we first need to find a bound for divut in L2(Ω×(0,T)), independently of R. To this purpose, we test (4.8) by ϕ=p, (4.9) by ψ=ut and sum up to obtain, with the notation of the previous subsection,

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpdx+ΩμR(p)|p|2dx+Ω(P[su]+Bsut):sutdx-Ω(p(χ+ρ(1-χ))+β(θ^-θc))divutdx=Ωα(x)(p-p)pds(x)+Ωg·utdx.

Note that some terms cancel out. Owing to (5.2) and exploiting also the energy identity (3.10), what we eventually get is

ddtΩ((χ+ρ(1-χ))(VR(p)+U0[p])+UP[su])dx+ΩμR(p)|p|2dx+ΩBsut:sutdx+Ω(1-ρ)χt(ΦR(p)+pG0[p]+pdivu-U0[p])dx+Ωα(x)(p-p)pds(x)Ωβ(θ^-θc)divutdx+Ωg·utdx.

Now, (4.11) yields

(1-ρ)χt(ΦR(p)+pG0[p]+pdivu-U0[p])=γR(p,θ,divu)χt2-Lχtθ^θc-1=γR(p,θ,divu)χt2-γR(p,θ,divu)χtLγR(p,θ,divu)θ^θc-112γR(p,θ,divu)χt2-C(1+θ^) 5.16

where in the last line we used Young’s inequality and (4.7), and where the constant C is independent of R. Moreover, from the pointwise inequality (4.32) and arguing as for (4.30) we get

Ωβ(θ^-θc)divutdxC1+Ωθ^2dx+B4Ω|sut|2dx,Ωg·utdxC+B4Ω|sut|2dx.

Hence we obtain, exploiting also Hypothesis 3.1 (i) to absorb the terms coming from the two estimates above,

ddtΩ((χ+ρ(1-χ))(VR(p)+U0[p])+UP[su])dx+μΩ|p|2dx+B2Ω|sut|2dx+12ΩγR(p,θ,divu)χt2dx+12Ωα(x)p2ds(x)C1+Ωθ^2dx

where the boundary term was handled as in (5.6). We now integrate in time 0τdt for some τ[0,T]. The right-hand side is bounded thanks to estimate (5.15), whereas the initial conditions are kept under control thanks to (2.18), (3.7), (3.9), (4.6) and Hypothesis 3.1 (v). Hence, neglecting some already estimated positive terms, we finally obtain

0TΩ|p|2+|sut|2(x,t)dxdtC 5.17

independently of R. This, together with (5.8) and Poincaré’s inequality (3.17), yields

pL2(0,T;W1,2(Ω))2C. 5.18

Estimate for the capillary pressure

We choose an even function λ:R(0,) such that λ(p)0 for p>0 and pλ(p)X. Then we test (4.8) by ϕ=pλ(p). We obtain

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpλ(p)dx+ΩμR(p)(λ(p)+pλ(p))|p|2dx=Ωα(x)(p-p)pλ(p)ds(x). 5.19

The term under the time derivative has the form

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpλ(p)dx=Ω(1-ρ)χtfR(p)+G0[p]+divupλ(p)dx+Ω(χ+ρ(1-χ))fR(p)ptpλ(p)dx+Ω(χ+ρ(1-χ))G0[p]tpλ(p)dx+Ω(χ+ρ(1-χ))divutpλ(p)dx. 5.20

We now define

Vλ,R(p):=0pfR(z)zλ(z)dz

so that

Ω(χ+ρ(1-χ))fR(p)ptpλ(p)dx=ddtΩ(χ+ρ(1-χ))Vλ,R(p)dx-Ω(1-ρ)χtVλ,R(p)dx,

and introduce the modified Preisach potential as a counterpart to (2.20)

Uλ[p]:=00fr[p]vλ(v)ψ(r,v)dvdr

which satisfies

G0[p]tpλ(p)-Uλ[p]t0a.\,e.

according to (2.19). Note that Vλ,R(p)>0 and Uλ[p]0 for all p0. Then (5.20) can be rewritten as

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpλ(p)dxddtΩ(χ+ρ(1-χ))Vλ,R(p)+Uλ[p]dx+Ω(χ+ρ(1-χ))divutpλ(p)dx+Ω(1-ρ)χt(pfR(p)+pG0[p]+pdivuλ(p)-Vλ,R(p)-Uλ[p])dx. 5.21

Defining

Ψλ,R(p):=VR(p)λ(p)-Vλ,R(p),

we see that from () and (4.11) it holds

(1-ρ)χt(pfR(p)+pG0[p]+pdivuλ(p)-Vλ,R(p)-Uλ[p])=(1-ρ)χt(ΦR(p)+pG0[p]-U0[p]+pdivuλ(p)+Ψλ,R(p)+U0[p]λ(p)-Uλ[p])=γR(p,θ,divu)χt2-Lχtθ^θc-1λ(p)+(1-ρ)χtΨλ,R(p)+U0[p]λ(p)-Uλ[p].

