Processing math: 68%
Research article Special Issues

Deformable porous media with degenerate hysteresis in gravity field

  • Hysteresis in the pressure-saturation relation in unsaturated porous media, which is due to surface tension on the liquid-gas interface, exhibits strong degeneracy in the resulting mass balance equation. Solutions to such degenerate equations have been recently constructed by the method of convexification even if the permeability coefficient depends on the hysteretic saturation. The model is extended here to the case that the solid matrix material is viscoelastic and that the system is coupled with a gravity-driven moisture flux. The existence of a solution is proved by compact anisotropic embedding involving Orlicz spaces with respect to the time variable.

    Citation: Chiara Gavioli, Pavel Krejčí. Deformable porous media with degenerate hysteresis in gravity field[J]. Mathematics in Engineering, 2025, 7(1): 35-60. doi: 10.3934/mine.2025003

    Related Papers:

    [1] Raimondo Penta, Ariel Ramírez-Torres, José Merodio, Reinaldo Rodríguez-Ramos . Effective governing equations for heterogenous porous media subject to inhomogeneous body forces. Mathematics in Engineering, 2021, 3(4): 1-17. doi: 10.3934/mine.2021033
    [2] Dmitrii Rachinskii . Bifurcation of relative periodic solutions in symmetric systems with hysteretic constitutive relations. Mathematics in Engineering, 2025, 7(2): 61-95. doi: 10.3934/mine.2025004
    [3] Luca Formaggia, Alessio Fumagalli, Anna Scotti . A multi-layer reactive transport model for fractured porous media. Mathematics in Engineering, 2022, 4(1): 1-32. doi: 10.3934/mine.2022008
    [4] Paola F. Antonietti, Chiara Facciolà, Marco Verani . Unified analysis of discontinuous Galerkin approximations of flows in fractured porous media on polygonal and polyhedral grids. Mathematics in Engineering, 2020, 2(2): 340-385. doi: 10.3934/mine.2020017
    [5] Daniel N. Riahi, Saulo Orizaga . On rotationally driven nonlinear inclined polymeric jet with gravity effect. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022014
    [6] Dheeraj Varma, Manikandan Mathur, Thierry Dauxois . Instabilities in internal gravity waves. Mathematics in Engineering, 2023, 5(1): 1-34. doi: 10.3934/mine.2023016
    [7] Daniele Del Santo, Francesco Fanelli, Gabriele Sbaiz, Aneta Wróblewska-Kamińska . On the influence of gravity in the dynamics of geophysical flows. Mathematics in Engineering, 2023, 5(1): 1-33. doi: 10.3934/mine.2023008
    [8] La-Su Mai, Suriguga . Local well-posedness of 1D degenerate drift diffusion equation. Mathematics in Engineering, 2024, 6(1): 155-172. doi: 10.3934/mine.2024007
    [9] David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036
    [10] Aleksandr Dzhugan, Fausto Ferrari . Domain variation solutions for degenerate two phase free boundary problems. Mathematics in Engineering, 2021, 3(6): 1-29. doi: 10.3934/mine.2021043
  • Hysteresis in the pressure-saturation relation in unsaturated porous media, which is due to surface tension on the liquid-gas interface, exhibits strong degeneracy in the resulting mass balance equation. Solutions to such degenerate equations have been recently constructed by the method of convexification even if the permeability coefficient depends on the hysteretic saturation. The model is extended here to the case that the solid matrix material is viscoelastic and that the system is coupled with a gravity-driven moisture flux. The existence of a solution is proved by compact anisotropic embedding involving Orlicz spaces with respect to the time variable.



    We pursue here the study started in [12,13] of degenerate diffusion in unsaturated porous media filled with liquid and gas. Here we additionally take into account the deformations of the solid matrix produced by the penetrating humidity, and the effects of gravity on fluid diffusion as a generalization of the Richards equation (see [24]). To be precise, our main modeling assumptions are the following:

    (a) The pressure-saturation relation exhibits hysteresis;

    (b) The solid skeleton is viscoelastic;

    (c) The permeability of the material depends on the moisture content;

    (d) A gravity-driven moisture flux takes part in the process.

    Our model is inspired by the experimental evidence about hysteresis in hydrogeology in, e.g., [2,16,20]. Figure 1 is taken from [20] and shows rate-independent hysteresis between the logarithm soil suction ψ, which is a decreasing function of the pressure, and the volumetric water content θ which, up to a linear transformation, can be identified with the saturation. The main characteristic of the hysteresis relation in Figure 1 is that at turning points where the pressure changes the orientation from decreasing to increasing or vice versa, the starting slope is horizontal, which makes the diffusion problem degenerate.

    Figure 1.  Typical experimental hysteresis dependence in porous media between the logarithm soil suction ψ and the volumetric water content θ.

    Mathematical investigation of diffusion problems with hysteresis goes back to A. Visintin's pioneering monograph [25] presenting basic methods of solving PDEs with hysteresis. However, two problems have remained unsolved until recently: the degeneracy at turning points and the dependence of the diffusion coefficient on the saturation. We have proved in [12] that the degeneracy in the case of constant permeability can be overcome using a method developed there to convexify the hysteresis relation. In [13], the convexification combined with a compact embedding theorem for anisotropic Orlicz spaces was shown to give a sufficient argument for the solvability also in the case of saturation-dependent permeability. Let us also mention some previous results on hysteresis dependence of the permeability coefficient in the non-degenerate case under additional regularization in time or space [4,5,9,26]. To our knowledge, only E. El Behi-Gornostaeva addressed in her thesis [8] the full problem of hysteresis-dependent saturation without regularization in the non-degenerate case and proposed a method for the existence proof.

    In this paper we extend the problem to the case of deformable porous media under gravity effects and full degeneracy of the saturation dependence. The process is driven by the mass conservation principle and the quasistatic mechanical equilibrium equation in the small deformations regime. We further simplify the system in order to make it mathematically tractable by assuming that shear stresses are negligible. The gravity term and the viscoelasticity of the deformable solid skeleton represent a new degree of complexity related to the fact that no obvious a priori upper bound for the solutions is available here. We show that such an estimate can be obtained using a hysteresis variant of the Moser iteration technique under an additional assumption on the admissible Preisach operators G, namely, that the Preisach density ϕ(x,r,v) in (3.6) of G decays sufficiently slowly at infinity, see Hypothesis 3.6. The proof of the existence theorem is then carried out by a convexity and compactness argument developed in [13].

    The structure of the present paper is the following. In Section 2 we set up the mathematical model for the phenomenon and show that it is consistent with the physical principles of mass and energy conservation. In Section 3 we recall the definitions of the main concepts including convexifiable Preisach operators, and state the main existence Theorem 3.7. In Section 4 we propose a time discretization scheme with time step τ>0 and derive estimates independent of τ. A uniform upper bound for the time-discrete approximation is established in Section 5 via Moser-type iterations. In Section 6 we show that, similarly to [13], we can derive the convexity estimate (6.26) for the time derivative of the pressure. As shown in the last Section 7, this is sufficient to let the time step τ tend to 0 and to prove that the limit is the desired solution.

    The physical quantities which appear in the model are listed below in Table 1.

    Table 1.  List of physical quantities.
    Symbol Quantity Physical dimension
    ρ liquid mass density [kgm3]
    θ volumetric water content []
    q liquid mass flux [kgm2s]
    σ stress tensor [kgms2]
    u displacement vector [m]
    μ bulk elasticity modulus [kgms2]
    γ bulk viscosity [kgms]
    p liquid pressure [kgms2]
    p0 standard pressure [kgms2]
    g0 gravity constant [ms2]
    κ permeability [s]
    b boundary permeability [sm]

     | Show Table
    DownLoad: CSV

    The fluid diffusion in a domain ΩR3 filled with a deformable unsaturated porous medium is driven by the liquid mass conservation principle

    ρθt=divq (2.1)

    with constant mass density ρ, and by the mechanical equilibrium equation for the solid

    divσ=0   in  Ω,σn(x)=0   on  Ω, (2.2)

    where n(x) denotes the unit outward normal vector to Ω at a point xΩ. We write (2.2) in variational form

    Ωσ:sξdx=0,ξW1,2(Ω;R3), (2.3)

    where s denotes the symmetric gradient.

    We assume that the deformations are small, so that divu is the relative local volume change, and that shear stresses can be neglected, that is,

    σ=(μdivu+γdivutp)δ, (2.4)

    where μ>0 is a constant bulk elasticity modulus, γ>0 is a bulk viscosity modulus, and δ=(δij), i,j=1,2,3 is the Kronecker symbol. Then (2.3) can be written as

    Ω(μdivu+γdivutp)divξdx=0,ξW1,2(Ω;R3). (2.5)

    Let hL2(Ω) be arbitrary, and let W be the solution to the problem

    ΔW=h,W=0  on  Ω. (2.6)

    Putting ξ=W in (2.5), we get the mechanical equilibrium equation in the form

    μdivu+γdivutp=0 (2.7)

    a.e. in Ω.