Now, using Young’s inequality as in (5.16) we obtain

γR(p,θ,divu)χt2-Lχtθ^θc-1λ(p)12γR(p,θ,divu)χt2λ(p)-C(1+θ^)λ(p),

and similarly

|(1-ρ)χtΨλ,R(p)+U0[p]λ(p)-Uλ[p]|14γR(p,θ,divu)χt2λ(p)+CΨλ,R(p)+U0[p]λ(p)-Uλ[p]2γR(p,θ,divu)λ(p),

so that (5.21) entails

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpλ(p)dxddtΩ(χ+ρ(1-χ))Vλ,R(p)+Uλ[p]dx+Ω(χ+ρ(1-χ))divutpλ(p)dx+14ΩγR(p,θ,divu)χt2λ(p)dx-CΩ(1+θ^)λ(p)+Ψλ,R(p)+U0[p]λ(p)-Uλ[p]2γR(p,θ,divu)λ(p)dx. 5.22

Note that

VR(p)V(p)+f2(p2-R2)+

for all pR, hence

Ψλ,R(p)+U0[p]λ(p)-Uλ[p]21+p2γR(p,θ,divu)λ2(p)VR(p)+U0[p]2(1+p2)γR(p,θ,divu)V(p)+f2(p2-R2)++U0[p]2γ(1+p2)1+(p2-R2)+C

independently of R. From (5.22) we conclude

Ω((χ+ρ(1-χ))(fR(p)+G0[p]+divu))tpλ(p)dxddtΩ(χ+ρ(1-χ))Vλ,R(p)+Uλ[p]dx+14ΩγR(p,θ,divu)χt2λ(p)dx-CΩ1+|p||divut|+θ^+p2λ(p)dx,

so that (5.19) and a time integration 0τdt for some τ[0,T] yields

Ω(χ+ρ(1-χ))Vλ,R(p)+Uλ[p](x,τ)dx+0τΩμR(p)(λ(p)+pλ(p))|p|2(x,t)dxdt+0τΩα(x)(p-p)pλ(p)(x,t)ds(x)dtC0τΩ1+|p||divut|+θ^+p2λ(p)(x,t)dxdt+ΩVλ,R(p)+Uλ[p](x,0)dx. 5.23

Now that we got rid of χt and derived a manageable estimate, we choose λ(p)=|p|2k with kν/2 which will be specified later. Here ν is as in Hypothesis 3.1. Note that this is an admissible choice, that is, pλ(p)X. Indeed, by estimates (4.45), (4.59) and by the anisotropic embedding formulas, see [3], we have

pLq(0,T;C(Ω¯)) 5.24

for any q[1,4). The bound also depends on R, but for a fixed R and each k>0 the function p|p|2k(·,t) belongs to X for a. e. t(0,T).

With this choice (5.23) takes the form

Ω(χ+ρ(1-χ))(VRk(p)+Uk[p])(x,τ)dx+(1+2k)0τΩμR(p)|p|2k|p|2(x,t)dxdt+0τΩα(x)(p-p)p|p|2k(x,t)ds(x)dtC0τΩ|divut||p|1+2k+(1+θ^)|p|2k+|p||p|1+2k(x,t)dxdt+ΩVRk(p)+Uk[p](x,0)dx

where, from Hypothesis 3.1 (vi),

VRk(p)(x,t):=0p(x,t)fR(z)z|z|2kdz0pf(1+|z|)1+νz|z|2kdz0pf2max{1,|z|}1+νz|z|2kdz=f201z|z|2kdz+1pz|z|2k-1-νdzf1+2k-ν|p|1+2k-ν-C,Uk[p](x,t):=00fr[p](x,t)v|v|2kψ(r,v)dvdr0.

Note also that, from Hypothesis 3.1 (vi) and an analogous version of (2.3),

VRk(p)(x,0)=0p0(x)fR(z)z|z|2kdzf2+2k|p0(x)|2+2k,Uk[p](x,0)=00fr[p0](x)v|v|2kψ(r,v)dvdrCψ|p0(x)|2+2k.

Moreover

|p|2k|p|2=1(1+k)2|(p|p|k)|2.

Finally, by Young’s inequality with conjugate exponents 2+2k1+2k,2+2k, we see that the boundary term is such that

0τΩα(x)(p-p)p|p|2kds(x)dt=0τΩα(x)|p|2+2kds(x)dt-0τΩα(x)pp|p|2k(x,t)ds(x)dt12+2k0τΩα(x)|p|2+2kds(x)dt-12+2k0τΩα(x)|p|2+2kds(x)dt.

Hence, using also Hypothesis 3.1 (vii), we obtain

f1+2k-νΩ|p(x,τ)|1+2k-νdx+μ1+2k(1+k)20τΩ|(p|p|k)|2dxdt+12+2k0τΩα(x)|p|2+2kds(x)dt12+2k0τΩα(x)|p|2+2kds(x)dt+f2+2k+CψΩ|p0(x)|2+2kdx+C0τΩ|divut||p|1+2k+(1+θ^)(1+|p|)1+2k+|p||p|1+2kdxdt.

From Hypothesis 3.1 (iv) and (v) it follows that the above inequality is of the form

Ω|p(x,τ)|1+2k-νdx+0τΩ|(p|p|k)|2dxdt+0τΩα(x)|p|2+2kds(x)dtC(1+k)Ck+0τΩ|h~||p|1+2kdxdt 5.25

with a constant C1 independent of R and k and with

h~=1+θ^+|divut|+|p|. 5.26

We have h~L2(Ω×(0,T)) and

h~L2(Ω×(0,T))C

independently of R by virtue of the estimates (4.32) and (5.17) for divut, (5.15) for θ^ and (5.18) for p.

Since we deal with anisotropic spaces Lq(0,T;Lr(Ω)), qr, it is convenient to introduce for the norm of a function vLq(0,T;Lr(Ω)) the symbol

graphic file with name 30_2022_805_Equ131_HTML.gif 5.27

For the function

wk(x,t)=p(x,t)|p(x,t)|k 5.28

we obtain from (5.25) using Hölder’s inequality and Poincaré’s inequality (3.17) that

supessτ(0,T)|wk(τ)|sksk+0τ|wk(t)|1;22dtC(1+k)Ck+0T|wk(t)|qkqkdt1/2 5.29

for all τ[0,T], with

sk=1+2k-ν1+k,qk=2+4k1+k

and with a constant C independent of τ, R and k. We now show that for a suitably chosen k, the right-hand side of (5.29) is dominated by the left-hand side, which will imply a bound for the left-hand side. By the Gagliardo–Nirenberg inequality (3.18) with q=qk, s=sk, r=2 and N=3 we have

|wk(t)|qkC|wk(t)|sk1-δk|wk(t)|1;2δk,δk=1sk-1qk1sk-16. 5.30

We now choose k in such a way that δkqk=2, that is, 3qk=6+2sk, which yields

k=1-ν,sk=3(1-ν)2-ν,qk=6-4ν2-ν,qk(1-δk)=qk-2=2(1-ν)2-ν=23sk.

By Hypothesis 3.1 we have sk1. Hence, by (5.30),

0T|wk(t)|qkqkdtCsupessτ(0,T)|wk(τ)|sk(2/3)sk0T|wk(t)|1;22dt.

Since k<1, we conclude from (5.29) that there exists a constant C independent of R such that, in particular,

supessτ(0,T)|wk(τ)|skC,0T|wk(t)|qkqkdtC.

Invoking (5.28), we obtain for p the estimates

supessτ(0,T)|p(τ)|3(1-ν)C,0T|p(t)|6-4ν6-4νdtC. 5.31

We now distinguish two cases: ν1/3 and ν>1/3. For ν1/3 (that is, 3(1-ν)2) we have

supessτ(0,T)|p(τ)|2C. 5.32

For ν>1/3 (that is, 3(1-ν)<2) we use again the Gagliardo–Nirenberg inequality (3.18) with q=2, s=3(1-ν), r=2 and N=3, obtaining

|p(t)|2C|p(t)|3(1-ν)1-δ|p(t)|1;2δ,δ=3ν-11+ν,

so that for

qν=2(1+ν)3ν-1

we have

0T|p(t)|2qνdt1/qνCsupesst(0,T)|p(t)|3(1-ν)1-δ0T|p(t)|1;22dt1/qν.

Hence by virtue of (5.18), (5.27), and (5.31) we obtain

graphic file with name 30_2022_805_Equ137_HTML.gif 5.33

with a constant C>0 independent of R, according to the notation (5.27). Note that qν6 thanks to Hypothesis 3.1 with ν1/2.

Further estimates

Now that we have obtained a suitable estimate for the capillary pressure in (5.33), we derive an analogous estimate for divut. To this aim we test (4.9) by ψ=ut, which yields

Ω(Bsut:sut)(x,t)dxΩ-P[su]:sut+|p||divut|+β|θ^-θc||divut|+|g||ut|(x,t)dx.

By (3.8), Hypothesis 3.1 (v) and (5.17) we have

Ω|P[su]|2(x,t)dxC1+0TΩ|sut|2(x,τ)dxdτC,

hence using Hypothesis 3.1 (i) and Young’s inequality as in Sect. 5.4 we conclude that the estimate

Ω|sut|2(x,t)dxC1+Ωp2+θ^2(x,t)dx 5.34

holds for a. e. t(0,T) with a constant C>0 independent of R. We want to find and estimate for sut in the norm of Lq(0,T;L2(Ω)) for a suitable q. To this aim we apply the Gagliardo–Nirenberg inequality (3.18) to θ^ with the choices q=r=2, s=1+b (with b from Hypothesis 3.1) and N=3. We obtain that, for t(0,T),

|θ^(t)|2C|θ^(t)|1+b1-δ|θ^(t)|1;2δ,δ=3-3b5-b,

so that for

qb=2(5-b)3-3b

we have

0T|θ^(t)|2qbdt1/qbCsupesst(0,T)|θ^(t)|1+b1-δ0T|θ^(t)|1;22dt1/qb.

Hence by virtue of (5.9) and (5.14) we get

graphic file with name 30_2022_805_Equ139_HTML.gif 5.35

independently of R. Note that our hypotheses on b imply qb6. Thus, coming back to (5.34) we have obtained that there exists q:=min{qν,qb}6 such that, thanks to (5.32) or (5.33) and (5.35),

graphic file with name 30_2022_805_Equ140_HTML.gif 5.36

independently of R, according to the notation (5.27).

We continue our analysis with the inequality (5.25) again. Unlike in Sect. 5.5, we do not keep the exponent k bounded, but we let k in a controlled way. As in (5.28), we define auxiliary functions wk=p|p|k and rewrite (5.25) for k1-ν as

Ω|wk(x,τ)|2akdx+0τΩ|wk|2dxdt+0τΩγ(x)|wk|2ds(x)dt(1+k)maxC1+k,0τΩ|h~||wk|2bkdxdt 5.37

with a constant C1 independent of k, with h~ given by (5.26), and with

ak=1+2k-ν2+2k,bk=1+2k2+2k.

It follows from (5.32) or (5.33), (5.35), and (5.36) that h~L6(0,T;L2(Ω)) and

graphic file with name 30_2022_805_Equ336_HTML.gif

Repeating exactly the argument of the proof of [8, Proposition 6.2] we obtain the following result.