    Assume now that the relative liquid mass flux qρut is proportional to the liquid pressure gradient p, that is,

    qρut=κp, (2.8)

    with proportionality factor κ=κ(x,θ)>0 depending on xΩ and θ. The liquid pressure p can be decomposed into two components

    p=pc+ph, (2.9)

    where pc is the capillary pressure and ph is the hydrostatic pressure. For the hydrostatic pressure we assume the classical relation

    ph=ρg0ν(xx0), (2.10)

    where ν is the unit vector in the gravity direction and x0 is a referential point.

    We introduce the dimensionless normalized capillary pressure

    u=pcp0 (2.11)

    and follow the modeling hypotheses of [2,24] which consists in representing the pcθ hysteresis relation by the formula

    θ=G[u], (2.12)

    where G is a Preisach operator defined below in Section 3.

    On the boundary Ω of Ω we assume that the unit outward normal n(x) is defined almost everywhere and that the normal component of the relative flux qρut in (2.8) is proportional to the difference of pressures p inside and p outside the body, that is,

    κ(x,θ)pn(x)=b(x)(pp)   on  Ω. (2.13)

    The weak formulation of the mass balance Eq (2.1) then reads

    Ω(ρ(θt+divut)y+κ(x,θ)py)dx+Ωb(x)(pp)yds(x)=0 (2.14)

    for every test function yW1,2(Ω)L(Ω).

    Assume for the moment that there is no mass exchange with the exterior corresponding to the choice b(x)=0 in (2.13), that is,

    κ(x,θ)pn(x)=0 (2.15)

    on Ω. Putting y=1 in (2.14), we get

    ddtΩρ(θ+divu)(x,t)dx=0. (2.16)

    The interpretation of (2.16) is mass conservation. The term Ωdivu(x,t)dx describes the evolution of the volume represented by Ω, and ρΩ(θ+divu)(x,t)dx is the total water mass in Ω at time t. Naturally, if the volume increases and mass is conserved, then the saturation decreases and vice versa.

    Consider now the energy balance. Under the boundary condition (2.15), no power is supplied to the system, and the total energy of the system should therefore decrease because all phenomena like viscosity, diffusion, and hysteresis dissipate energy. We put y=p/ρ in (2.14) and obtain using (2.15) that

    Ω(θt+divut)pdx+Ωκ(x,θ)ρ|p|2dx=0. (2.17)

    For a Preisach operator G in (2.12) there exists a Preisach potential operator V and a dissipation operator D such that

    G[u]tu=V[u]t+|D[u]t|  a.e. (2.18)

    Using (2.7) and (2.9)–(2.11), we can rewrite (2.17) in the form

    ddtΩ(p0V[u]+ρg0θν(xx0)+μ2|divu|2)dx+Ω(p0|D[u]t|+κ(x,θ)ρ|p|2+γ|divut|2)dx=0. (2.19)

    The energetic interpretation of (2.19) is the following: The term

    Ω(p0|D[u]t|+κ(x,θ)ρ|p|2+γ|divut|2)dx (2.20)

    is the dissipation rate which is positive in agreement with the principles of thermodynamics, while

    Ω(p0V[u]+ρg0θν(xx0)+μ2|divu|2)dx (2.21)

    is the potential energy of the system which is therefore decreasing as expected.

    Putting v=divuph/μ, from (2.7) and (2.9)–(2.14) we get the following system, consisting of a PDE coupled with an ODE, for the unknowns u and v:

    Ω((G[u]+v)ty+κ(x,G[u])p0ρ(u+ρg0p0ν)y)dx+Ωp0b(x)ρ(uu)yds(x)=0, (2.22)
    γvt+μvp0u=0, (2.23)

    for every yW1,2(Ω)L(Ω), where we set u=(pph)/p0.

    The existence proof for (2.22)-(2.23) is independent of the exact values of the physical constants. We therefore normalize all constants to 1 for simplicity and consider the system

    Ω((G[u]+v)ty+κ(x,G[u])(u+ν)y)dx+Ωb(x)(uu)yds(x)=0, (2.24)
    vt+v=u, (2.25)

    for every yW1,2(Ω)L(Ω), coupled with the initial conditions

    u(x,0)=u0(x),v(x,0)=v0(x)  a.e. in  Ω. (2.26)

    We study here Problem (2.24)–(2.26) in a bounded Lipschitzian domain ΩRN and time interval (0,T), where NN can be arbitrary. Recall first the definition of the Preisach operator introduced in [21] in the variational setting of [18].

    Definition 3.1. Let λL(Ω×(0,)) be a given function with the following properties:

    Λ>0: λ(x,r)=0  for rΛ,xΩ, (3.1)
    ˉλ>0: |λ(x1,r1)λ(x2,r2)|(ˉλ|x1x2|+|r1r2|) r1,r2(0,),x1,x2Ω. (3.2)

    For a given r>0, we call the play operator with threshold r and initial memory λ the mapping which with a given function uL1(Ω;W1,1(0,T)) associates the solution ξrL1(Ω;W1,1(0,T)) of the variational inequality

    |u(x,t)ξr(x,t)|r,ξrt(x,t)(u(x,t)ξr(x,t)z)0  a.e. z[r,r], (3.3)

    with initial condition

    ξr(x,0)=λ(x,r)  a.e., (3.4)

    and we denote

    ξr(x,t)=pr[λ,u](x,t). (3.5)

    Given a measurable function ϕ:Ω×(0,)×R[0,) and a constant ˉGR, the Preisach operator G is defined as a mapping G:L2(Ω;W1,1(0,T))L2(Ω;W1,1(0,T)) by the formula

    G[u](x,t)=ˉG+ 0ξr(x,t)0ϕ(x,r,v)dvdr. (3.6)

    The Preisach operator is said to be regular if the density function ϕ of G in (3.6) belongs to L(Ω×(0,)×R), and there exist constants ϕ1,ˉϕ>0 and a decreasing function ϕ0:RR such that for all U>0, all x,x1,x2Ω, and a.e. (r,v)(0,U)×(U,U) we have

    0<ϕ0(U)<ϕ(x,r,v)<ϕ1, (3.7)
    |ϕ(x1,r,v)ϕ(x2,r,v)|ˉϕ|x1x2|. (3.8)

    In applications, the natural physical condition θ=G[u][0,1] is satisfied for each input function u if and only if ˉG(0,1) and

    00ϕ(x,r,v)dvdr1ˉG,00ϕ(x,r,v)dvdrˉG, (3.9)

    for a.e. xΩ. However, for the existence result in Theorem 3.7 only (3.7) and (3.8) are substantial.

    Let us mention the following classical result (see [18, Proposition Ⅱ.3.11]).

    Proposition 3.2. Let G be a regular Preisach operator in the sense of Definition 3.1. Then it can be extended to a Lipschitz continuous mapping G:Lp(Ω;C[0,T])Lp(Ω;C[0,T]) for every p[1,).

    The Preisach operator is rate-independent. Hence, for input functions u(x,t) which are monotone in a time interval t(a,b), a regular Preisach operator G can be represented by a superposition operator G[u](x,t)=B(x,u(x,t)) with an increasing function uB(x,u) called a Preisach branch. Indeed, the branches may be different at different points x and different intervals (a,b). The branches corresponding to increasing inputs are said to be ascending (the so-called wetting curves in the context of porous media), the branches corresponding to decreasing inputs are said to be descending (drying curves).

    Definition 3.3. Let U>0 be given. A Preisach operator is said to be uniformly counterclockwise convex on [U,U] if for all inputs u such that |u(x,t)|U a.e., all ascending branches are uniformly convex and all descending branches are uniformly concave in the sense that there exists β>0 such that for every u(U,U) we have the implications

    ut>02Bu2β,ut<02Bu2β

    in the sense of distributions, see [12].

    A regular Preisach operator G is called convexifiable if for every U>0 there exist a uniformly counterclockwise convex Preisach operator P on [U,U], positive constants g(U),g(U),ˉg(U), and a twice continuously differentiable mapping g:[U,U][U,U] such that

    g(0)=0,0<g(U)g(u)g(U),|g(u)|ˉg(U)  u[U,U], (3.10)

    and G=Pg.

    A typical example of a uniformly counterclockwise convex operator is the so-called Prandtl-Ishlinskii operator characterized by positive density functions ϕ(x,r) independent of v, see [18, Section 4.2]. Operators of the form Pg with a Prandtl-Ishlinskii operator P and an increasing function g are often used in control engineering because of their explicit inversion formulas, see [3,6,19]. They are called the generalized Prandtl-Ishlinskii operators (GPI) and represent an important subclass of Preisach operators. Note also that for every Preisach operator P and every Lipschitz continuous increasing function g, the superposition operator G=Pg is also a Preisach operator, and there exists an explicit formula for its density, see [17, Proposition 2.3]. Another class of convexifiable Preisach operators is shown in [12, Proposition 1.3].

    The technical hypotheses on the data in (2.24)-(2.25) can be stated as follows.

    Hypothesis 3.4. The boundary permeability b belongs to L(Ω), and is such that b(x)0 a.e. and Ωb(x)ds(x)>0. The boundary source u belongs to L(Ω×(0,T)); in addition, utL2(Ω×(0,T)). The permeability κ:Ω×RR is a bounded Lipschitz continuous function, more precisely, there exist constants ˉκ>0, κ1>κ0>0 such that for all θ,θ1,θ2R and all x,x1,x2Ω we have

    κ0κ(x,θ)κ1,|κ(x1,θ1)κ(x2,θ2)|ˉκ(|x1x2|+|θ1θ2|).

    Note that we can rewrite (2.24)-(2.25) for yW1,2(Ω)L(Ω) as

    Ω((G[u]t+uv)y+κ(x,G[u])(u+ν)y)dx+Ωb(x)(uu)yds(x)=0,vt+v=u.

    Therefore, taking into account (2.26), even a local solution to Problem (2.24)–(2.26) may fail to exist if for example λ(x,r)0 and div(κ(x,G[u](x,0))(u0(x)+ν))+v0(x)u0(x)0, and we need an initial memory compatibility condition which we state here following [12]. A more detailed discussion on this issue can be found in the introduction to [12].

    Hypothesis 3.5. Let the initial memory λ and the Preisach density function ϕ be as in Definition 3.1. The initial pressure u0 belongs to W2,(Ω), the initial volume strain v0 belongs to L(Ω), and there exist a constant L>0 and a function r0L(Ω) such that, for Λ>0 as in (3.1), supessxΩ|u0(x)|Λ and for a.e. xΩ the following initial compatibility conditions hold:

    λ(x,0)=u0(x)  a.e., (3.11)
    θ0(x)=G[u](x,0)=ˉG+ 0λ(x,r)0ϕ(x,r,v)dvdr  a.e., (3.12)
    1L|div(κ(x,θ0(x))(u0(x)+ν))+v0(x)u0(x)|r0(x)Λ  a.e., (3.13)
    rλ(x,r)sign(div(κ(x,θ0(x))(u0(x)+ν))+v0(x)u0(x))  a.e.  for r(0,r0(x)), (3.14)
    κ(x,θ0(x))(u0(x)+ν)n(x)=b(x)(u0(x)u(x,0))  a.e. <italic>on</italic>Ω. (3.15)

    Unlike [12], here we do not need to assume div(κ(x,θ0(x))(u0(x)+ν))L(Ω) since it follows from the fact that u0W2,(Ω) together with assumptions (3.2), (3.8), and Hypothesis 3.4 on κ. Instead, in addition to the hypotheses of [12], we have to assume a polynomial decay of the Preisach density function at infinity that will allow us to perform a hysteresis variant of the Moser iterations in Section 5.

    Hypothesis 3.6. The Preisach density ϕ:Ω×(0,)×R[0,) in (3.6) is such that

    m>0 ϕ0>0:ϕ(x,r,v)ϕ0max{1,r+|v|}m  a.e.  for  (r,v)(0,)×R. (3.16)

    Condition (3.16) is indeed compatible with (3.9) if m>2, as

    00max{1,r+v}mdvdr=0rmax{1,z}mdzdr=12+1m2,

    so that (3.9) is satisfied if for example ˉG=1/2 and ϕ012/m.

    Our main existence result reads as follows.

    Theorem 3.7. Let Hypotheses 3.4–3.6 hold, and let G be a convexifiable Preisach operator in the sense of Definition 3.3. Then there exists a solution (u,v) to Problem (2.24)–(2.26) such that u,vL(Ω×(0,T)), uL2(Ω×(0,T);RN), ut and θt=G[u]t belong to the Orlicz space LΦlog(Ω×(0,T)) generated by the function Φlog(v)=vlog(1+v), and vtL(Ω×(0,T)).

    Basic properties of Orlicz spaces are summarized in [13, Section 5]. For a more comprehensive discussion, we refer the interested reader to the monographs [1,22].

    We proceed as in [12], choose a discretization parameter nN, define the time step τ=T/n, and replace system (2.24)-(2.25) with the following time-discrete counterpart for the unknowns {ui,vi:i=1,,n}:

    Ω(1τ((G[u]iG[u]i1)+(vivi1))y+κ(x,G[u]i)(ui+ν)y)dx+Ωb(x)(uiui)yds(x)=0, (4.1)
    1τ(vivi1)+vi=ui, (4.2)

    for every test function yW1,2(Ω), where ui(x)=u(x,iτ) for i{1,,n}, u0 and v0 are as in (2.26). Here, the time-discrete Preisach operator G[u]i is defined for an input sequence {ui:iN{0}} by a formula of the form (3.6), namely,

    G[u]i(x)=ˉG+ 0ξri(x)0ϕ(x,r,v)dvdr, (4.3)

    where ξri denotes the output of the time-discrete play operator

    ξri(x)=pr[λ,u]i(x) (4.4)

    defined as the solution operator of the variational inequality

    |ui(x)ξri(x)|r,(ξri(x)ξri1(x))(ui(x)ξri(x)z)0,iN,  z[r,r], (4.5)

    with a given initial condition

    ξr0(x)=λ(x,r)  a.e., (4.6)

    similarly as in (3.3)-(3.4). Note that the discrete variational inequality (4.5) can be interpreted as weak formulation of (3.3) for piecewise constant inputs in terms of the Kurzweil integral, and details can be found in [10, Section 2].

    Proposition 4.1. System (4.1)-(4.2) with initial conditions u0 and v0 as in (2.26) admits at least one solution {ui,vi}W1,2(Ω) for each i{1,,n}.

    Proof. We proceed by induction, and assume that the sequences {ui,vi} have already been constructed for i=1,,j1. Note that there is no hysteresis in the passage from uj1 to uj, so that there exists a function Gj:Ω×RR, continuous and nondecreasing in the second variable, such that G[u]j=Gj(x,uj). Furthermore, by (4.2) we have

    vj(x)=11+τvj1(x)+τ1+τuj(x). (4.7)

    Hence, (4.1)-(4.2) can be interpreted as a quasilinear elliptic equation for the unknown u:=uj of the form

    Ω((1τGj(x,u)+11+τu)y+κ(x,Gj(x,u))(u+ν)y)dx+Ωb(x)(uuj)yds(x)=Ωajydx (4.8)

    for every yW1,2(Ω) with given functions ajL2(Ω).

    Let {ek:kN}W1,2(Ω) be an orthonormal basis of L2(Ω). It is convenient to choose eigenfunctions of the following problem, which is compatible with boundary conditions (2.13), namely

    κ0Ωekydx+Ωb(x)ekyds(x)=μkΩekydx

    with eigenvalues 0<μ1<μ2. For each mN we look for coefficients uk, k=1,,m, such that the function

    u(m)(x)=mk=1ukek(x)

    satisfies the identity

    Ω((1τGj(x,u(m))+11+τu(m))ekdx+κ(x,Gj(x,u(m)))(u(m)+ν)ek)dx+Ωb(x)(u(m)uj)ekds(x)Ωajekdx=0 (4.9)

    for every k=1,,m. The existence of a solution to (4.9) can be proved in a classical way using the Brouwer degree theory. An introduction to topological methods for solving nonlinear partial differential equations can be found in [11, Chapter Ⅴ], but we will mainly refer to [14], where detailed proofs are also provided.

    We begin by noting that we can define a continuous mapping T:RmRm associating to u:=(u1,,um) the left-hand side of (4.9) for k=1,,m. Let γ[0,1], and consider the homotopy

    Tγ(u)k=Ω(γτGj(x,u(m))+11+τu(m))ekdx+Ω((1γ)κ0+γκ(x,Gj(x,u(m))))(u(m)+γν)ekdx+Ωb(x)(u(m)γuj)ekds(x)γΩajekdx. (4.10)

    Testing (4.10) by uk and summing up over k=1,,m, by Hypothesis 3.4 we obtain

    Tγ(u)u=Ω(γτGj(x,u(m))u(m)+11+τ|u(m)|2)dx+Ω((1γ)κ0+γκ(x,Gj(x,u(m))))(|u(m)|2+γνu(m))dx+Ωb(x)(|u(m)|2γuju(m))ds(x)γΩaju(m)dxΩ(11+τ|u(m)|2+κ0|u(m)|2)dx+Ωb(x)|u(m)|2ds(x)Ω(1τ|Gj(x,u(m))|+|aj|)|u(m)|dx(κ0+κ1)Ω|u(m)|dxΩb(x)|uj||u(m)|ds(x).

    Hence, by Young's inequality and again Hypothesis 3.4, we see that Tγ(u)u>0 outside the ball B(m)RRm with a sufficiently large radius R independent of γ and m. Therefore, the equation Tγ(u)=0 has no solution on B(m)R, so that by [14, Theorem 3] the Brouwer degree d(Tγ,B(m)R,0) is a constant for γ[0,1]. The mapping T0 is linear, hence its degree is odd. In particular, d(T1,B(m)R,0)=d(T0,B(m)R,0)0, which implies (see [14, Theorem 2]) that the equation T1(u)=0, equivalently, (4.9), has at least one solution in B(m)R.