Proposition 5.1

Let Hypothesis 3.1 hold and let (p,u,θ,χ) be a solution of (4.8)–(4.11) with the regularity from Proposition 4.1. Then the function p admits an L-bound independent of R, more precisely,

|p(x,t)|C(ν¯H)σ/(ν¯(σ-1))σσ/(ν¯(σ-1)2)=:Rσ 5.38

for a. e. (x,t)Ω×(0,T) with σ=19/18, Inline graphic, and with positive constants ν¯,C depending only on the data.

The main consequence of Proposition 5.1 is that, since we aim at taking the limit as R in (4.8)–(4.11), we can restrict ourselves to parameter values R>Rσ, with Rσ from (5.38), so that the cut-off (4.2), (), (4.34.4) is never active and γR(p,θ,divu)=γ(θ^,divu). Hence we can rewrite (4.8)–(4.11) in the form

Ω((χ+ρ(1-χ))(f(p)+G0[p]+divu))tϕdx+Ω1ρWμ(p)p·ϕdx=Ωα(x)(p-p)ϕds(x), 5.39
Ω(P[su]+Bsut):sψdx-Ω(p(χ+ρ(1-χ))+β(θ^-θc))divψdx=Ωg·ψdx, 5.40
ΩCV(θ)tζ+κ(θ^)θ·ζdx+Ωω(x)(θ-θ)ζds(x)=Ω(Bsut:sut+DP[su]t+1ρWμ(p)QR(|p|2)+(χ+ρ(1-χ))|D0[p]t|+γ(θ^,divu)χt2-Lθcχt+βdivutθ^)ζdx, 5.41
γ(θ^,divu)χt+I[0,1](χ)(1-ρ)Φ(p)+pG0[p]-U0[p]+pdivu+Lθ^θc-1a.\,e. 5.42

for all test functions ϕ,ζX, ψX0, with θ^=QR(θ) and with initial conditions (1.6). In order to pass to the limit as R, we still need to derive some higher order estimates.

Higher order estimates for the capillary pressure

Let us define

M(p):=0pμ(z)dz 5.43

for pR, so that μ(p)p=M(p). We would like to test (5.39) by ϕ=M(p)t=μ(p)pt which, however, is not an admissible test function since ptX. Hence we choose a small h>0 and test by ϕ=1hM(p)(t)-M(p)(t-h), where

ϕ(x,t)=1hM(p)(t)-M(p)(t-h)(x):=1hM(p(x,t))-M(p(x,t-h)),

with the intention to let h0. We obtain

Ω((χ+ρ(1-χ))(f(p)+G0[p]+divu))t1hM(p)(t)-M(p)(t-h)dx+Ω1ρWM(p)·1hM(p)(t)-M(p)(t-h)dx=Ωα(x)(p-p)1hM(p)(t)-M(p)(t-h)ds(x). 5.44

Concerning the second summand on the left-hand side of (5.44), note that

M(p)(x,t)·1hM(p)(x,t)-M(p)(x,t-h)12h|M(p)|2(x,t)-|M(p)|2(x,t-h).

We now deal with the boundary term. It holds

p(x,t)1hM(p)(x,t)-M(p)(x,t-h)=1h(p(x,t)M(p)(x,t)-p(x,t-h)M(p)(x,t-h))-1hp(x,t)-p(x,t-h)M(p)(x,t-h),

To handle the term p1hM(p)(t)-M(p)(t-h), we use the inequality F(y)-F(z)F(y)(y-z) which holds for every convex function F and every yz. We interpret M(p)(t) as y, M(p)(t-h) as z and F(y)=M-1(M(p(t)))=M-1(y). The function M-1 is increasing, hence its antiderivative F is convex. Thus

pM(p)(t)-M(p)(t-h)M(p)(t-h)M(p)(t)M-1(z)dz=p(t-h)p(t)ξM(ξ)dξ.

Defining

μ^(p):=0pzμ(z)dz

for pR, we obtain

p(x,t)1hM(p)(x,t)-M(p)(x,t-h)1hμ^(p)(x,t)-μ^(p)(x,t-h).

Thus (5.44) and the above estimates entail

Ω(χ+ρ(1-χ))(f(p)t+G0[p]t)(x,t)1hM(p)(x,t)-M(p)(x,t-h)dx+12ρWΩ1h(|M(p)|2(x,t)-|M(p)|2(x,t-h))dx+Ωα(x)1hμ^(p)(x,t)-μ^(p)(x,t-h)ds(x)-Ωα(x)1h(p(x,t)M(p)(x,t)-p(x,t-h)M(p)(x,t-h))ds(x)-Ω(1-ρ)χt(f(p)+G0[p]+divu)(x,t)1hM(p)(x,t)-M(p)(x,t-h)dx-Ω(χ+ρ(1-χ))divut(x,t)1hM(p)(x,t)-M(p)(x,t-h)dx-Ωα(x)1hp(x,t)-p(x,t-h)M(p)(x,t-h)ds(x).

We are now ready to integrate in time from h to some τ(0,T) and then let h0. Note that estimate (4.45) entails that the function M(p)t=μ(p)pt is in L2, so that the convergence is strong in L2. We obtain

0τΩ(χ+ρ(1-χ))(f(p)t+G0[p]t)μ(p)ptdxdt+12ρWΩ|M(p)|2(x,τ)dx-12ρWΩ|M(p0)|2(x)dx+Ωα(x)μ^(p)-pM(p)(x,τ)ds(x)-Ωα(x)μ^(p0)(x)-p(x,0)M(p0)(x)ds(x)-0τΩ(1-ρ)χt(f(p)+G0[p]+divu)μ(p)ptdxdt-0τΩ(χ+ρ(1-χ))divutμ(p)ptdxdt-0τΩα(x)ptM(p)ds(x)dt.