    Testing (4.9) by uk, summing up over k=1,,m, and employing Hypothesis 3.4 we see that the sequence {u(m)} is uniformly bounded in W1,2(Ω), hence it is compact in L2(Ω). Let u be the limit of any convergent subsequence of {u(m)}. Passing to the limit in (4.9) as m, we check that (4.8) holds with y=ek for every kN. Since the system {ek} is complete in L2(Ω), (4.8) holds for u=uj, and recalling (4.7) we conclude that system (4.1)-(4.2) has a solution.

    The situation is different from the case without gravity studied in [13], where the upper bound could be derived in an elementary way from Hilpert's inequality following [15]. Here, we propose a hysteresis variant of the Moser iteration procedure which is new to our knowledge. The price we pay is that we have to restrict the class of admissible Preisach densities ϕ(x,r,v) assuming that they have a sufficiently slow decay at infinity, see Hypothesis 3.6.

    We first test (4.1) by y=ui and get, using (4.2),

    Ω(1τ((G[u]iG[u]i1)ui+(vivi1)vi)+κ(x,G[u]i)(ui+ν)ui)dx+1τ2Ω|vivi1|2dx+Ωb(x)(uiui)uids(x)=0 (5.1)

    for all i{1,,n}. Let us define the functions

    ψ(x,r,ξ):=ξ0ϕ(x,r,v)dv,Ψ(x,r,ξ):=ξ0vϕ(x,r,v)dv. (5.2)

    In terms of the sequence ξri(x)=pr[λ,u]i we have

    G[u]i(x)=ˉG+0ψ(x,r,ξri(x))dr. (5.3)

    Choosing in (4.5) z=0 and using the fact that the function ψ in (5.2) is increasing, we obtain in both cases ξriξri1, ξriξri1 the inequalities

    (ψ(x,r,ξri)ψ(x,r,ξri1))ui(ψ(x,r,ξri)ψ(x,r,ξri1))ξriΨ(x,r,ξri)Ψ(x,r,ξri1), (5.4)

    and (5.1) yields, by Hypothesis 3.4 and the inequality (vivi1)vi(v2iv2i1)/2,

    1τΩ(0(Ψ(x,r,ξri)Ψ(x,r,ξri1))dr+v2iv2i12)dx+κ0Ω|ui|2dx+Ωb(x)|ui|2ds(x)C (5.5)

    for i{1,,n}, with a constant C>0 independent of τ. Summing up over i and exploiting the fact that, by the definition of ξr0 in (4.6) and the assumptions on ϕ and λ in Definition 3.1, we have

    0Ψ(x,r,ξr0(x))dr=Λ0λ(x,r)0vϕ(x,r,v)dvdrϕ12Λ0λ2(x,r)drC, (5.6)

    and we are assuming v0L(Ω), we get the estimate

    maxi=1,,nΩ|vi|2dx+τni=0(Ω|ui|2dx+Ωb(x)|ui|2ds(x))C (5.7)

    with a constant C>0 independent of τ.

    For parameters R>1 and k>1 to be specified later, we define the function

    Uk,R(u):={u|u|2k for  |u|R,(2k+1)R2ku2kR2k+1 for  u>R,(2k+1)R2ku+2kR2k+1 for  u<R. (5.8)

    The function Uk,R is odd, increasing, continuously differentiable, and its derivative has the form

    Uk,R(u)={(2k+1)|u|2k for  |u|R,(2k+1)R2k for  |u|>R. (5.9)

    We test (4.1) with y=Uk,R(ui), which is an admissible test function, and we obtain

    Ω(1τ(G[u]iG[u]i1)Uk,R(ui)+(vivi1)Uk,R(ui)+Uk,R(ui)κ(x,G[u]i)(ui+ν)ui)dx+Ωb(x)(uiui)Uk,R(ui)ds(x)=0. (5.10)

    For every α,βR and for every nondecreasing function h:RR, we have the elementary inequality

    α(h(α+β)h(β))0. (5.11)

    Formula (5.11) for α=(vivi1)/τ, β=vi, h=Uk,R together with (4.2) enables us to estimate

    (vivi1)Uk,R(ui)(vivi1)Uk,R(vi)Wk,R(vi)Wk,R(vi1), (5.12)

    where we put

    Wk,R(v)=v0Uk,R(z)dz0. (5.13)

    The next term in (5.10) can be estimate from below by means of Young's inequality

    Uk,R(ui)κ(x,G[u]i)(ui+ν)uiaUk,R(ui)|ui|2bUk,R(ui) (5.14)

    with positive constants a,b independent of R and k. To estimate the boundary term in (5.10) from below, we first notice that for all uR we have, by virtue of (5.8),

    |Uk,R(u)|(2k+2)/(2k+1)uUk,R(u)=|Uk,R(u)|1/(2k+1)|u|1. (5.15)

    Since u is bounded by Hypothesis 3.4, we may use (5.15) and Young's inequality with conjugate exponents 2k+2 and (2k+2)/(2k+1) to check that there exists a constant C>0 independent of R and k such that

    (uiui)Uk,R(ui)12uiUk,R(ui)C2k+2 (5.16)

    for all i{1,,n}.

    Finally, in order to estimate the time-discrete hysteresis term in (5.10), we define the function

    Ψk,R(x,r,ξ):=ξ0Uk,R(v)ϕ(x,r,v)dv. (5.17)

    Similarly as in (5.4), we have

    (ψ(x,r,ξri)ψ(x,r,ξri1))Uk,R(ui)(ψ(x,r,ξri)ψ(x,r,ξri1))Uk,R(ξri)Ψk,R(x,r,ξri)Ψk,R(x,r,ξri1). (5.18)

    Summing up in (5.10) over i=1,,n and using the above estimates (5.12), (5.14), (5.16), and (5.18), we get, using again an estimate of the initial step at i=0 similar to (5.6) and the assumption v0L(Ω),

    maxi=1,,nΩ0Ψk,R(x,r,ξri)drdx+aτni=1(ΩUk,R(ui)|ui|2dx+Ωb(x)uiUk,R(ui)ds(x))C(C2k+τni=1ΩUk,R(ui)dx) (5.19)

    with constants a>0 and C>0 independent of τ, k, and R.

    For uR put

    Vk,R(u):={u|u|k for  |u|R,(k+1)RkukRk+1 for  u>R,(k+1)Rku+kRk+1 for  u<R. (5.20)

    We have

    |Vk,R(ui)|2={(k+1)2|ui|2k|ui|2 for  |ui|<R,(k+1)2R2k|ui|2 for  |ui|R, (5.21)
    |Vk,R(ui)|2={|ui|2k+2 for  |ui|<R,(k+1)2R2k|ui|2+k2R2k+22k(k+1)R2k+1|ui|R2k+2 for  |ui|R. (5.22)

    Then

    Uk,R(ui)|ui|2=2k+1(k+1)2|Vk,R(ui)|2, (5.23)
    uiUk,R(ui)2k+1(k+1)2|Vk,R(ui)|2. (5.24)

    We define an equivalent norm of an element w from the Sobolev space W1,2(Ω) by the formula

    w1,2:=(aΩ|w|2dx+Ωb(x)|w|2ds(x))1/2

    and using (5.23)-(5.24) rewrite inequality (5.19) in the form

    maxi=1,,nΩ0Ψk,R(x,r,ξri)drdx+τk+1ni=1Vk,R(ui)21,2C(C2k+τni=1ΩUk,R(ui)dx). (5.25)

    We define

    wi(x):=min{R,|ui(x)|},u(k,R)i(x):=wki(x). (5.26)

    With this notation, by (5.9) we can rewrite the term at the right-hand side of (5.25) as

    ΩUk,R(ui)dx=(2k+1)Ω|u(k,R)i(x)|2dx. (5.27)

    Let us now compare the W1,2-norm of u(k,R)i with the W1,2-norm of Vk,R(ui) on the left-hand side of (5.25). We have

    |u(k,R)i|2={k2|ui|2k2|ui|2 for  |ui|<R0 for  |ui|R,|u(k,R)i|2={|ui|2k for  |ui|<RR2k for  |ui|R,

    and, by Young's inequality with conjugate exponents k and k/(k1),

    k2|ui|2k2k2(1k+k1k|ui|2k)k+(k+1)2|ui|2k.

    Hence, recalling (5.21)-(5.22), we estimate

    u(k,R)i21,2Vk,R(ui)21,2+kui21,2 (5.28)

    for i=1,,n. From (5.25), (5.27), and (5.7) we thus get

    maxi=1,,nΩ0Ψk,R(x,r,ξri)drdx+τk+1ni=1u(k,R)i21,2C0(C2k+(k+1)τni=1Ω|u(k,R)i|2dx), (5.29)

    for some constants C0,C>0 independent of τ, k, and R. To simplify the notation, we denote by ||q the norm in Lq(Ω). There exists a constant K>0 such that for every σ(0,1) the interpolation inequality

    |w|22K(σN|w|21+σ2w21,2) (5.30)

    holds for every wW1,2(Ω), see [7]. Choosing σ=(C0K)1/2(1+k)1, we apply this inequality in (5.29) and obtain

    maxi=1,,nΩ0Ψk,R(x,r,ξri)drdxC(C2k+(k+1)N+1τni=1(Ω|u(k,R)i|dx)2), (5.31)

    with a constant C>0 depending on C0 and K.

    So far, this has been a standard Moser argument. The crucial point in the derivation of an upper bound for the time-discrete problem is to find an efficient lower bound for the hysteresis term on the left-hand side of (5.31). By (5.17) we have for xΩ that

    0Ψk,R(x,r,ξri(x))dr=0ξri(x)0Uk,R(v)ϕ(x,r,v)dvdr, (5.32)

    where the function Uk,R is defined in (5.8). Assume first that ui(x)>0. By (4.5) for all r>0 we have

    ξri(x)0Uk,R(v)ϕ(x,r,v)dv(ξri(x))+0Uk,R(v)ϕ(x,r,v)dv(ui(x)r)+0Uk,R(v)ϕ(x,r,v)dv,

    where ()+ denotes the positive part. From (5.32) we thus get

    0Ψk,R(x,r,ξri(x))drui(x)0ui(x)r0Uk,R(v)ϕ(x,r,v)dvdr. (5.33)

    In the case ui(x)<0 we argue similarly and obtain the formula which is valid for both cases

    0Ψk,R(x,r,ξri(x))dr|ui(x)|0|ui(x)|r0Uk,R(v)ϕ(x,r,v)dvdrϕ0|ui(x)|0|ui(x)|r0Uk,R(v)max{1,r+v}mdvdr=:ϕ0Ak,R(|ui(x)|), (5.34)

    according to Hypothesis 3.6. Our goal is to show that the function

    Ak,R(u)=u0ur0max{1,r+v}mUk,R(v)dvdr=u0uv0max{1,r+v}mUk,R(v)drdv=u0Uk,R(v)uvmax{1,z}mdzdv, (5.35)

    of the argument u0, where we have first used Fubini's theorem and then substituted z=r+v, is dominant for k>(m3)/2 over the function

    Bk,R(u):=min{R,u}2k+3m,

    that is, there exists a constant Ck independent of R and possibly dependent on k such that

    Bk,R(u)1+CkAk,R(u) (5.36)

    for all u0. We have in particular

    Ak,R(u)=max{1,u}mu0Uk,R(v)dv,Bk,R(u)=(2k+3m)u2k+2m  for  u<R,Bk,R(u)=0  for u>R.

    We obviously have (5.36) for u1. Indeed, in this case u1<R, and we can compute

    Ak,R(u)=u0Uk,R(v)(uv)dv=u2k+3(2k+2)(2k+3),Bk,R(u)=u2k+3m,

    so that (5.36) holds with Ck=6(k+1)2. For u>1 it suffices to prove that

    Bk,R(u)CkAk,R(u). (5.37)

    Indeed, then (5.36) follows from the formula Bk,R(u)Bk,R(1)Ck(Ak,R(u)Ak,R(1)) and the fact that (5.36) holds for u=1 from the previous step. Note that (5.37) holds automatically for u>R. For u(1,R) we have by (5.8) that

    Ak,R(u)=umu0v2k+1dv=12k+2u2k+2m,

    hence (5.37) and (5.36) hold with Ck as above. Therefore, combining (5.34) and (5.36), in terms of the functions wi(x) defined in (5.26), estimate (5.31) implies

    maxi=1,,n|wi|2k+3m2k+3mCkN+3max{C2k,τni=1|wi|2kk} (5.38)

    for all k>(m3)/2 with a constant C>0 independent of k and R.

    We are ready to start the Moser iterations. Note that we can assume m3. Indeed, if (3.16) holds with m<3, it will certainly be true with m=3. Since we have an anisotropic norm on the right-hand side of (5.38), we have to replace the left-hand side by an expression which is compatible with the right-hand side. We shall see that, since

    (τni=1|wi|4k+62m2k+3m)1/2(τnmaxi=1,,n(|wi|2k+3m2k+3m)2)1/2,

    the right choice is

    max{C2k+3m,(τni=1|wi|4k+62m2k+3m)1/2}CTkN+3max{C2k,τni=1|wi|2kk} (5.39)

    with a constant CT depending on T. We now choose k0>m3 and define a sequence kj for jN recurrently by the formula kj=2kj1+3m, that is,

    kj=2jKm+m3,Km=k0+3m. (5.40)

    From (5.39) we obtain

    \begin{equation} \left(\max\left\{C, \left(\tau{\sum\limits_{i = 1}^n}\left|w_i\right|_{k_j}^{2k_j}\right)^{1/2k_j}\right\}\right)^{\omega_j} \le \left(C_T k_{j-1}^{N+3}\right)^{1/2k_{j-1}}\max\left\{C, \left(\tau{\sum\limits_{i = 1}^n}\left|w_i\right|_{k_{j-1}}^{2k_{j-1}}\right)^{1/2 k_{j-1}} \right\} \end{equation} (5.41)

    with exponent

    \omega_j = \frac{2k_{j-1}+3-m}{2k_{j-1}} = 1- \frac{m-3}{2k_{j-1}}.

    Note that (m-3)/2k_{j-1} \in [0, 1/2) , so that \omega_j \in (1/2, 1] . We define the quantities

    L_j { :=} \log\left(\max\left\{C, \left(\tau{\sum\limits_{i = 1}^n}\left|w_i\right|_{k_j}^{2k_j}\right)^{1/2k_j} \right\}\right).

    Then from (5.41) we get the recurrent relation

    \begin{equation} \omega_j L_{j} \le \frac1{2k_{j-1}} (\log C_T + (N+3) \log k_{j-1}) + L_{j-1} \end{equation} (5.42)

    for all j \in {\mathbb{N}} . We rewrite (5.42) in the form

    \begin{equation} \left(\prod\limits_{i = 1}^j \omega_i\right) L_{j} \le \left(\prod\limits_{i = 1}^{j-1}\omega_i\right) \frac1{2k_{j-1}} (\log C_T + (N+3) \log k_{j-1}) + \left(\prod\limits_{i = 1}^{j-1}\omega_i\right) L_{j-1}. \end{equation} (5.43)

    Summing up the above inequality over j = 2, \dots, k yields

    \begin{equation} L_k \le \left(\prod\limits_{i = 1}^{k} \frac{1}{\omega_i}\right) \sum\limits_{j = 2}^k \left(\prod\limits_{i = 1}^{j-1}\omega_i\right) \frac1{2k_{j-1}} (\log C_T + (N+3) \log k_{j-1}) + \left(\prod\limits_{i = 2}^{k} \frac{1}{\omega_i}\right) L_1. \end{equation} (5.44)

    To estimate L_1 , we observe that by virtue of (5.39) we have

    \begin{equation} \begin{aligned} \left(\tau{\sum\limits_{i = 1}^n}\left|w_i\right|_{k_1}^{2k_1}\right)^{1/2k_1} & = \left(\left(\tau{\sum\limits_{i = 1}^n}\left|w_i\right|_{2k_0+3-m}^{4k_0+6-2m}\right)^{1/2}\right)^{1/(2k_0+3-m)} \\ &\le C\left(1 + \tau{\sum\limits_{i = 1}^n}\left|w_i\right|_{k_0}^{2k_0}\right)^{1/(2k_0+3-m)}. \end{aligned} \end{equation} (5.45)

    Poincaré's inequality implies

    \begin{equation} |{V_{k, R}}(u_i)|_2^2 \le C \|{V_{k, R}}(u_i)\|_{1, 2}^2, \end{equation} (5.46)

    so that from (5.25) and (5.27) it follows that for all k\ge 1 we have

    \begin{equation} \begin{aligned} \tau{\sum\limits_{i = 1}^n} |w_i|_{2k+2}^{2k+2} &\le \tau{\sum\limits_{i = 1}^n}|{V_{k, R}}(u_i)|_2^2 \\ &\le C \tau{\sum\limits_{i = 1}^n}\|{V_{k, R}}(u_i)\|_{1, 2}^2 \\ &\le C(k+1)\Big(C^{2k} + (2k+1)\tau{\sum\limits_{i = 1}^n} |w_i|_{2k}^{2k}\Big). \end{aligned} \end{equation} (5.47)

    By induction over k we prove that for k_0 > m-3 as above the estimate

    \begin{equation} \tau{\sum\limits_{i = 1}^n} |w_i|_{2k_0}^{2k_0} \le C \end{equation} (5.48)

    holds true with a constant C depending only on m in Hypothesis 3.6 and on the data of the problem. Hence, by (5.45) and (5.48) we obtain that L_1 \le C . Coming back to (5.44), we note also that

    \frac{1}{\omega_i} = 1 + \frac{m-3}{2k_{i-1}+3-m},

    and since

    \prod\limits_{i = 1}^\infty \frac{1}{\omega_i} < \infty \ \Longleftrightarrow \ \sum\limits_{i = 1}^\infty \log\bigg(\frac{1}{\omega_i}\bigg) < \infty \ \Longleftrightarrow \ \sum\limits_{i = 1}^\infty \frac{m-3}{2k_{i-1}+3-m} < \infty,

    we conclude that \prod_{i = 1}^\infty 1/\omega_i < \infty . We additionally have

    \prod\limits_{i = 1}^\infty \omega_i > 0 \ \Longleftrightarrow \ \sum\limits_{i = 1}^\infty \frac{m-3}{2k_{i-1}} < \infty,

    see [23, Theorem 15.5], so that \prod_{i = 1}^\infty \omega_i > 0 . To summarize, we have