Combining (2.8) with the identity (2.7) for the play, we see that it holds μ(p)G0[p]tpt0, thus

(χ+ρ(1-χ))f(p)t+G0[p]tμ(p)ptρμf2max{1,Rσ}1+ν|pt|2

thanks to Hypothesis 3.1 (vi), (vii) and estimate (5.38). Hence there exists a constant c>0 such that for every t(0,T) we have

c0τΩ|pt|2(x,t)dxdt+(μ)22ρWΩ|p|2(x,τ)dx+μ2Ωα(x)p2(x,τ)ds(x)C(1+0τΩ|χt|(1+|divu|)|pt|+|divut||pt|dxdt+0τΩα(x)|pt||p|ds(x)dt)

thanks to Hypothesis 3.1 (v), (vi) and (vii), where we handled the boundary term on the left-hand side as in (4.42). Arguing as for estimate (4.34), we obtain

|χt(x,t)|(1-ρ)Φ(p)+pG0[p]-U0[p]+pdivu+Lθ^/θc-1γ(θ^,divu)C 5.45

for a. e. (x,t)Ω×(0,T), this time independently of R thanks to (5.38). Thus, employing also Young’s inequality, estimates (5.7), (5.8), (5.17) and Hypothesis 3.1 (iii) and (iv), we conclude that

supessτ(0,T)Ω|p|2(x,τ)dx+Ωα(x)p2(x,τ)ds(x)C, 5.46
0TΩpt2(x,t)dxdtC. 5.47

By (5.7), (5.17), (5.18), (5.47) and by comparison in equation (5.39), we see that the term ΔM(p) is bounded in L2(Ω×(0,T)) independently of R. In terms of the new variable p~=M(p), the boundary condition in (1.7) is nonlinear, and from considerations similar to those used in the proof of [14, Theorem 4.1] it follows

M(p)L2(0,T;W2,2(Ω))2C. 5.48

We thus may employ the Gagliardo–Nirenberg inequality (3.18) with s=r=2, N=3 obtaining

|M(p)(t)|qC|M(p)(t)|2+|M(p)(t)|21-δ|ΔM(p)(t)|2δ,δ=312-1q,

which holds for all 2<q<6. Elevating to some power s such that δs=2 and integrating in time yield

0T|p(t)|qsdtCforq(2,6)and1q+23s=12,

thanks to estimates (5.46), (5.48) and Hypothesis 3.1 (vii). In particular, for s=4,q=3 and s=q=103 we obtain, respectively,

graphic file with name 30_2022_805_Equ153_HTML.gif 5.49

according to the notation (5.27).

Higher order estimates for the displacement

Let us consider Eq. (5.40). Setting

w(x,t):=p(χ+ρ(1-χ))+β(θ^-θc)(x,t)+G^(x,t)

and arguing as for (4.50) we deduce

Ω|sut|r(x,t)dxCΩ|su0|r(x)dx+Ctr-10tΩ|sut|r(x,τ)dxdτ+CΩ|w|r(x,t)dxa.\,e. 5.50

Thus, by choosing r=8/3+2b/3 (with b[1/2,1) from Hypothesis 3.1) in the above inequality we obtain from (5.15), (5.38), Hypothesis 3.1 (ii), (v) and Grönwall’s lemma

sut8/3+2b/3C. 5.51

We now derive an estimate for sut in a suitable anisotropic Lebesgue space. To this aim we need to derive first an additional estimate for θ^. We use Gagliardo–Nirenberg inequality (3.18) with the choices s=1+b, r=2 and N=3 obtaining, for t(0,T),

|θ^(t)|qC|θ^(t)|1+b+|θ^(t)|1+b1-δ|θ^(t)|2δ,δ=11+b-1q6(1+b)5-b,

which holds for all 1+b<q<6. This yields, elevating to some power s such that δs=2 and integrating in time,

0T|θ^(t)|qsdtCforq(1+b,6)and1q+5-b3s(1+b)=11+b

thanks to estimates (5.9), (5.14). In particular, for q=12(1+b)7+b and s=4 we obtain

graphic file with name 30_2022_805_Equ156_HTML.gif 5.52

Note that

12(1+b)7+b<83+2b3-7<b<2b>5,

which is certainly true under our hypotheses. Therefore, choosing r=12(1+b)7+b in (5.50) we obtain, thanks to (5.51) and Hypothesis 3.1 (v),

Ω|sut|12(1+b)/(7+b)(x,t)dxC1+Ω|w|12(1+b)/(7+b)(x,t)dx

for a. e. t(0,T). Hypothesis 3.1 (ii) and estimates (5.38), (5.52) then yield

graphic file with name 30_2022_805_Equ157_HTML.gif 5.53

independently of R.

Higher order estimates for the temperature

Note that (5.40) with ψ=ut and (5.42) entail, respectively,

ΩBsut:sutdx=-ΩP[su]:sutdx+Ω(p(χ+ρ(1-χ))+β(θ^-θc))divutdx+Ωg·utdx,γ(θ^,divu)χt2=(1-ρ)χt(Φ(p)+pG0[p]-U0[p]+pdivu)+Lχtθ^θc-1.

Plugging these identities into (5.41) we obtain

ΩCV(θ)tζ+κ(θ^)θ·ζdx+Ωω(x)(θ-θ)ζds(x)=Ω(-P[su]:sut+(p(χ+ρ(1-χ))-βθc)divut+g·ut+DP[su]t+1ρWμ(p)QR(|p|2)+(χ+ρ(1-χ))|D0[p]t|+χt((1-ρ)(Φ(p)+pG0[p]-U0[p]+pdivu)-L))ζdx=:ΩΓ(x,t)ζdx 5.54

for every ζX, where Γ(x,t) has the regularity of the worst term. Estimates (5.38), (5.45), (3.16), (3.8), (3.11), (2.18) yield

|Γ|C(1+|su|2+|sut|2+|p|2+|pt|),

which from (5.47), (5.49), (5.51), (5.53) implies

graphic file with name 30_2022_805_Equ159_HTML.gif 5.55

independently of R, with b as in Hypothesis 3.1.

Assume now that for some p08/3+2b/3 we have proved

θ^p0C. 5.56

We know that this is true for p0=8/3+2b/3 by virtue of (5.15). Set

r0=1+b4+bp0, 5.57

and set ζ=θ^r0 in (5.54). We obtain

ΩCV(θ)tθ^r0+κ(θ^)θ·θ^r0dx+Ωω(x)(θ-θ)θ^r0ds(x)=ΩΓθ^r0dx. 5.58

It holds

ΩCV(θ)tθ^r0dx=ddtΩFr0(θ)dx

where

Fr0(θ):=0θcV(s)(QR(s))r0ds.

Observe that

Fr0(θ)θ^r0+1+br0+1+b

by Hypothesis 3.1 (viii). Moreover, Hypothesis 3.1 (ix) entails

Ωκ(θ^)θ·θ^r0dx=Ωκ(θ^)r0θ^r0-1θ·θ^dxr0κΩθ^r0+a|θ^|2dx.

Concerning the boundary term, we use Young’s inequality with exponents r0+1r0,r0+1, and obtain

Ωω(x)(θ-θ)θ^r0ds(x)Ωω(x)θ^r0+1ds(x)-Ωω(x)θθ^r0ds(x)Ωω(x)θ^r0+1ds(x)-r0r0+1Ωω(x)θ^r0+1ds(x)-1r0+1Ωω(x)(θ)r0+1ds(x)1r0+1Ωω(x)θ^r0+1ds(x)-C

by Hypothesis 3.1 (iii) and (iv). We now integrate (5.58) in time 0τdt for some τ[0,T]. Thanks to the choice (5.57) and Hölder’s inequality with exponents 4+b3,4+b1+b, the right-hand side is such that

0τΩΓθ^r0dxdt=0τΩΓ(θ^p0)(1+b)/(4+b)dxdtΓ(4+b)/3θ^p0r0C

by estimates (5.55), (5.56). Hence we have obtained

1r0+1+bΩθ^r0+1+b(x,τ)dx+r00τΩθ^r0+a|θ^|2(x,t)dxdt+1r0+10τΩω(x)θ^r0+1(x,t)ds(x)dtC. 5.59

We now denote

r=1+r0+a2,s=r0+1+br,v=θ^r

and rewrite (5.59) as

Ω|v|s(x,τ)dx+0τΩ|v|2(x,t)dxdtC(r0+1+b). 5.60

We now apply Gagliardo–Nirenberg inequality (3.18) to v(t), t(0,T), with r=2 and N=3. Choosing q in such a way that δq=2, that is, q=23s+2, and integrating in time from 0 to T we obtain

vqCsupt[0,T]|v(t)|s+supt[0,T]|v(t)|s(q-2)/qv22/qCsupt[0,T]|v(t)|s+v2.

Estimate (5.60) yields

supt[0,T]|v(t)|sC(r0+1+b)1/s,v2C(r0+1+b)1/2,

so that vqC(r0+1+b). Coming back to the variable θ^, we have proved that

θ^p1C(r0+1+b)forp1=rq=5(1+b)p03(4+b)+83+a+2b3.

We now proceed by induction according to the rule

pj+1=5(1+b)pj3(4+b)+83+a+2b3,rj=(1+b)pj(4+b).

We have the implication

pj<(8+3a+2b)(4+b)7-2bpj+1>pj.

Hence, the sequence {pj} is increasing and limjpj=(8+3a+2b)(4+b)7-2b. It follows that choosing p¯=pj for some j sufficiently large we obtain

r¯:=(1+b)p¯(4+b)>a^, 5.61

with a^ from Hypothesis 3.1 (ix), and using also (5.59) we obtain

θ^p¯+supesst(0,T)|θ^(t)|r¯+1+bC 5.62

with p¯ arbitrarily close to (8+3a+2b)(4+b)7-2b. We now come back to (5.54), which we test by ζ=θ (note that this is an admissible choice by Proposition 4.1). It holds

ΩCV(θ)tθ(x,t)dx=ΩcV(θ)θθt(x,t)dx=ddtΩ0θ(x,t)cV(s)sdsdx,

hence from Hypothesis 3.1 (ix) and (5.55) we obtain, after a time integration,

Ωθ2+b(x,τ)dx+0TΩκ(θ^)|θ|2(x,t)dxdt+0TΩω(x)θ2(x,t)ds(x)dtC1+θ(4+b)/(1+b). 5.63

Using the Gagliardo–Nirenberg inequality (3.18) again with q=4+b1+b, s=1+b (note that 1+b<4+b1+b<6 under our hypotheses), r=2 and N=3 we have that, for each fixed t(0,T),

|θ(t)|(4+b)/(1+b)C1+|θ(t)|2δ

with δ=6(3-b2-b)(5-b)(4+b) and where we used estimate (5.9). Raising to the power (4+b)/(1+b) and integrating 0Tdt we get

θ(4+b)/(1+b)C1+0TΩ|θ|2dxdtδ/2C1+0TΩκ(θ^)|θ|2dxdtδ/2.

Plugging this back into (5.63) and using Young’s inequality we deduce

Ωθ2+b(x,τ)dx+0TΩκ(θ^)|θ|2(x,t)dxdt+0TΩω(x)θ2(x,t)ds(x)dtC. 5.64

This enables us to derive an upper bound for the integral Ωκ(θ^)θ·ζdx, which we need for getting an estimate for θt from equation (5.54). By Hölder’s inequality and Hypothesis 3.1 (ix) we have that

Ω|κ(θ^)θ·ζ|dx=Ω|κ1/2(θ^)θ·κ1/2(θ^)ζ|dxCΩκ(θ^)|θ|2dx1/2Ωmax{1,θ^1+a^}|ζ|2dx1/2. 5.65

Let us now choose q^>1 such that (1+a^)q^=1+r¯+b, where r¯ is defined in (5.61). Note that such a q^ exists since 1+r¯+b>1+a^+b>1+a^. Defining

q:=2q^q^-1=2+2q^-1>2, 5.66

we get from Hölder’s inequality with conjugate exponents q^,q2 that

Ωθ^1+a^|ζ|2dxΩθ^1+r¯+bdx1/q^Ω|ζ|qdx2/qCΩ|ζ|qdx2/q

by virtue of (5.62). Inequality (5.65) then yields the bound

Ω|κ(θ^)θ·ζ|dxCΩκ(θ^)|θ|2dx1/2Ω|ζ|qdx1/q.

Hence, by (5.64),

0TΩ|κ(θ^)θ·ζ|dxdtCζL2(0,T;W1,q(Ω)).

From (5.55) it follows that testing with ζL2(0,T;W1,q(Ω)) is admissible, in the sense that the term Γζ is integrable. This is obvious if q3. For q<3 the space W1,q(Ω) is embedded in LqS(Ω) with

1qS=1q-13=16-12q^,

so that 34+b+1qS1. We thus obtain from (5.54) that

0TΩθtζdxdtCζL2(0,T;W1,q(Ω)). 5.67

Passage to the limit as R

In this section we conclude the proof of Theorem 3.3 by passing to the limit in (5.39)–(5.42) as R. Most of the convergences can be handled as at the end of Sect. 4.3, hence we focus here on the main differences.

Let Ri be a sequence such that Ri>Rσ, with Rσ as in (5.38), and let (p,u,χ,θ)=(p(i),u(i),χ(i),θ(i)) be solutions of (5.39)–(5.42) corresponding to R=Ri, with θ^=θ^(i)=QRi(θ(i)) and test functions ϕ,ζX, ψX0. Our aim is to check that at least a subsequence converges as i to a solution of (3.2)–(3.5) with test functions ϕX, ψX0 and ζXq.

First, for the capillary pressure p=p(i) we have the estimates (5.38), (5.46), (5.47), (5.48) and (5.49), which imply that, passing to a subsequence if necessary,

pt(i)ptweakly inL2(Ω×(0,T)),p(i)pstrongly inLq(Ω;C[0,T])for allq[1,),p(i)pstrongly inLq(Ω×(0,T);R3)for allq1,103,

by compact embedding. We easily show that

QRi(|p(i)|2)|p|2strongly inLq(Ω×(0,T);R3)for allq1,53. 6.1

Indeed, let ΩT(i)Ω×(0,T) be the set of all (x,t)Ω×(0,T) such that |p(i)(x,t)|2>Ri. By (5.49) we have

C0TΩ|p(i)(x,t)|10/3dxdtΩT(i)|p(i)(x,t)|10/3dxdt|ΩT(i)|Ri5/3,

hence |ΩT(i)|CRi-5/3. For q<53 we use Hölder’s inequality to get the estimate

0TΩQRi(|p(i)|2)-|p(i)|2qdxdt=ΩT(i)Ri-|p(i)|2qdxdtΩT(i)|p(i)|2qdxdtΩT(i)|p(i)|10/3dxdt3q/5|ΩT(i)|1-3q/5CRi-(5-3q)/3,

and (6.1) follows.

For the temperature θ=θ(i) we proceed in a similar way. By estimates (5.64) and (5.67) we obtain

θ(i)θweakly inL2(Ω×(0,T);R3),θt(i)θtweakly inL2(0,T;W-1,q(Ω)),θ(i)θstrongly inL2(Ω×(0,T)),

where for the last convergence we exploited [18, Theorem 5.1] and the embedding W-1,q(Ω)W-1,2(Ω) (recall that q>2). Furthermore, estimate (5.62) entails that θ^(i) are uniformly bounded in Lq(Ω×(0,T)) for every q<(8+3a+2b)(4+b)7-2b. Hence a similar argument as above yields that

θ^(i)=QRi(θ(i))θstronglyinLq(Ω×(0,T))forallq1,(8+3a+2b)(4+b)7-2b.

Estimate (5.48) and the Sobolev embeddings yield an inequality similar to (4.55), but with a constant independent of both η and R. This is enough to obtain that su(i)su, sut(i)sut strongly in L2(Ω;C([0,T];Rsym3×3)) and in L2(Ω×(0,T);Rsym3×3), respectively. The strong convergences χ(i)χ, χt(i)χt then follow as at the end of Sect. 4.2, as well as the convergence of the hysteresis terms.

Therefore the limit as i yields a solution to (3.2)–(3.5), and the proof of Theorem 3.3 is completed.

Funding Information

Open access funding provided by Austrian Science Fund (FWF).

Declarations

Conflict of interest

The authors declare that they have no conflicts of interests.

Footnotes

The support from the Austrian Science Fund (FWF) projects V662 and F65, from the GAČR Grant No. 20-14736S, and from the European Regional Development Fund, Project No. CZ.02.1.01/0.0/0.0/16_019/0000778 is gratefully acknowledged.

Publisher's Note

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

Contributor Information

Chiara Gavioli, Email: chiara.gavioli@tuwien.ac.at.

Pavel Krejčí, Email: Pavel.Krejci@cvut.cz.

References

  • 1.Adams RA. Anisotropic Sobolev inequalities. Časopis pro pěstování matematiky. 1988;113(3):267–279. doi: 10.21136/CPM.1988.108786. [DOI] [Google Scholar]
  • 2.Albers B, Krejčí P. Unsaturated porous media flow with thermomechanical interaction. Math. Meth. Appl. Sci. 2016;39(9):2220–2238. doi: 10.1002/mma.3635. [DOI] [Google Scholar]
  • 3.Besov, O.V., Il’in, V.P., Nikol’skii, S.M.: Integral Representations of Functions and Imbedding Theorems. Scripta Series in Mathematics. Halsted Press (John Wiley & Sons): New York-Toronto, Ont.-London; 1978 (Vol. I), 1979 (Vol. II). Russian version Nauka: Moscow; 1975
  • 4.Detmann B, Krejčí P, Rocca E. Solvability of an unsaturated porous media flow problem with thermomechanical interaction. SIAM J. Math. Anal. 2016;48:4175–4201. doi: 10.1137/16M1056365. [DOI] [Google Scholar]
  • 5.Edmunds DE, Edmunds RM. Embeddings of anisotropic Sobolev spaces. Arch. Ration. Mech. Anal. 1986;94:245–252. doi: 10.1007/BF00279865. [DOI] [Google Scholar]
  • 6.Flynn D, McNamara H, O’Kane JP, Pokrovskiĭ AV. Application of the Preisach model to soil-moisture hysteresis. In: Bertotti G, Mayergoyz I, editors. The Science of Hysteresis. Oxford: Academic Press; 2006. pp. 689–744. [Google Scholar]
  • 7.Fučík, S., Kufner, A.: Nonlinear Differential Equations. Studies in Applied Mechanics, vol 2, Amsterdam /Elsevier Scientific Publ. Co., New York (1980)
  • 8.Gavioli C, Krejčí P. On a viscoelastoplastic porous medium problem with nonlinear interaction. SIAM J. Math. Anal. 2021;53(1):1191–1213. doi: 10.1137/20M1340617. [DOI] [Google Scholar]
  • 9.Gilbarg D, Trudinger NS. Elliptic Partial Differential Equations of Second Order. Berlin: Springer; 1983. [Google Scholar]
  • 10.Krasnosel’skii MA, Pokrovskii AV. Systems with Hysteresis. Moscow: Nauka; 1983. [Google Scholar]
  • 11.Krejčí P. Hysteresis operators—a new approach to evolution differential inequalities. Comment. Math. Univ. Carolinae. 1989;33:525–536. [Google Scholar]
  • 12.Krejčí P. Hysteresis, Convexity and Dissipation in Hyperbolic Equations. Tokyo: Gakkōtosho; 1996. [Google Scholar]
  • 13.Krejčí P. The Preisach hysteresis model: Error bounds for numerical identification and inversion. Discrete Cont. Dyn. Syst. Ser. S. 2013;6:101–119. [Google Scholar]
  • 14.Krejčí P, Panizzi L. Regularity and uniqueness in quasilinear parabolic systems. Appl. Math. 2011;56(4):341–370. doi: 10.1007/s10492-011-0020-5. [DOI] [Google Scholar]
  • 15.Krejčí P, Rocca E, Sprekels J. Unsaturated deformable porous media flow with thermal phase transition. Math. Models Methods Appl. Sci. 2017;27(14):2675–2710. doi: 10.1142/S0218202517500555. [DOI] [Google Scholar]
  • 16.Ladyzhenskaya OA, Ural’tseva NN. Linear and Quasilinear Equations of Elliptic Type. Moscow: Nauka; 1973. [Google Scholar]
  • 17.Lemaitre J, Chaboche J-L. Mechanics of Solid Materials. Cambridge: Cambridge University Press; 1990. [Google Scholar]
  • 18.Lions JL. Quelques méthodes de résolution des problèmes aux limites non linéaires. Paris: Dunod; Gauthier-Villars; 1969. [Google Scholar]
  • 19.Nečas J, Hlaváček I. Mathematical Theory of Elastic and Elasto-Plastic Bodies: An Introduction. Studies in Applied Mechanics. New York: Elsevier Science Ltd; 1981. [Google Scholar]
  • 20.Showalter RE, Stefanelli U. Diffusion in poro-plastic media. Math. Methods Appl. Sci. 2004;27:2131–2151. doi: 10.1002/mma.541. [DOI] [Google Scholar]
  • 21.Simader CG. On Dirichlet’s Boundary Value Problem. Lecture Notes in Mathematics. New York: Springer; 1972. [Google Scholar]
  • 22.Visintin A. Models of Phase Transitions, Progress in Nonlinear Differential Equations and Their Applications. Boston: Birkhäuser; 1996. [Google Scholar]

Articles from Nonlinear Differential Equations and Applications are provided here courtesy of Springer

RESOURCES