    \sum\limits_{j = 1}^\infty \frac1{2k_{j-1}} (\log C_T + (N+3) \log k_{j-1}) < \infty, \quad \prod\limits_{i = 1}^\infty\frac{1}{\omega_i} < \infty, \quad \prod\limits_{i = 1}^\infty\omega_i > 0,

    and we conclude from (5.44) that there exists a constant L^* > 0 such that L_k \le L^* for all k\in {\mathbb{N}} , hence

    \begin{equation} |u_i(x)| \le U { :=} {{{\rm{e}}}}^{L^*} \end{equation} (5.49)

    for a.e. x \in \Omega and all i\in \{1, \dots, n\} . By comparison in Eq (4.2), we also have

    \begin{equation} |v_i(x)| \le C \end{equation} (5.50)

    for a.e. x \in \Omega and all i\in \{1, \dots, n\} , with a constant C > 0 depending on U . Estimate (5.49) additionally implies |\xi_i^r(x)| \le (\hat{U}-r)^+ for \hat{U}{ :=}\max\{U, \Lambda\} , and we can write in (5.3)

    G[u]_i(x) = \bar G + \int_0^{\hat{U}} \psi(x, r, \xi^r_i(x)){\, \mathrm{d}} r.

    Hence, even if we do not assume the a priori boundedness of G as in (3.9), by assumption (3.7) we obtain

    \begin{equation} \quad |G[u]_i| \le C \end{equation} (5.51)

    with a constant C > 0 independent of i and \tau .

    Recall that the operator G is convexifiable in the sense of Definition 3.3, that is, for every U > 0 there exists a twice continuously differentiable mapping g:[-U, U] \to [-U, U] such that g(0) = 0 , 0 < g_* \le g'(u) \le g^* < \infty , |g''(U)| \le \bar{g} , and G is of the form

    \begin{equation} G = P \circ g, \end{equation} (6.1)

    where P is a uniformly counterclockwise convex Preisach operator on [-U, U] . Let us fix U from (5.49) and the corresponding function g . The following result is a straightforward consequence of [12, Proposition 3.6].

    Proposition 6.1. Let P be uniformly counterclockwise convex on [-U, U] , and let f be an odd increasing function such that f(0) = 0 . Then there exists \beta > 0 such that for every sequence \{w_i: i = -1, 0, \dots, n-1\} in [-U, U] we have

    \begin{equation} \begin{aligned} &{\sum\limits_{i = 0}^{n-1}} (P[w]_{i+1} - 2P[w]_i + P[w]_{i-1})f(w_{i+1} - w_i) + \frac{P[w]_0 - P[w]_{-1}}{w_0 - w_{-1}}F(w_0 - w_{-1}) \\ &\qquad \ge \frac{\beta}{2}{\sum\limits_{i = 0}^{n-1}} \Gamma(w_{i+1} - w_i), \end{aligned} \end{equation} (6.2)

    where we set for w\in {\mathbb{R}}

    \begin{equation} F(w) { :=} \int_0^w f(v){\, \mathrm{d}} v, \qquad \Gamma(w) { :=} |w|(wf(w) - F(w)) = |w|\int_0^{|w|}vf'(v){\, \mathrm{d}} v. \end{equation} (6.3)

    We need to define backward steps u_{-1} and v_{-1} satisfying the strong formulation of (4.1)-(4.2) for i = 0 , that is,

    \begin{align} \frac1\tau\Big((G[u]_0(x) - G[u]_{-1}(x)) +(v_0(x)- v_{-1}(x))\Big) & = {\mathrm{\, div\, }}\!\big(\kappa(x, \theta_0(x))\left(\nabla u_0(x) + {\boldsymbol{\nu}}\right)\big) \ \mbox{ in } \Omega, \end{align} (6.4)
    \begin{align} v_{-1}(x) & = (1+\tau)v_0(x) - \tau u_0(x) \ \mbox{ in } \Omega, \end{align} (6.5)

    with boundary condition (3.15). Repeating the argument of [12, Proposition 3.3], we use assumptions (3.7) and (3.14) to find for each 0 < \tau < \phi_0(U)/2L^2 functions u_{-1} and G[u]_{-1} satisfying

    \begin{equation} \frac1\tau(G[u]_0(x) - G[u]_{-1}(x)) = {\mathrm{\, div\, }}\!\big(\kappa(x, \theta_0(x))\left(\nabla u_0(x) + {\boldsymbol{\nu}}\right)\big) + v_0(x) - u_0(x) \ \mbox{ in } \Omega, \end{equation} (6.6)

    as well as, thanks to (3.13), the estimate

    \begin{equation} \frac1\tau |u_0(x) - u_{-1}(x)| \le C \end{equation} (6.7)

    with a constant C > 0 independent of \tau and x . Finally, we construct v_{-1} according to (6.5).

    We extend the discrete system (4.1)-(4.2) to i = 0 and write it in the form

    \begin{align} & {\int_{\Omega}} \left(\frac1\tau\Big((P[w]_i - P[w]_{i-1}) + (v_i - v_{i-1})\Big){y} + \kappa(x, \theta_{i})\big(\nabla u_i + {\boldsymbol{\nu}}\big)\cdot\nabla{y}\right){\, \mathrm{d}} x\\ & \quad \quad \quad \quad + {\int_{\partial\Omega}} b^*(x)(u_i - u^*_i){y} {\, \mathrm{d}} s(x) = 0, \end{align} (6.8)
    \begin{align} & \quad \frac1\tau (v_i - v_{i-1}) + v_i = u_i, \end{align} (6.9)

    with w_i = g(u_i) , \theta_i = G[u]_i , for i\in \{0, 1, \dots, n\} and for an arbitrary test function {y} \in W^{1, 2}(\Omega) . We proceed as in [12] and test the difference of (6.8) taken at discrete times i+1 and i

    \begin{equation} \begin{aligned} &{\int_{\Omega}} \frac1\tau\big(P[w]_{i{+}1} - 2P[w]_i + P[w]_{i{-}1} +v_{i{+}1} - 2v_i + v_{i{-}1}\big){y} {\, \mathrm{d}} x \\ &\qquad + {\int_{\Omega}} \Big(\kappa(x, \theta_{i{+}1})\nabla u_{i{+}1} - \kappa(x, \theta_i) \nabla u_i + (\kappa(x, \theta_{i{+}1}) - \kappa(x, \theta_{i})){\boldsymbol{\nu}}\Big) \cdot\nabla{y}{\, \mathrm{d}} x\\ &\qquad + {\int_{\partial\Omega}} b^*(x)(u_{i+1} -u_i){y} {\, \mathrm{d}} s(x) = {\int_{\partial\Omega}} b^*(x)(u^*_{i+1} - u^*_i){y} {\, \mathrm{d}} s(x) \end{aligned} \end{equation} (6.10)

    by {y} = f(w_{i+1} - w_i) with

    \begin{equation} f(w) { :=} \frac{w}{\tau + |w|}. \end{equation} (6.11)

    In agreement with (6.3), we have

    \begin{align} F(w) & = |w| - \tau\log\left(1 + \frac{|w|}{\tau}\right), \end{align} (6.12)
    \begin{align} \Gamma(w) & = \tau |w|\left(\log\left(1 + \frac{|w|}{\tau}\right) - \frac{|w|}{\tau + |w|}\right). \end{align} (6.13)

    The hysteresis term is estimated from below by virtue of Proposition 6.1 as follows:

    \begin{equation} \begin{aligned} \frac1\tau {\sum\limits_{i = 0}^{n-1}} &(P[w]_{i+1} - 2P[w]_i + P[w]_{i-1})f(w_{i+1} - w_i) + \frac{1}{\tau}\frac{P[w]_0-P[w]_{-1}}{w_0-w_{-1}}F(w_0-w_{-1}) \\ &\ge \frac{\beta}{2\tau}{\sum\limits_{i = 0}^{n-1}} \Gamma(w_{i+1} - w_i). \end{aligned} \end{equation} (6.14)

    Note that for |w| \ge \tau({{{\rm{e}}}}^2 - 1) , we have

    \frac{|w|}{\tau + |w|} < 1 \le \frac12\log\left(1 + \frac{|w|}{\tau}\right),

    so that

    \Gamma(w) \ge \frac{\tau}{2}|w|\log\left(1 + \frac{|w|}{\tau}\right).

    We denote by J the set of all i \in \{0, 1, \dots, n-1\} such that |w_{i+1} - w_i| \ge \tau({{{\rm{e}}}}^2 - 1) , and by J^\perp its complement. We thus have

    \begin{align} \frac12\sum\limits_{i\in J} |w_{i+1} - w_i|\log\left(1 + \frac{|w_{i+1} {-} w_i|}{\tau}\right) &\le \frac1\tau\sum\limits_{i\in J}\Gamma(w_{i+1} - w_i), \end{align} (6.15)
    \begin{align} \frac12\sum\limits_{i\in J^\perp} |w_{i+1} - w_i|\log\left(1 + \frac{|w_{i+1} {-} w_i|}{\tau}\right) &\le T({{{\rm{e}}}}^2 - 1), \end{align} (6.16)

    hence

    \frac{\beta}{2\tau}{\sum\limits_{i = 0}^{n-1}} \Gamma(w_{i+1} - w_i) \ge \frac\beta{4}{\sum\limits_{i = 0}^{n-1}} |w_{i+1} - w_i|\log\left(1 + \frac{|w_{i+1} {-} w_i|}{\tau}\right) - C

    with a constant C > 0 independent of x and \tau . Moreover, for w_0 \neq w_{-1} we have

    0 < \frac{F(w_0-w_{-1})}{|w_0-w_{-1}|} \le 1,

    which yields, together with identity (6.6) and assumption (3.13),

    \begin{equation} 0 < \frac{1}{\tau}\left|\frac{P[w]_0-P[w]_{-1}}{w_0-w_{-1}}\right| F(w_0-w_{-1}) \le C \end{equation} (6.17)

    with a constant C > 0 independent of x and \tau . For w_0 = w_{-1} , we interpret (P[w]_0-P[w]_{-1})/(w_0-w_{-1}) as B'_+(w_{-1}) or B'_-(w_{-1}) , according to the notation for the Preisach branches introduced after Proposition 3.2, see [12] for more details. From (6.14)–(6.17) we thus get

    \begin{equation} \frac1\tau {\sum\limits_{i = 0}^{n-1}} (P[w]_{i+1} - 2P[w]_i + P[w]_{i-1})f(w_{i+1} {-} w_i) \ge \frac{\beta}{4}{\sum\limits_{i = 0}^{n-1}} |w_{i+1} {-} w_i|\log\left(1 + \frac{|w_{i+1} {-} w_i|}{\tau}\right) - C \end{equation} (6.18)

    with a constant C > 0 independent of x and \tau .

    We further have by (6.9) and by the monotonicity of f and g that

    \begin{align} &\frac1\tau\big(v_{i{+}1} - 2v_i + v_{i{-}1}\big)f(w_{i+1} {-} w_i) = \Big((u_{i{+}1} - u_i) - (v_{i{+}1} - v_i)\Big)f(w_{i+1} {-} w_i)\\ &\qquad \ge \tau (v_{i+1} - u_{i+1})f(w_{i+1} {-} w_i) \ge -C\tau, \end{align} (6.19)

    where in the last step we employed estimates (5.49)-(5.50), and the constant C > 0 depends on U from (5.49).

    The boundary source term

    {\int_{\partial\Omega}} b^*(x)(u^*_{i+1} - u^*_i)f(w_{i+1} - w_i) {\, \mathrm{d}} s(x)

    is bounded by a constant by virtue of Hypothesis 3.4, while the boundary term on the left-hand side

    {\int_{\partial\Omega}} b^*(x)(u_{i+1} - u_i)f(w_{i+1} - w_i) {\, \mathrm{d}} s(x)

    is non-negative by monotonicity of both functions f and g . Using (6.18)-(6.19) we thus get from (6.10)

    \begin{align} &{\sum\limits_{i = 0}^{n-1}} {\int_{\Omega}} |w_{i+1} - w_i|\log\left(1 + \frac{|w_{i+1} {-} w_i|}{\tau}\right){\, \mathrm{d}} x\\ &\qquad + {\sum\limits_{i = 0}^{n-1}}{\int_{\Omega}} \Big(\kappa(x, \theta_{i{+}1})\nabla u_{i{+}1}{-}\kappa(x, \theta_i) \nabla u_i {+}(\kappa(x, \theta_{i{+}1}){-} \kappa(x, \theta_{i})){\boldsymbol{\nu}}\Big)\cdot \nabla f(w_{i+1} - w_i){\, \mathrm{d}} x \le C \end{align} (6.20)

    with a constant C > 0 independent of \tau . We further have

    \begin{align*} &\Big(\kappa(x, \theta_{i+1})\nabla u_{i+1}-\kappa(x, \theta_i) \nabla u_i\Big)\cdot \nabla f(w_{i+1} {-} w_i)\\ & = f'(w_{i+1} {-} w_i)\Bigg(\Big(\big(\kappa(x, \theta_{i+1}) {-} \kappa(x, \theta_{i})\big)\nabla u_i + \kappa(x, \theta_{i+1})\nabla (u_{i+1} {-} u_i)\Big)\\ &\qquad \times \Big(\big(g'(u_{i+1}) {-} g'(u_{i})\big)\nabla u_i + g'(u_{i+1})\nabla (u_{i+1} {-} u_i)\Big)\Bigg)\\ & = f'(w_{i+1} {-} w_i)\Bigg(g'(u_{i+1})\kappa(x, \theta_{i+1})|\nabla (u_{i+1} {-} u_i)|^2 + \big(g'(u_{i+1}) {-} g'(u_{i})\big)\big(\kappa(x, \theta_{i+1}) {-} \kappa(x, \theta_{i})\big)|\nabla u_i|^2\\ &\qquad + \Big(\kappa(x, \theta_{i+1})\big(g'(u_{i+1}) {-} g'(u_{i})\big) + g'(u_{i+1})\big(\kappa(x, \theta_{i+1}) {-} \kappa(x, \theta_{i})\big)\Big)\nabla u_i\cdot \nabla (u_{i+1} {-} u_i)\Bigg). \end{align*}

    The functions \kappa and g' are bounded and Lipschitz continuous, and

    \begin{equation} f'(w_{i+1} - w_i) = \frac{\tau}{(\tau + |w_{i+1} - w_i|)^2}. \end{equation} (6.21)

    Moreover, since \theta_i = P[w]_i admits a representation similar to (4.3), by Hypothesis 3.4, estimate (5.51), and the Lipschitz continuity of the time-discrete play implied by (4.5), we obtain

    \begin{equation} |\kappa(x, \theta_{i+1}) {-} \kappa(x, \theta_{i})| \le C|\theta_{i+1} {-} \theta_i| \le C|w_{i+1} {-} w_i|, \end{equation} (6.22)

    whereas, from assumption (3.10),

    \begin{equation} |g'(u_{i+1}) {-} g'(u_{i})| \le \bar{g}(U)|u_{i+1} {-} u_i| \le \bar{g}(U)(g_*(U))^{-1}|w_{i+1} {-} w_i|. \end{equation} (6.23)

    Thus, employing the Cauchy-Schwarz and Young's inequalities and assumption (3.10) on g , we conclude that there exist constants \delta > 0 and K > 0 independent of \tau such that

    \begin{equation} \begin{aligned} \Big(\kappa(x, \theta_{i+1})\nabla u_{i+1}&-\kappa(x, \theta_i) \nabla u_i\Big)\cdot \nabla f(w_{i+1} {-} w_i) \\ &\ge \delta f'(w_{i+1} {-} w_i)|\nabla (u_{i+1} {-} u_i)|^2 - K\frac{\tau}{(\tau + |u_{i+1} {-} u_i|)^2}|u_{i+1} {-} u_i|^2|\nabla u_i|^2 \\ &\ge \delta f'(w_{i+1} {-} w_i)|\nabla (u_{i+1} {-} u_i)|^2 -\tau K |\nabla u_i|^2. \end{aligned} \end{equation} (6.24)

    The remaining term (\kappa(x, \theta_{i{+}1}){-} \kappa(x, \theta_{i})){\boldsymbol{\nu}}\cdot \nabla f(w_{i+1} - w_i) in (6.20) can be treated as follows. By (3.10) we have

    \begin{aligned} |\nabla f(w_{i+1} - w_i)| & = f'(w_{i+1} - w_i)\big|g'(u_{i+1})\nabla u_{i+1} - g'(u_i)\nabla u_i\big| \\ & = f'(w_{i+1} - w_i)\big|(g'(u_{i+1})-g'(u_i))\nabla u_i + g'(u_{i+1})(\nabla u_{i+1}-\nabla u_i)\big| \\ &\le C f'(w_{i+1} - w_i)\Big(|u_{i+1}-u_{i}||\nabla u_i| + |\nabla(u_{i+1} - u_{i})|\Big), \end{aligned}

    and employing (6.22) and arguing as in (6.24) we get

    \begin{equation} |\kappa(x, \theta_{i{+}1}){-} \kappa(x, \theta_{i})| |\nabla f(w_{i+1} - w_i)| \le \frac\delta{2}f'(w_{i+1} {-} w_i)|\nabla (u_{i+1} {-} u_i)|^2 + C \tau\left(1{+} |\nabla u_i|^2\right) \end{equation} (6.25)

    with some constant C > 0 independent of \tau . As a consequence of (3.10), (6.20)–(6.25), and (5.7), we thus have the crucial estimate

    \begin{equation} {\sum\limits_{i = 0}^{n-1}} {\int_{\Omega}} |u_{i+1} - u_i|\log\left(1 + \frac{|u_{i+1} {-} u_i|}{\tau}\right){\, \mathrm{d}} x \le C\left(1+ \tau {\sum\limits_{i = 0}^n} {\int_{\Omega}}|\nabla u_i|^2 {\, \mathrm{d}} x\right) \le C \end{equation} (6.26)

    with a constant C > 0 independent of \tau .

    The estimates we have derived in Sections 5 and 6 are sufficient for passing to the limit as \tau\to 0 . For the time-discrete sequence \{u_i\} , we repeat literally the compactness argument in Sobolev and Orlicz spaces for piecewise linear and piecewise constant interpolates from [13, Sections 5 and 6]. The most delicate step is the convergence of the hysteresis terms. Estimates (5.7), (5.49), and (6.26) guarantee that the piecewise linear interpolates are bounded in the space

    \mathcal{B} { :=} L^\infty(\Omega\times (0, T))\cap {\mathcal X} \cap L^1(\Omega; W^{1, \Phi_{log}}(0, T)),

    where

    {\mathcal X} = \{u \in L^2(\Omega\times (0, T)) : \nabla u \in L^2(\Omega\times (0, T);{\mathbb{R}}^N)\}

    and

    W^{1, \Phi_{log}}(0, T) = \{u \in L^1(0, T): \dot u \in L^{\Phi_{log}}(0, T)\}.

    From [13, Proposition 6.1], the space \mathcal{B} is compactly embedded in L^1(\Omega; C[0, T]) , and this is enough to ensure the convergence of the hysteresis terms in view of Proposition 3.2. As for the piecewise linear and piecewise constant interpolates of the sequence \{v_i\} , their convergence follows from the (strong) convergence of the interpolates of \{u_i\} (compare in (4.2)) combined with the Lebesgue dominated convergence theorem. The limits are the desired solution to Problem (2.24)–(2.26), and this concludes the proof of Theorem 3.7.

    In this paper, we have investigated a model for unsaturated flow in a viscoelastic porous medium exhibiting degenerate hysteresis in the pressure-saturation relationship, coupled with gravity-driven moisture flux. This extends previous work on degenerate diffusion in porous media with hysteresis by incorporating the effects of solid matrix deformation and gravity, which introduce significant challenges in the analysis. The key difficulty lies in obtaining a priori bounds for the solutions, which we address by employing a hysteresis variant of the Moser iteration technique. Specifically, we require a condition on the Preisach density, namely, that it decays sufficiently slowly at infinity (Hypothesis 3.6). This condition allows us to establish a uniform upper bound for the solutions. Building upon this estimate, we leverage the convexity and compactness arguments developed in [13] to prove the existence of a weak solution to the coupled system. Our approach involves a time discretization scheme, the derivation of estimates independent of the time step, and a passage to the limit.

    The results presented here contribute to a deeper understanding of flow phenomena in deformable porous media, relevant to various applications in hydrogeology and related fields. Future work could explore the impact of more complex constitutive models for the solid matrix, as well as the development of efficient numerical methods for the computation of solutions.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The support from the European Union's Horizon Europe research and innovation programme under the Marie Skłodowska-Curie grant agreement No 101102708, from the MŠMT grant 8X23001, and from the GAČR project 24-10586S is gratefully acknowledged.

    The authors declare no conflicts of interest.



    [1] R. A. Adams, Sobolev spaces, Academic Press, 1975.
    [2] B. Albers, Modeling the hysteretic behavior of the capillary pressure in partially saturated porous media: a review, Acta Mech., 225 (2014), 2163–2189. https://doi.org/10.1007/s00707-014-1122-4 doi: 10.1007/s00707-014-1122-4
    [3] M. Al Saaideh, N. Alatawneh, M. Al Janaideh, A generalized Prandtl-Ishlinskii model for hysteresis modeling in electromagnetic devices, 2021 IEEE Energy Conversion Congress and Exposition (ECCE), 2021, 4513–4518. https://doi.org/10.1109/ECCE47101.2021.9595830 doi: 10.1109/ECCE47101.2021.9595830
    [4] F. Bagagiolo, A. Visintin, Hysteresis in filtration through porous media, Z. Anal. Anwend., 19 (2000), 977–997. https://doi.org/10.4171/ZAA/993 doi: 10.4171/ZAA/993
    [5] F. Bagagiolo, A. Visintin, Porous media filtration with hysteresis, Adv. Math. Sci. Appl., 14 (2004), 379–403.
    [6] M. Balato, S. Perna, C. Petrarca, C. Visone, A simple scheme for the inversion of a Preisach like hysteresis operator in saturation conditions, AIP Adv., 12 (2022), 035047. https://doi.org/10.1063/9.0000331 doi: 10.1063/9.0000331
    [7] O. V. Besov, V. P. Il'in, S. M. Nikol'skii, Integral representations of functions and embedding theorems, Nauka, Moscow, 1975.
    [8] E. El Behi-Gornostaeva, Well-posedness for flows in porous media with a hysteretic constitutive relation, Ph.D. Thesis, Technische Universität München, 2017. Available at: https://mediatum.ub.tum.de/doc/1335260/document.pdf.
    [9] E. El Behi-Gornostaeva, K. Mitra, B. Schweizer, Traveling wave solutions for the Richards equation with hysteresis, IMA J. Appl. Math., 84 (2019), 797–812. https://doi.org/10.1093/imamat/hxz015 doi: 10.1093/imamat/hxz015
    [10] M. Eleuteri, P. Krejčí, Asymptotic behaviour of a Neumann parabolic problem with hysteresis, Z. Angew. Math. Mech., 87 (2007), 261–277. https://doi.org/10.1002/zamm.200610299 doi: 10.1002/zamm.200610299
    [11] S. Fučík, A. Kufner, Nonlinear differential equations, Elsevier, 1980.
    [12] C. Gavioli, P. Krejčí, Degenerate diffusion with Preisach hysteresis, Discrete Contin. Dyn. Syst.-S, 16 (2023), 3677–3708. https://doi.org/10.3934/dcdss.2023154 doi: 10.3934/dcdss.2023154
    [13] C. Gavioli, P. Krejčí, Degenerate diffusion in porous media with hysteresis-dependent permeability, Discrete Contin. Dyn. Syst., 45 (2025), 1523–1542. https://doi.org/10.3934/dcds.2024137 doi: 10.3934/dcds.2024137
    [14] E. Heinz, An elementary analytic theory of the degree of mapping in n-dimensional space, Indiana Univ. Math. J., 8 (1959), 231–247. https://doi.org/10.1512/iumj.1959.8.58017 doi: 10.1512/iumj.1959.8.58017
    [15] M. Hilpert, On uniqueness for evolution problems with hysteresis, In: J. F. Rodrigues, Mathematical models for phase change problems, International Series of Numerical Mathematics, Birkhäuser, Basel, 88 (1989), 377–388. https://doi.org/10.1007/978-3-0348-9148-6_19
    [16] B. Hölting, W. G. Coldewey, Hydrogeology, Springer, 2019.
    [17] P. Krejčí, The Preisach hysteresis model: error bounds for numerical identification and inversion, Discrete Contin. Dyn. Syst.-S, 6 (2013), 101–119. https://doi.org/10.3934/dcdss.2013.6.101 doi: 10.3934/dcdss.2013.6.101
    [18] P. Krejčí, Hysteresis, convexity and dissipation in hyperbolic equations, Vol. 8, Tokyo: Gakkotosho, 1996.
    [19] K. Kuhnen, Modeling, identification and compensation of complex hysteretic nonlinearities: a modified Prandtl-Ishlinskii approach, Eur. J. Control, 9 (2003), 407–418. https://doi.org/10.3166/ejc.9.407-418 doi: 10.3166/ejc.9.407-418
    [20] H. Q. Pham, D. G. Fredlund, S. L. Barbour, A study of hysteresis models for soil-water characteristic curves, Can. Geotech. J., 42 (2005), 1548–1568. https://doi.org/10.1139/t05-071 doi: 10.1139/t05-071
    [21] F. Preisach, Über die magnetische Nachwirkung, Z. Phys., 94 (1935), 277–302. https://doi.org/10.1007/bf01349418 doi: 10.1007/bf01349418
    [22] M. M. Rao, Z. D. Ren, Applications of Orlicz spaces, 1 Ed., CRC Press, 2002. https://doi.org/10.1201/9780203910863
    [23] W. Rudin, Real and complex analysis, India: McGraw-Hill, 1966.
    [24] B. Schweizer, Hysteresis in porous media: modelling and analysis, Interfaces Free Bound., 19 (2017), 417–447. https://doi.org/10.4171/ifb/388 doi: 10.4171/ifb/388
    [25] A. Visintin, Differential models of hysteresis, Springer, Berlin Heidelberg, 1994. https://doi.org/10.1007/978-3-662-11557-2
    [26] A. Visintin, Mathematical models of hysteresis, In: G. Bertotti, I. Mayergoyz, The science of hysteresis, (2006), 1–123. https://doi.org/10.1016/b978-012480874-4/50004-x
  • This article has been cited by:

    1. Chiara Gavioli, Pavel Krejčí, Diffusion in porous media with hysteresis and bounded speed of propagation, 2025, 76, 0044-2275, 10.1007/s00033-025-02492-z
  • Reader Comments
  • © 2025 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(584) PDF downloads(118) Cited by(1)

Figures and Tables

Figures(1)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog