This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Homogenization of a multiscale model for water transport in vegetated soil

Andrew Maira\,{}^{\mathrm{a}}, Mariya Ptashnykb\,{}^{\mathrm{b}}
a Department of Conservation of Natural Resources, NEIKER, Derio, Basque Country, Spain
b Maxwell Institute for Mathematical Sciences, Department of Mathematics,
Heriot-Watt University, Edinburgh, Scotland, UK
Abstract

In this paper we consider the multiscale modelling of water transport in vegetated soil. In the microscopic model we distinguish between subdomains of soil and plant tissue, and use the Richards equation to model the water transport through each. Water uptake is incorporated by means of a boundary condition on the surface between root tissue and soil. Assuming a simplified root system architecture, which gives a cylindrical microstructure to the domain, the two-scale convergence and periodic unfolding methods are applied to rigorously derive a macroscopic model for water transport in vegetated soil. The degeneracy of the Richards equation and the dependence of root tissue permeability on the small parameter introduce considerable challenges in deriving macroscopic equations, especially in proving strong convergence. The variable-doubling method is used to prove the uniqueness of solutions to the model, and also to show strong two-scale convergence in the non-linear terms of the equation for water transport through root tissue.

Keywords: Richards equation, degenerate parabolic equations, variable-doubling, dual-porosity, homogenization, two-scale convergence and periodic unfolding

MSC codes: 35B27, 35K51, 35K65

1 Introduction

Mathematical models for water transport in soil are used to inform practice within different environmental contexts such as calculating crop yields [22], monitoring flood risk [15], or assessing hill slope stability [16]. Moreover, predictions from the models are becoming even more important when developing strategies to deal with the increasing impact of climate change. The commonly employed models for soil water transport account for root-soil interactions only by adding a root water uptake sink term to Richards equation, whereas existing dual-porosity models [20, 42] do not relate the architecture of the macropores to the distribution and properties of roots within the soil.

Thus, the main aim of the paper is to derive a microscopic model that explicitly distinguishes between water flow through subdomains of soil and periodically arranged root branches and apply homogenization techniques to derive a macroscopic model for water flow through vegetated soil. For more general root systems, this simple geometry can be considered locally and, with a suitable transformation, a macroscopic model can be derived. In the microscopic model, the Richards equation, with different permeability and water retention functions, is used to model water flow through the soil and root subdomains and water uptake is described by the boundary conditions at the root surface. Pressure gradients within the root tissue are driven by transpiration, incorporated through a boundary condition at the root crown. The difference between the properties of the bulk soil and the rhizosphere is reflected through the explicit spatial dependence in the water retention and hydraulic conductivity functions.

The Richards equation, with its non-linearities in the time derivative and elliptic operator, belongs to a larger class of degenerate parabolic equations. Well-posedness results for these types of equations and systems have been obtained by many authors, see e.g. [4, 8, 14, 23, 27]. Existence of strong solutions to the Richards equation was addressed in [33], the hysteresis problem for the Richards equation was analysed in [40], and viscosity solutions for a general class of non-linear degenerate parabolic equations were studied in [28]. In [38, 29] the uniqueness of solutions for degenerate parabolic equations and systems was shown using the variable-doubling method [30], and proving the L1L^{1}-contraction principle.

There are several results on multiscale analysis for the Richards equation and degenerate parabolic equations in general. The homogenization of non-linear parabolic equations with Dirichlet bounday conditions in a domain periodically perforated by small holes was studied in [35] and in [5, 7] the two-scale convergence method was applied to non-linear equations with oscilating coefficients modelling the two-phase flow in a porous medium. In [24] the homogenization of the Richards equation in a medium with deterministic almost periodic microstructure of coarse and fine impermeable soft inclusions is obtained using sigma-convergence method and an upscaling of the Richards equation to describe the flow in fractured porous media was considered in [31]. Parabolic and elliptic equations defined in domains with cylindrical microstructures, similar to the one considered here, were analysed in [13, 19]. Furthermore, in [34] homogenization techniques have been applied to dual-porosity models with the flow through porous particles and the inter-particle space described by the Richards equation. Similar to the model considered here, the permeability of the particles was dependent on the small parameter and, to show convergence of the non-linear terms, a specific form of test function was used in conjunction with the unfolding operator, also named as dilation operator and periodic modulation method [6, 10].

We consider a cylindrical microstructure representing the root branches where permeability of the root tissue is proportional to root branch radius and, hence, depends on the small parameter ε\varepsilon. The ε\varepsilon-dependence of the permeability tensor means that for the proof of the strong convergence, required to pass to the limit in the non-linear terms, a standard compactness argument cannot be used. Using the time-doubling method it is possible to show the equicontinuity of the unfolded sequence of solutions to the microscopic problem and prove the strong two-scale convergence. This is a novel application of the variable-doubling method in homogenization theory, which will allow the derivation of macroscopic equations for other degenerate parabolic problems. We would also like to remark that the ideas used to show the convergence results in [34] cannot be applied to our problem due to different scaling in the permeability function and different boundary conditions on the surfaces of microstructure. The paper is organised as follows. In section 2 we formulate the microscopic model for the water flow in vegetated soil. Section 3 is devoted to the analysis of the microscopic problem. The homogenization results and the derivation of the macroscopic model are then presented in section 4. Typical examples for the nonlinear hydraulic conductivity and water retention functions are given in Appendix.

2 Formulation of microscopic model

For the microscopic description of water flow in a vegetated soil we consider a simplified root system architecture where the root branches are periodically distributed throughout the domain. Consider Ω=(0,L1)×(0,L2)×(L3,0)\Omega=(0,L_{1})\times(0,L_{2})\times(-L_{3},0) representing a section of vegetated soil, with Γ0=Ω{x3=0}\Gamma_{0}=\partial\Omega\cap\{x_{3}=0\} and ΓL3=Ω{x3=L3}\Gamma_{L_{3}}=\partial\Omega\cap\{x_{3}=-L_{3}\} . The regions occupied by root branches PεP^{\varepsilon}, the rhizosphere soil around each root branch RεR^{\varepsilon}, the bulk soil BεB^{\varepsilon}, and the combined total soil subdomain SεS^{\varepsilon} are defined as Bε=Ω(P¯εR¯ε)B^{\varepsilon}=\Omega\setminus\big{(}\overline{P}^{\varepsilon}\cup\overline{R}^{\varepsilon}\big{)}, Sε=Int(R¯εB¯ε)S^{\varepsilon}={\rm Int}\big{(}\overline{R}^{\varepsilon}\cup\overline{B}^{\varepsilon}\big{)},

Pε=ξΞεε(YP+ξ)×(L3,0),Rε=ξΞεε(YRY¯P+ξ)×(L3,0),P^{\varepsilon}=\bigcup_{\xi\in\Xi_{\varepsilon}}\varepsilon(Y_{P}+\xi)\times(-L_{3},0),\quad R^{\varepsilon}=\bigcup_{\xi\in\Xi_{\varepsilon}}\varepsilon(Y_{R}\setminus\overline{Y}_{P}+\xi)\times(-L_{3},0),

see Figure 1, where Ξε={ξ2:ε(Y¯+ξ)(0,L1)×(0,L2)}\Xi_{\varepsilon}=\{\xi\in\mathbb{Z}^{2}\,:\,\varepsilon(\overline{Y}+\xi)\subset(0,L_{1})\times(0,L_{2})\}, the unit cell Y=(0,1)2Y=(0,1)^{2} and Y¯PY¯RY\overline{Y}_{P}\subset\overline{Y}_{R}\subset Y with boundaries ΓP\Gamma_{P} and ΓR\Gamma_{R} respectively, R^=YRY¯P\hat{R}=Y_{R}\setminus\overline{Y}_{P}, and B^=YY¯R\hat{B}=Y\setminus\overline{Y}_{R}. The lateral surfaces of the root branch microstructure and the boundaries between the rhizosphere and bulk soil are given by

ΓPε=ξΞεε(ΓP+ξ)×(L3,0) and ΓRε=ξΞεε(ΓR+ξ)×(L3,0),\Gamma_{P}^{\varepsilon}=\bigcup_{\xi\in\Xi_{\varepsilon}}\varepsilon(\Gamma_{P}+\xi)\times(-L_{3},0)\quad\text{ and }\quad\Gamma_{R}^{\varepsilon}=\bigcup_{\xi\in\Xi_{\varepsilon}}\varepsilon(\Gamma_{R}+\xi)\times(-L_{3},0),

and, with J=B,P,R,SJ=B,~{}P,~{}R,~{}S, the vertical boundaries are defined as

ΓJ,0ε={xΩJε:x3=0},ΓJ,L3ε={xΩJε:x3=L3}.\Gamma_{J,0}^{\varepsilon}=\{x\in\partial\Omega\cap\partial J^{\varepsilon}:x_{3}=0\},\qquad\Gamma_{J,L_{3}}^{\varepsilon}=\{x\in\partial\Omega\cap\partial J^{\varepsilon}:x_{3}=-L_{3}\}.
x2x_{2}x1x_{1}ε\varepsilonPεP^{\varepsilon}RεR^{\varepsilon}BεB^{\varepsilon}L1L_{1}
(a)
x3x_{3}x2x_{2}L2L_{2}x3=0x_{3}=0x3=L3x_{3}=-L_{3}PεP^{\varepsilon}PεP^{\varepsilon}ε\varepsilonRεR^{\varepsilon}BεB^{\varepsilon}
(b)
Figure 1: An illustration of the domain Ω\Omega comprised of bulk soil BεB^{\varepsilon}, rhizosphere soil RεR^{\varepsilon}. and root tissue PεP^{\varepsilon}; (a) from above; (b) cross-section of the x2x3x_{2}-x_{3} plane at x1=L1/2x_{1}={L_{1}}/{2}.

The water transport in unsaturated soil is modelled by the Richards equation

tθSε(hSε)(KSε(hSε)(hSε+e3))\displaystyle\partial_{t}\theta_{S}^{\varepsilon}(h_{S}^{\varepsilon})-\nabla\cdot(K_{S}^{\varepsilon}(h_{S}^{\varepsilon})(\nabla h_{S}^{\varepsilon}+e_{3})) =0\displaystyle=0 in Sε×(0,T],\displaystyle\text{ in }\;S^{\varepsilon}\times(0,T], (2.1)

where e3=(0,0,1)e_{3}=(0,0,1)^{\top}. In root tissue, water is transported through the xylem, which can be regarded as a porous medium [21], and hence is also modelled by the Richards equation

tθP(hPε)(IεKP(hPε)(hPε+e3))=0 in Pε×(0,T].\partial_{t}\theta_{P}(h_{P}^{\varepsilon})-\nabla\cdot\big{(}I_{\varepsilon}K_{P}(h_{P}^{\varepsilon})(\nabla h_{P}^{\varepsilon}+e_{3})\big{)}=0\quad\text{ in }\;\;P^{\varepsilon}\times(0,T]. (2.2)

In equations (2.1) and (2.2), KSεK_{S}^{\varepsilon} and KPK_{P} denote the hydraulic conductivity functions of soil and root tissue, and θSε\theta_{S}^{\varepsilon} and θP\theta_{P} are the soil and root water retention functions, respectively. The hydraulic properties of root tissue are anisotropic [18] and are defined in terms of root segment radius εr\varepsilon r and the intrinsic axial conductance of the root tissue kaxk_{\text{ax}} as

IεKP(hPε)=(2πεrρgkr0002πεrρgkr000kaxρgL3)κP(hPε), where Iε=(ε000ε0001),\displaystyle I_{\varepsilon}K_{P}(h_{P}^{\varepsilon})=\begin{pmatrix}2\pi\varepsilon r\rho gk_{\text{r}}&0&0\\ 0&2\pi\varepsilon r\rho gk_{\text{r}}&0\\ 0&0&\frac{k_{\text{ax}}\rho g}{L_{3}}\end{pmatrix}\kappa_{P}(h_{P}^{\varepsilon}),\;\text{ where }\;I_{\varepsilon}=\begin{pmatrix}\varepsilon&0&0\\ 0&\varepsilon&0\\ 0&0&1\\ \end{pmatrix},

see e.g. [45, 44], and the function κP\kappa_{P} incorporates the dependence of hydraulic conductivity on xylem pressure head hPεh_{P}^{\varepsilon}. To address the soil heterogeneity, we consider different hydraulic conductivity and retention functions for rhizosphere θR,KR\theta_{R},K_{R} and bulk soil θB,KB\theta_{B},K_{B}:

θSε(x,hSε)=θS(x^ε,hSε) and KSε(x,hSε)=KS(x^ε,hSε)\theta_{S}^{\varepsilon}(x,h_{S}^{\varepsilon})=\theta_{S}\big{(}\frac{\hat{x}}{\varepsilon},h_{S}^{\varepsilon}\big{)}\quad\text{ and }\quad K_{S}^{\varepsilon}(x,h_{S}^{\varepsilon})=K_{S}\big{(}\frac{\hat{x}}{\varepsilon},h_{S}^{\varepsilon}\big{)}

Here x^=(x1,x2)\hat{x}=(x_{1},x_{2}) and θS,KS\theta_{S},K_{S} are YY-periodic in 3\mathbb{R}^{3} such that for yYy\in Y and hh\in\mathbb{R}

θS(y,h)=θR(h)χR^(y)+θB(h)χB^(y),KS(y,h)=KR(h)χR^(y)+KB(h)χB^(y),\theta_{S}(y,h)=\theta_{R}(h)\chi_{\hat{R}}(y)+\theta_{B}(h)\chi_{\hat{B}}(y),\;\qquad K_{S}(y,h)=K_{R}(h)\chi_{\hat{R}}(y)+K_{B}(h)\chi_{\hat{B}}(y),

with χJ\chi_{J} denoting the characteristic function of the subdomain J=B,RJ=B,R. Hence,

hSε(t,x)=hRε(t,x)χRε(x)+hBε(t,x)χBε(x),h_{S}^{\varepsilon}(t,x)=h_{R}^{\varepsilon}(t,x)\chi_{R^{\varepsilon}}(x)+h_{B}^{\varepsilon}(t,x)\chi_{B^{\varepsilon}}(x),

where hRεh_{R}^{\varepsilon} and hBεh_{B}^{\varepsilon} denote the pressure head in the rhizosphere and bulk soil, respectively. We have the continuity of the pressure head and flux on the interface between rhizosphere and bulk soil

hRε=hBε,KRε(hRε)(hRε+e3)ν=KBε(hBε)(hBε+e3)ν on ΓRε×(0,T],h_{R}^{\varepsilon}=h_{B}^{\varepsilon},\qquad K_{R}^{\varepsilon}(h_{R}^{\varepsilon})(\nabla h_{R}^{\varepsilon}+{e}_{3})\cdot\nu=K_{B}^{\varepsilon}(h_{B}^{\varepsilon})(\nabla h_{B}^{\varepsilon}+e_{3})\cdot\nu\quad\text{ on }\;\;\Gamma_{R}^{\varepsilon}\times(0,T],

and across the rhizosphere-root boundary ΓPε\Gamma_{P}^{\varepsilon}, we set the flux of water as proportional to the difference between water pressure head in the soil hSεh_{S}^{\varepsilon} and root tissue hPεh_{P}^{\varepsilon} [45, 44].

At the soil-atmosphere interface ΓS,0ε\Gamma_{S,0}^{\varepsilon}, precipitation, evaporation and surface run-off are all accounted for by a function ff, and no-flux conditions are imposed on the lateral components of the external soil boundary ΓSε\Gamma_{S}^{\varepsilon}. We also assume that the lower soil boundary ΓS,L3ε\Gamma_{S,L_{3}}^{\varepsilon} is at the interface between the soil and the water table. At the root collars ΓP,0ε\Gamma_{P,0}^{\varepsilon} the outward flux is set as equal to the potential transpiration demand 𝒯pot\mathcal{T}_{\text{pot}}\geq 0, often taken to be the product of a species-dependent crop coefficient Kcb\text{K}_{\text{cb}} and a climate-dependent reference evapotranspiration constant ETo\text{ET}_{o} [3], and we prescribe a constant water pressure head a0a\leq 0 on the lower boundary at the root cap ΓP,L3ε\Gamma_{P,L_{3}}^{\varepsilon}. The full boundary conditions are therefore stated as

KSε(hSε)(hSε+e3)ν\displaystyle-K_{S}^{\varepsilon}(h_{S}^{\varepsilon})(\nabla h_{S}^{\varepsilon}+e_{3})\cdot\nu =εkΓ(hSεhPε)\displaystyle=\varepsilon k_{\Gamma}(h_{S}^{\varepsilon}-h_{P}^{\varepsilon}) on ΓPε×(0,T],\displaystyle\text{ on }\Gamma_{P}^{\varepsilon}\times(0,T], (2.3)
KSε(hSε)(hSε+e3)ν\displaystyle-K_{S}^{\varepsilon}(h_{S}^{\varepsilon})(\nabla h_{S}^{\varepsilon}+e_{3})\cdot\nu =f(hSε)\displaystyle=f(h_{S}^{\varepsilon}) on ΓS,0ε×(0,T],\displaystyle\text{ on }\Gamma_{S,0}^{\varepsilon}\times(0,T],
KSε(hSε)(hSε+e3)ν\displaystyle-K_{S}^{\varepsilon}(h_{S}^{\varepsilon})(\nabla h_{S}^{\varepsilon}+e_{3})\cdot\nu =0\displaystyle=0 on ΓSε×(0,T],\displaystyle\text{ on }\Gamma_{S}^{\varepsilon}\times(0,T],
hSε\displaystyle h_{S}^{\varepsilon} =0\displaystyle=0 on ΓS,L3ε×(0,T],\displaystyle\text{ on }\Gamma_{S,L_{3}}^{\varepsilon}\times(0,T],
hSε(0)\displaystyle h_{S}^{\varepsilon}(0) =hS,0\displaystyle=h_{S,0} in Sε,\displaystyle\text{ in }S^{\varepsilon},

and

IεKP(hPε)(hPε+e3)ν\displaystyle-I_{\varepsilon}K_{P}(h_{P}^{\varepsilon})(\nabla h_{P}^{\varepsilon}+{e}_{3})\cdot\nu =εkΓ(hPεhSε)\displaystyle=\varepsilon k_{\Gamma}(h_{P}^{\varepsilon}-h_{S}^{\varepsilon}) on ΓPε×(0,T],\displaystyle\text{ on }\Gamma_{P}^{\varepsilon}\times(0,T], (2.4)
IεKP(hPε)(hPε+e3)ν\displaystyle-I_{\varepsilon}K_{P}(h_{P}^{\varepsilon})(\nabla h_{P}^{\varepsilon}+{e}_{3})\cdot\nu =𝒯pot\displaystyle=\mathcal{T}_{\text{pot}} on ΓP,0ε×(0,T],\displaystyle\text{ on }\Gamma_{P,0}^{\varepsilon}\times(0,T],
hPε\displaystyle h_{P}^{\varepsilon} =a\displaystyle=a on ΓP,L3ε×(0,T],\displaystyle\text{ on }\Gamma_{P,L_{3}}^{\varepsilon}\times(0,T],
hPε(0)\displaystyle h_{P}^{\varepsilon}(0) =hP,0\displaystyle=h_{P,0} in Pε,\displaystyle\text{ in }P^{\varepsilon},

where the term kΓ=krρgk_{\Gamma}=k_{\rm r}\rho g is the product of the intrinsic radial conductance of the root tissue krk_{r}, the water density ρ\rho, and gravitational acceleration gg.

3 Existence and uniqueness results

In this section we prove the well-posedness of model (2.1), (2.3) and (2.2), (2.4). There are many well-posedness results for degenerate parabolic equations, but due to specific assumptions on the non-linear functions and boundary conditions considered here, we present the main steps of the proofs, using the methods as in [4, 23]. We consider the space

V(Jε)={hH1(Jε):h=0 on ΓJ,L3ε}, for J=S,P.V(J^{\varepsilon})=\big{\{}h\in H^{1}(J^{\varepsilon})\;:\;h=0\;\text{ on }\;\Gamma^{\varepsilon}_{J,L_{3}}\big{\}},\qquad\text{ for }\;J=S,P.

For uV(Jε)u\in V(J^{\varepsilon}) and ψV(Jε)\psi\in V(J^{\varepsilon})^{\prime} the dual product is denoted by u,ψV(Jε)\langle u,\psi\rangle_{V(J^{\varepsilon})^{\prime}} and the inner product for ϕLp(A)\phi\in L^{p}(A) and ψLq(A)\psi\in L^{q}(A) by ϕ,ψA=Aϕψ𝑑x\langle\phi,\psi\rangle_{A}=\int_{A}\phi\psi dx, given a domain A3A\subset\mathbb{R}^{3} and 1/p+1/q=11/p+1/q=1. Similar notation is used for the inner product for ϕLp(A)\phi\in L^{p}(\partial A) and ψLq(A)\psi\in L^{q}(\partial A). We also denote u,ψV(Jε),τ=0τu(t),ψ(t)V(Jε)𝑑t\langle u,\psi\rangle_{V(J^{\varepsilon})^{\prime},\tau}=\int_{0}^{\tau}\langle u(t),\psi(t)\rangle_{V(J^{\varepsilon})^{\prime}}dt and Aτ=(0,τ)×AA_{\tau}=(0,\tau)\times A, where AA is a domain or boundary of a domain.

Assumption 3.1.
  • (A1)

    The functions θR,θB,θP:{\theta}_{R},\theta_{B},{\theta}_{P}:\mathbb{R}\to\mathbb{R} are strictly increasing and Lipschitz continuous, θJ(0)=0{\theta}_{J}(0)=0, and <θJ,minθJ(z)θJ,max<+-\infty<{\theta}_{J,\text{min}}\leq{\theta}_{J}(z)\leq{\theta}_{J,\text{max}}<+\infty for all zz\in\mathbb{R}, where J=R,B,PJ=R,B,P.

  • (A2)

    The functions KR,KB,KP:[0,)K_{R},K_{B},K_{P}:\mathbb{R}\to[0,\infty) are continuous and 0<KJ,0KJ(z)KJ,sat0<K_{J,0}\leq K_{J}(z)\leq K_{J,\text{sat}} for all zz\in\mathbb{R}, for J=R,B,PJ=R,B,P.

  • (A3)

    The function f:f:\mathbb{R}\to\mathbb{R} is continuous and |f(z)|fm\lvert f(z)\rvert\leq f_{m} for all zz\in\mathbb{R}.

  • (A4)

    The initial conditions hS,0V(Sε)h_{S,0}\in V(S^{\varepsilon}), hP,0aV(Pε)h_{P,0}-a\in V(P^{\varepsilon}) are non-positive.

Remark 3.2.

The assumption θJ(0)=0\theta_{J}(0)=0, with J=R,B,PJ=R,B,P, is not restrictive, since we can define shifted functions θJθJ,sat\theta_{J}-\theta_{J,\text{sat}} that will satisfy the same problem, where θJ,sat\theta_{J,\text{sat}} denotes the saturated water content for the soil and root xylem respectively.

Definition 3.3.

A weak solution of model (2.1)–(2.4) is (hSε,hPε)(h_{S}^{\varepsilon},h_{P}^{\varepsilon}) with hSεL2(0,T;V(Sε))h_{S}^{\varepsilon}\in L^{2}(0,T;V(S^{\varepsilon})), hPεaL2(0,T;V(Pε))h_{P}^{\varepsilon}-a\in L^{2}(0,T;V(P^{\varepsilon})), tθSε(,hSε)L2(0,T;V(Sε))\partial_{t}{\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon})\in L^{2}\big{(}0,T;V(S^{\varepsilon})^{\prime}\big{)}, tθP(hPε)L2(0,T;V(Pε))\partial_{t}{\theta}_{P}(h_{P}^{\varepsilon})\in L^{2}\big{(}0,T;V(P^{\varepsilon})^{\prime}\big{)}, and satisfying

0T[tθSε(x,hSε),ϕV(Sε)+KSε(x,hSε)(hSε+e3),ϕSε\displaystyle\int_{0}^{T}\Big{[}\big{\langle}\partial_{t}{\theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon}),\phi\big{\rangle}_{V(S^{\varepsilon})^{\prime}}+\big{\langle}K_{S}^{\varepsilon}(x,h_{S}^{\varepsilon})(\nabla h_{S}^{\varepsilon}+{e}_{3}),\nabla\phi\big{\rangle}_{S^{\varepsilon}} (3.1)
+εkΓhSεhPε,ϕΓPε+f(hSε),ϕΓS,0ε]dt\displaystyle+\varepsilon k_{\Gamma}\big{\langle}h_{S}^{\varepsilon}-h_{P}^{\varepsilon},\phi\big{\rangle}_{\Gamma_{P}^{\varepsilon}}+\big{\langle}f(h_{S}^{\varepsilon}),\phi\big{\rangle}_{\Gamma_{S,0}^{\varepsilon}}\Big{]}dt =0,\displaystyle=0,
0T[tθP(hPε),ψV(Pε)+IεKP(hPε)(hPε+e3),ψPε\displaystyle\int_{0}^{T}\Big{[}\big{\langle}\partial_{t}{\theta}_{P}(h_{P}^{\varepsilon}),\psi\big{\rangle}_{V(P^{\varepsilon})^{\prime}}+\big{\langle}I_{\varepsilon}K_{P}(h_{P}^{\varepsilon})(\nabla h_{P}^{\varepsilon}+{e}_{3}),\nabla\psi\big{\rangle}_{P^{\varepsilon}} (3.2)
+εkΓhPεhSε,ψΓPε+𝒯pot,ψΓP.0ε]dt\displaystyle+\varepsilon k_{\Gamma}\big{\langle}h_{P}^{\varepsilon}-h_{S}^{\varepsilon},\psi\big{\rangle}_{\Gamma_{P}^{\varepsilon}}+\big{\langle}\mathcal{T}_{\text{pot}},\psi\big{\rangle}_{\Gamma^{\varepsilon}_{P.0}}\Big{]}dt =0,\displaystyle=0,

for all ϕL2(0,T;V(Sε))\phi\in L^{2}(0,T;V(S^{\varepsilon})) and ψL2(0,T;V(Pε))\psi\in L^{2}(0,T;V(P^{\varepsilon})).

In the proofs we will use the functions ΘP{\Theta}_{P}, ΘR\Theta_{R}, ΘB{\Theta}_{B}, and ΘSε\Theta_{S}^{\varepsilon}, defined as

ΘJ(hJ)\displaystyle{\Theta}_{J}(h_{J}) =θJ(hJ)hJ+hJ0θJ(z)𝑑z0, for J=P,R,B,\displaystyle={\theta}_{J}(h_{J})h_{J}+\int_{h_{J}}^{0}{\theta}_{J}(z)dz\geq 0,\quad\text{ for }\;J=P,R,B, (3.3)
ΘSε(x,hS)\displaystyle{\Theta}^{\varepsilon}_{S}(x,h_{S}) =ΘR(hS)χRε(x)+ΘB(hS)χBε(x).\displaystyle={\Theta}_{R}(h_{S})\chi_{R^{\varepsilon}}(x)+{\Theta}_{B}(h_{S})\chi_{B^{\varepsilon}}(x).

The definition of ΘJ\Theta_{J} and monotonicity assumptions on θJ\theta_{J}, for J=P,R,BJ=P,R,B, imply

ΘJ(u)ΘJ(v)=θJ(u)u+u0θJ(z)𝑑z[θJ(v)v+v0θJ(z)𝑑z](θJ(u)θJ(v))u.\displaystyle{\Theta}_{J}(u)-{\Theta}_{J}(v)={\theta}_{J}(u)u+\int_{u}^{0}{\theta}_{J}(z)dz-\Big{[}{\theta}_{J}(v)v+\int_{v}^{0}{\theta}_{J}(z)dz\Big{]}\leq\big{(}{\theta}_{J}(u)-{\theta}_{J}(v)\big{)}u. (3.4)
Remark 3.4.

For simplicity of presentation, in the proofs of existence, uniqueness, and non-positivity results, obtained for each fixed ε>0\varepsilon>0, we shall use the notation θS(u):=θSε(x,u)\theta_{S}(u):=\theta_{S}^{\varepsilon}(x,u), ΘS(u):=ΘSε(x,u)\Theta_{S}(u):=\Theta_{S}^{\varepsilon}(x,u), and KS(u):=KSε(x,u)K_{S}(u):=K^{\varepsilon}_{S}(x,u).

Theorem 3.5.

Under Assumption 3.1 for each fixed ε>0\varepsilon>0 there exists a weak solution (hSε,hPε)(h_{S}^{\varepsilon},h_{P}^{\varepsilon}) to model (2.1), (2.3), (2.2), and (2.4).

Proof.

The existence of a solution can be shown using the Rothe-Galerkin method, see e.g. [26, 17]. Since ε>0\varepsilon>0 is fixed, for the simplicity of notation we omit the superscript ε\varepsilon in hSεh^{\varepsilon}_{S} and hPεh^{\varepsilon}_{P}. With i=1,,Ni=1,\ldots,N, where N=T/τN=T/\tau\in\mathbb{N} for some τ>0\tau>0, and mm\in\mathbb{N}, we consider

hS,im=k=1mαi,kmvk and hP,ima=k=1mβi,kmwk,h_{S,i}^{m}=\sum_{k=1}^{m}\alpha_{i,k}^{m}v_{k}\quad\text{ and }\quad h_{P,i}^{m}-a=\sum_{k=1}^{m}\beta_{i,k}^{m}w_{k},

where {vk}k=1\{v_{k}\}_{k=1}^{\infty} and {wk}k=1\{w_{k}\}_{k=1}^{\infty} are orthogonal bases for V(Sε)V(S^{\varepsilon}) and V(Pε)V(P^{\varepsilon}), respectively, and are orthonormal in L2(Sε)L^{2}(S^{\varepsilon}) and L2(Pε)L^{2}(P^{\varepsilon}), and the discrete-in-time equations

1τθS(hS,im)θS(hS,i1m),ϕSε+KS(hS,i1m)(hS,im+e3),ϕSε\displaystyle\frac{1}{\tau}\big{\langle}\theta_{S}(h_{S,i}^{m})-\theta_{S}(h_{S,i-1}^{m}),\phi\big{\rangle}_{S^{\varepsilon}}+\big{\langle}K_{S}(h_{S,i-1}^{m})(\nabla h_{S,i}^{m}+{e}_{3}),\nabla\phi\big{\rangle}_{S^{\varepsilon}} (3.5)
+εkΓhS,imhP,i1m,ϕΓPε+f(hS,i1m),ϕΓS,0ε=0,\displaystyle+\varepsilon\,k_{\Gamma}\big{\langle}h_{S,i}^{m}-h_{P,i-1}^{m},\phi\big{\rangle}_{\Gamma_{P}^{\varepsilon}}+\big{\langle}f(h_{S,i-1}^{m}),\phi\big{\rangle}_{\Gamma_{S,0}^{\varepsilon}}=0,
1τθP(hP,im)θP(hP,i1m),ψPε+IεKP(hP,i1m)(hP,im+e3),ψPε\displaystyle\frac{1}{\tau}\big{\langle}\theta_{P}(h_{P,i}^{m})-{\theta}_{P}(h_{P,i-1}^{m}),\psi\big{\rangle}_{P^{\varepsilon}}+\big{\langle}I_{\varepsilon}K_{P}(h_{P,i-1}^{m})(\nabla h_{P,i}^{m}+{e}_{3}),\nabla\psi\big{\rangle}_{P^{\varepsilon}} (3.6)
εkΓhS,i1mhP,im,ψΓPε+𝒯pot,ψΓP,0ε=0,\displaystyle-\varepsilon\,k_{\Gamma}\langle h_{S,i-1}^{m}-h_{P,i}^{m},\psi\rangle_{\Gamma_{P}^{\varepsilon}}+\langle\mathcal{T}_{\text{pot}},\psi\rangle_{\Gamma_{P,0}^{\varepsilon}}=0,

for all ϕVm\phi\in V_{m}ψUm\psi\in U_{m}, with Vm=span{v1,,vm}V_{m}={\rm span}\{v_{1},...,v_{m}\} and Um=span{w1,,wm}U_{m}={\rm span}\{w_{1},...,w_{m}\} and hS,0mh_{S,0}^{m} and hP,0mah_{P,0}^{m}-a being the projections of hS,0h_{S,0} and hP,0ah_{P,0}-a onto VmV_{m} and UmU_{m} respectively. The regularity of hS,0h_{S,0} and hP,0h_{P,0} implies hS,0mhS,0h_{S,0}^{m}\to h_{S,0} in V(Sε)V(S^{\varepsilon}) and hP,0mahP,0ah_{P,0}^{m}-a\to h_{P,0}-a in V(Pε)V(P^{\varepsilon}) as mm\to\infty.

Similar arguments as in [4, 39], imply that for given (hS,i1m,hP,i1m)(h^{m}_{S,i-1},h^{m}_{P,i-1}), under Assumption 3.1, there exists solution (hS,im,hP,im)(h_{S,i}^{m},h_{P,i}^{m}) to (3.5) and (3.6), for all i=1,,Ni=1,\ldots,N. To derive a priori estimates we consider hS,imh_{S,i}^{m} and hP,imah_{P,i}^{m}-a as test functions in (3.5) and (3.6). Summing over i=1,,ni=1,\ldots,n, for 1nN1\leq n\leq N, and using Assumption 3.1 and inequality (3.4) yields

J=S,P[ΘJ(hJ,nm)L1(Jε)+i=1nτhJ,imL2(Jε)2+kΓε(hJ,nmL2(ΓPε)2hJ,0mL2(ΓPε)2)]\displaystyle\sum_{J=S,P}\!\!\Big{[}\big{\|}{\Theta}_{J}\big{(}h_{J,n}^{m}\big{)}\big{\|}_{L^{1}(J^{\varepsilon})}+\sum_{i=1}^{n}\tau\|\nabla h_{J,i}^{m}\|^{2}_{L^{2}(J^{\varepsilon})}+k_{\Gamma}\varepsilon\big{(}\big{\|}h_{J,n}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}-\big{\|}h_{J,0}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}\big{)}\Big{]}
Cδ+J=S,P(i=1nτδ[hJ,imL2(ΓJ,0ε)2+εhJ,imL2(ΓPε)2]+ΘJ(hJ,0m)L1(Jε)),\displaystyle\leq C_{\delta}+\sum_{J=S,P}\Big{(}\sum_{i=1}^{n}\tau\delta\Big{[}\big{\|}h_{J,i}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{J,0}^{\varepsilon})}+\varepsilon\big{\|}h_{J,i}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}\Big{]}+\big{\|}\Theta_{J}\big{(}h_{J,0}^{m}\big{)}\big{\|}_{L^{1}(J^{\varepsilon})}\Big{)},

for some δ>0\delta>0, where Cauchy’s inequality and properties of telescoping sum imply

i=1nτΓPε(|hS,im|2hP,i1mhS,imhS,i1mhP,im+|hP,im|2)𝑑γ\displaystyle\sum_{i=1}^{n}\tau\int_{\Gamma_{P}^{\varepsilon}}\Big{(}\big{\lvert}h_{S,i}^{m}\big{\rvert}^{2}-h_{P,i-1}^{m}h_{S,i}^{m}-h_{S,i-1}^{m}h_{P,i}^{m}+\big{\lvert}h_{P,i}^{m}\big{\rvert}^{2}\Big{)}d\gamma
τ2(hS,nmL2(ΓPε)2+hP,nmL2(ΓPε)2hS,0mL2(ΓPε)2hP,0mL2(ΓPε)2).\displaystyle\qquad\geq\frac{\tau}{2}\Big{(}\big{\|}h_{S,n}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}+\big{\|}h_{P,n}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}-\big{\|}h_{S,0}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}-\big{\|}h_{P,0}^{m}\big{\|}^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}\Big{)}.

To estimate the boundary integrals, we use the trace theorem, followed by the generalised Poincaré inequality, and the fact that hS,im=0h_{S,i}^{m}=0 on ΓS,L3\Gamma_{S,L_{3}} and hP,im=ah_{P,i}^{m}=a on ΓP,L3\Gamma_{P,L_{3}} to obtain

i=1nτ(εhS,imL2(ΓPε)2+hS,imL2(ΓS,0ε)2)C1i=1nτhS,imL2(Sε)2+τεhS,0mL2(ΓPε)2,\displaystyle\sum_{i=1}^{n}\tau\Big{(}\varepsilon\|h_{S,i}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}+\|h_{S,i}^{m}\|^{2}_{L^{2}(\Gamma^{\varepsilon}_{S,0})}\Big{)}\leq C_{1}\sum_{i=1}^{n}\tau\|\nabla h_{S,i}^{m}\|^{2}_{L^{2}(S^{\varepsilon})}+\tau\varepsilon\|h_{S,0}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}, (3.7)
i=1nτ(εhP,imL2(ΓPε)2+hP,imL2(ΓP,0ε)2)C1i=1nτhP,imL2(Pε)2+C2a2,\displaystyle\sum_{i=1}^{n}\tau\Big{(}\varepsilon\|h_{P,i}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}+\|h_{P,i}^{m}\|^{2}_{L^{2}(\Gamma_{P,0}^{\varepsilon})}\Big{)}\leq C_{1}\sum_{i=1}^{n}\tau\|\nabla h_{P,i}^{m}\|^{2}_{L^{2}(P^{\varepsilon})}+C_{2}a^{2},

where C1,C2>0C_{1},C_{2}>0 may depend on ε\varepsilon. Combining the estimates from above and using assumptions on initial conditions, together with the Poincaré inequality, yield

max1iNΘJ(hJ,im)L1(Jε)C,i=1Nτ[hJ,imL2(Jε)2+hJ,imL2(Jε)2]C,\displaystyle\max\limits_{1\leq i\leq N}\|\Theta_{J}(h_{J,i}^{m})\|_{L^{1}(J^{\varepsilon})}\leq C,\quad\sum_{i=1}^{N}\tau\Big{[}\|h_{J,i}^{m}\|^{2}_{L^{2}(J^{\varepsilon})}+\|\nabla h_{J,i}^{m}\|^{2}_{L^{2}(J^{\varepsilon})}\Big{]}\leq C, (3.8)

for J=S,PJ=S,P, where C>0C>0 may depend on ε\varepsilon. Summing (3.5) over j=i+1,,i+lj=i+1,...,i+l, with l=1,,Nl=1,...,N, and using the properties of the telescoping sum, then taking ϕ=hS,i+lmhS,im\phi=h_{S,i+l}^{m}-h_{S,i}^{m} as a test function, and summing over i=1,,Nli=1,...,N-l, gives

i=1Nl1τθSε(hS,i+lm)θSε(hS,im),hS,i+lmhS,imSε\displaystyle\sum_{i=1}^{N-l}\frac{1}{\tau}\big{\langle}\theta^{\varepsilon}_{S}(h_{S,i+l}^{m})-\theta^{\varepsilon}_{S}(h_{S,i}^{m}),h_{S,i+l}^{m}-h_{S,i}^{m}\big{\rangle}_{S^{\varepsilon}}
i=1Nlj=i+1i+l|KS(hS,j1m)(hS,jm+e3),(hS,i+lmhS,im)Sε|\displaystyle\leq\sum_{i=1}^{N-l}\sum_{j=i+1}^{i+l}\Big{|}\big{\langle}K_{S}(h_{S,j-1}^{m})(\nabla h_{S,j}^{m}+{e}_{3}),\nabla\big{(}h_{S,i+l}^{m}-h_{S,i}^{m}\big{)}\big{\rangle}_{S^{\varepsilon}}\Big{|}
+i=1Nlj=i+1i+l|εkΓhS,jmhP,j1m,hS,i+lmhS,imΓPε+f(hS,j1m),hS,i+lmhS,imΓS,0ε|.\displaystyle\quad+\sum_{i=1}^{N-l}\sum_{j=i+1}^{i+l}\Big{|}\varepsilon k_{\Gamma}\big{\langle}h_{S,j}^{m}-h_{P,j-1}^{m},h_{S,i+l}^{m}-h_{S,i}^{m}\big{\rangle}_{\Gamma_{P}^{\varepsilon}}+\big{\langle}f(h_{S,j-1}^{m}),h_{S,i+l}^{m}-h_{S,i}^{m}\big{\rangle}_{\Gamma_{S,0}^{\varepsilon}}\Big{|}.

Performing similar calculations with ϕ=hP,i+lmhP,im\phi=h_{P,i+l}^{m}-h_{P,i}^{m} as test function in (3.6) and using estimates on hS,jmh_{S,j}^{m} and hP,jmh_{P,j}^{m} in (3.8) imply

i=1NlτθJ(hJ,i+lm)θJ(hJ,im),hJ,i+lmhJ,imJεClτ, for J=S,P.\sum_{i=1}^{N-l}\tau\big{\langle}{\theta}_{J}(h_{J,i+l}^{m})-{\theta}_{J}(h_{J,i}^{m}),h_{J,i+l}^{m}-h_{J,i}^{m}\big{\rangle}_{J^{\varepsilon}}\leq Cl\tau,\qquad\text{ for }\;J=S,P. (3.9)

To pass to the limit in (3.5) and (3.6) as m,Nm,N\to\infty we consider the piecewise constant interpolations in time

h¯J,Nm(x,t)=hJ,im(x) for t(τ(i1),τi],i=1,,N,\overline{h}_{J,N}^{m}(x,t)=h_{J,i}^{m}(x)\qquad\text{ for }\;\;t\in(\tau(i-1),~{}\tau i],\;\;\;i=1,...,N, (3.10)

and h^J,Nm(x,t)=h¯J,Nm(x,tτ)=hJ,i1m(x)\hat{h}_{J,N}^{m}(x,t)=\overline{h}_{J,N}^{m}(x,t-\tau)=h_{J,i-1}^{m}(x), where h^J,Nm(t)=hJ,0m\hat{h}_{J,N}^{m}(t)=h_{J,0}^{m} for t(0,τ]t\in(0,\tau] and J=S,PJ=S,P. Then (3.8) and (3.9) yield

θJ(h¯J,Nm(+λ))θJ(h¯J,Nm),h¯J,Nm(+λ)h¯J,Nm(0,Tλ)×JεCλ,\displaystyle\Big{\langle}{\theta}_{J}(\overline{h}_{J,N}^{m}(\cdot+\lambda))-{\theta}_{J}(\overline{h}_{J,N}^{m}),\,\overline{h}_{J,N}^{m}(\cdot+\lambda)-\overline{h}_{J,N}^{m}\Big{\rangle}_{(0,T-\lambda)\times J^{\varepsilon}}\leq C\lambda, (3.11)
sup0tTΘJ(h¯J,Nm(t))L1(Jε)+h¯J,NmL2(0,T;H1(Jε))+h^J,NmL2(0,T;H1(Jε))C,\displaystyle\sup_{0\leq t\leq T}\big{\|}{\Theta}_{J}(\overline{h}_{J,N}^{m}(t))\big{\|}_{L^{1}(J^{\varepsilon})}+\big{\|}\overline{h}_{J,N}^{m}\big{\|}_{L^{2}(0,T;H^{1}(J^{\varepsilon}))}+\big{\|}\hat{h}_{J,N}^{m}\big{\|}_{L^{2}(0,T;H^{1}(J^{\varepsilon}))}\leq C,

for J=S,PJ=S,P, uniform with respect to mm and NN, where λ(lτ,(l+1)τ)\lambda\in(l\tau,(l+1)\tau) for some l=l= 0,,N10,...,N-1. Using a priori estimates (3.11) and equations (3.5) and (3.6) we obtain

tτθSε(,h¯S,Nm)L2(0,T;V(Sε))+tτθP(h¯P,Nm)L2(0,T;V(Pε))C,\big{\|}\partial_{t}^{\tau}{\theta}^{\varepsilon}_{S}\big{(}\cdot,\overline{h}_{S,N}^{m}\big{)}\big{\|}_{L^{2}(0,T;V(S^{\varepsilon})^{\prime})}+\big{\|}\partial_{t}^{\tau}{\theta}_{P}\big{(}\overline{h}_{P,N}^{m}\big{)}\big{\|}_{L^{2}(0,T;V(P^{\varepsilon})^{\prime})}\leq C, (3.12)

where

tτθJ(h¯J,Nm)=1τ[θJ(h¯J,Nm)θJ(h^J,Nm)], for J=S,P.\partial_{t}^{\tau}{\theta}_{J}(\overline{h}_{J,N}^{m})=\frac{1}{\tau}\Big{[}{\theta}_{J}(\overline{h}_{J,N}^{m})-{\theta}_{J}(\hat{h}_{J,N}^{m})\Big{]},\qquad\text{ for }J=S,P.

Combining estimates (3.11) and (3.12) yields

h^J,Nmh^J,h¯J,NmhJ\displaystyle\hat{h}_{J,N}^{m}\rightharpoonup\hat{h}_{J},\quad\overline{h}_{J,N}^{m}\rightharpoonup h_{J}  weakly in L2(0,T;H1(Jε)),\displaystyle\;\text{ weakly in }L^{2}(0,T;H^{1}(J^{\varepsilon})), J=S,P,\displaystyle J=S,P, (3.13)
tτθJ(h¯J,Nm)μJ\displaystyle\partial_{t}^{\tau}{\theta}_{J}(\overline{h}_{J,N}^{m})\rightharpoonup\mu_{J}  weakly in L2(0,T;V(Jε)),\displaystyle\;\text{ weakly in }L^{2}(0,T;V(J^{\varepsilon})^{\prime}), as m,N.\displaystyle\text{ as }m,N\to\infty.

The boundedness, Lipschitz continuity and monotonicity of θJ{\theta}_{J} imply

θJ(h¯J,Nm)L2((0,T)×Jε)C,\displaystyle\big{\|}{\theta}_{J}(\overline{h}_{J,N}^{m})\big{\|}_{L^{2}((0,T)\times J^{\varepsilon})}\leq C, (3.14)
θJ(h¯J,Nm)L2((0,T)×Jε)LθJh¯J,NmL2((0,T)×Jε),\displaystyle\big{\|}\nabla{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)}\big{\|}_{L^{2}((0,T)\times J^{\varepsilon})}\leq L_{\theta_{J}}\|\nabla\overline{h}_{J,N}^{m}\|_{L^{2}((0,T)\times J^{\varepsilon})},
θJ(h¯J,Nm(+λ))θJ(h¯J,Nm)L2((0,Tλ)×Jε)2\displaystyle\big{\|}{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}(\cdot+\lambda)\big{)}-{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)}\big{\|}_{L^{2}((0,T-\lambda)\times J^{\varepsilon})}^{2}
LθJθJ(h¯J,Nm(+λ))θJ(h¯J,Nm),h¯J,Nm(+λ)h¯J,NmJTλεCλ,\displaystyle\quad\leq L_{\theta_{J}}\big{\langle}\theta_{J}\big{(}\overline{h}_{J,N}^{m}(\cdot+\lambda)\big{)}-{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)},\overline{h}_{J,N}^{m}(\cdot+\lambda)-\overline{h}_{J,N}^{m}\big{\rangle}_{J^{\varepsilon}_{T-\lambda}}\leq C\lambda,

for J=P,R,BJ=P,R,B, where C>0C>0 is independent of mm and NN and LθJL_{\theta_{J}} are the Lipschitz constants for θJ\theta_{J}. Using [41, Theorem 1], there exists zJL2((0,T)×Jε)z_{J}\in~{}L^{2}((0,T)\times J^{\varepsilon}), for J=P,R,BJ=P,R,B, such that, upto a subsequence,

θJ(h¯J,Nm)zJstrongly inL2((0,T)×Jε) as m,N.\displaystyle{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)}\to z_{J}\qquad\text{strongly in}~{}L^{2}((0,T)\times J^{\varepsilon})\quad\text{ as }\;m,N\to\infty. (3.15)

Since θJ{\theta}_{J} is strictly increasing and continuous, it admits a continuous inverse and

h¯J,Nm=θJ1(θJ(h¯J,Nm))θJ1(zJ) a.e. in (0,T)×Jε as m,N,\overline{h}_{J,N}^{m}={\theta}_{J}^{-1}\big{(}{\theta}_{J}(\overline{h}_{J,N}^{m})\big{)}\to{\theta}_{J}^{-1}(z_{J})\quad\text{ a.e.~{}in }\;(0,T)\times J^{\varepsilon}\quad\text{ as }\;m,N\to\infty, (3.16)

and, because h¯J,NmhJ\overline{h}_{J,N}^{m}\rightharpoonup h_{J} in L2(0,T;H1(Jε))L^{2}(0,T;H^{1}(J^{\varepsilon})) as m,Nm,N\to\infty, we have θJ1(zJ)=hJ{\theta}_{J}^{-1}\big{(}z_{J}\big{)}=h_{J} and zJ=θJ(hJ)z_{J}={\theta}_{J}\big{(}h_{J}\big{)}, for J=P,R,BJ=P,R,B. Applying the change of variables t=t+λt=t+\lambda, with λ=τ\lambda=\tau, in the last estimate in (3.14) and using the monotonicity and Lipschitz continuity of θJ\theta_{J} yield

θJ(h¯J,Nm)θJ(h^J,Nm)L2((0,T)×Jε)2LθJ[τθJ(hJ,1m)θJ(hJ,0m),hJ,1mhJ,0mJε\displaystyle\big{\|}{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)}-{\theta}_{J}\big{(}\hat{h}_{J,N}^{m}\big{)}\big{\|}_{L^{2}((0,T)\times J^{\varepsilon})}^{2}\leq L_{\theta_{J}}\Big{[}\tau\langle{\theta}_{J}(h_{J,1}^{m})-{\theta}_{J}(h_{J,0}^{m}),h_{J,1}^{m}-h_{J,0}^{m}\rangle_{J^{\varepsilon}} (3.17)
+τTθJ(h¯J,Nm)θJ(h^J,Nm),h¯J,Nmh^J,NmJεdt]Cτ12,\displaystyle+\int_{\tau}^{T}\!\!\!\langle{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)}-{\theta}_{J}\big{(}\hat{h}_{J,N}^{m}\big{)},\overline{h}_{J,N}^{m}-\hat{h}_{J,N}^{m}\rangle_{J^{\varepsilon}}dt\Big{]}\leq C\tau^{\frac{1}{2}},

for J=P,R,BJ=P,R,B. Thus, using (3.16), we have

θJ(h^J,Nm)θJ(hJ)strongly in L2((0,T)×Jε) as m,N,{\theta}_{J}\big{(}\hat{h}_{J,N}^{m}\big{)}\to{\theta}_{J}\big{(}h_{J}\big{)}\qquad\text{strongly in }L^{2}\big{(}(0,T)\times J^{\varepsilon}\big{)}\quad\text{ as }\;m,N\to\infty, (3.18)

and h^J,Nm=θJ1(θJ(h^J,Nm))θJ1(θJ(hJ))=hJ\hat{h}_{J,N}^{m}=\theta_{J}^{-1}(\theta_{J}(\hat{h}_{J,N}^{m}))\to\theta_{J}^{-1}(\theta_{J}(h_{J}))=h_{J} a.e. in (0,T)×Jε(0,T)\times J^{\varepsilon} as m,Nm,N\to\infty, and hence

h^J,NmhJweakly in L2(0,T;H1(Jε)) as m,N, for J=P,S.\hat{h}_{J,N}^{m}\rightharpoonup h_{J}\qquad\text{weakly in }\;L^{2}\big{(}0,T;H^{1}(J^{\varepsilon})\big{)}\quad\text{ as }\;m,N\to\infty,\;\text{ for }\;J=P,S. (3.19)

Boundedness of θJ\theta_{J} ensures tτθJ(h¯J,Nm)L2((0,T)×Jε)\partial_{t}^{\tau}{\theta}_{J}(\overline{h}_{J,N}^{m})\in L^{2}((0,T)\times J^{\varepsilon}) and

0Ttτ(θJ(h¯J,Nm)),wV(Jε)𝑑t0TθJ(hJ),twJε𝑑t=0TθJ(hJ),twV(Jε)𝑑t,\displaystyle\int_{0}^{T}\!\!\!\big{\langle}\partial_{t}^{\tau}\big{(}{\theta}_{J}\big{(}\overline{h}_{J,N}^{m}\big{)}\big{)},w\big{\rangle}_{V(J^{\varepsilon})^{\prime}}dt\to-\int_{0}^{T}\!\!\!\big{\langle}{\theta}_{J}\big{(}h_{J}\big{)},\partial_{t}w\rangle_{J^{\varepsilon}}dt=-\int_{0}^{T}\!\!\!\big{\langle}{\theta}_{J}\big{(}h_{J}\big{)},\partial_{t}w\big{\rangle}_{V(J^{\varepsilon})^{\prime}}dt, (3.20)

as m,Nm,N\to\infty, for wC0((0,T);CΓJ,L3(Jε¯))w\in C_{0}^{\infty}((0,T);C_{\Gamma_{J,L_{3}}}^{\infty}(\overline{J^{\varepsilon}})) where CΓJ,L3(J¯)={vC(J¯)|v=0 on ΓJ,L3ε}C_{\Gamma_{J,L_{3}}}^{\infty}(\overline{J})=\{v\in C^{\infty}(\overline{J})\;|\;v=0\text{ on }\Gamma^{\varepsilon}_{J,L_{3}}\} and J=P,SJ=P,S. Considering w=κφw=\kappa\varphi with κC0(0,T)\kappa\in C^{\infty}_{0}(0,T) and φCΓJ,L3(Jε¯)\varphi\in C_{\Gamma_{J,L_{3}}}^{\infty}(\overline{J^{\varepsilon}}) and using the last two convergence results in (3.13), together with the definition of the weak derivative, yields μJ=tθJ(hJ)\mu_{J}=\partial_{t}{\theta}_{J}\big{(}h_{J}\big{)} for J=P,SJ=P,S. Continuity of KJK_{J} implies KJ(h^J,Nm)KJ(hJ)K_{J}\big{(}\hat{h}_{J,N}^{m}\big{)}\to K_{J}(h_{J}) a.e. in (0,T)×Jε(0,T)\times J^{\varepsilon} and, since |KJ(h^J,Nm)|KJ,sat|K_{J}(\hat{h}_{J,N}^{m})|\leq K_{J,\text{sat}}, by the dominated convergence theorem we have

KJ(h^J,Nm)KJ(hJ)\displaystyle K_{J}(\hat{h}_{J,N}^{m})\to K_{J}(h_{J}) strongly in Lp((0,T)×Jε),\displaystyle\text{strongly in }L^{p}((0,T)\times J^{\varepsilon}), (3.21)
KJ(h^J,Nm)(h¯J,Nm+e3)KJ(hJ)(hJ+e3)\displaystyle K_{J}(\hat{h}_{J,N}^{m})(\nabla\overline{h}_{J,N}^{m}+{e}_{3})\rightharpoonup K_{J}(h_{J})(\nabla h_{J}+{e}_{3}) weakly in Lq((0,T)×Jε)3,\displaystyle\text{weakly in }L^{q}((0,T)\times J^{\varepsilon})^{3},

as m,Nm,N\to\infty, with p>2p>2 and q=2p/(p+2)q=2p/(p+2), and J=P,SJ=P,S. Using the trace inequality applied to |θS(h^S,nm)θS(hS)|2|{\theta}_{S}\big{(}\hat{h}_{S,n}^{m}\big{)}-{\theta}_{S}\big{(}h_{S}\big{)}|^{2}, along with the Lipschitz continuity of θS\theta_{S} and boundedness of h^S,Nm\hat{h}_{S,N}^{m} in L2(0,T;H1(Sε))L^{2}(0,T;H^{1}(S^{\varepsilon})), yields

θS(h^S,Nm)θS(hS)L2((0,T)×ΓS,0ε)2C[θS(h^S,Nm)θS(hS)L2(STε)2\displaystyle\big{\|}{\theta}_{S}\big{(}\hat{h}_{S,N}^{m}\big{)}-{\theta}_{S}\big{(}h_{S}\big{)}\big{\|}_{L^{2}((0,T)\times\Gamma^{\varepsilon}_{S,0})}^{2}\leq C\Big{[}\big{\|}{\theta}_{S}\big{(}\hat{h}_{S,N}^{m}\big{)}-{\theta}_{S}\big{(}h_{S}\big{)}\big{\|}_{L^{2}(S^{\varepsilon}_{T})}^{2}
+θS(h^S,Nm)θS(hS)L2(STε)(h^S,NmL2(STε)+hSL2(STε))].\displaystyle+\|{\theta}_{S}\big{(}\hat{h}_{S,N}^{m}\big{)}-{\theta}_{S}\big{(}h_{S}\big{)}\big{\|}_{L^{2}(S^{\varepsilon}_{T})}\big{(}\|\nabla\hat{h}_{S,N}^{m}\|_{L^{2}(S^{\varepsilon}_{T})}+\|\nabla h_{S}\|_{L^{2}(S^{\varepsilon}_{T})}\big{)}\Big{]}.

Combining this with the convergence of θS(h^S,Nm){\theta}_{S}(\hat{h}_{S,N}^{m}) in L2((0,T)×Sε)L^{2}((0,T)\times S^{\varepsilon}) implies

θS(h^S,Nm)θS(hS)strongly in L2((0,T)×ΓS,0ε),{\theta}_{S}(\hat{h}_{S,N}^{m})\to{\theta}_{S}(h_{S})\qquad\text{strongly in }L^{2}((0,T)\times\Gamma^{\varepsilon}_{S,0}), (3.22)

and h^S,Nm=θS1(θS(h^S,Nm))θS1(θS(hS))=hS\hat{h}_{S,N}^{m}={\theta}_{S}^{-1}({\theta}_{S}(\hat{h}_{S,N}^{m}))\to{\theta}_{S}^{-1}({\theta}_{S}(h_{S}))=h_{S} a.e. in (0,T)×ΓS,0ε(0,T)\times\Gamma^{\varepsilon}_{S,0}, as m,Nm,N\to\infty. Then the continuity and boundedness of ff ensure the convergence of f(h^S,Nm)f(\hat{h}_{S,N}^{m}).

Taking ϕC0((0,T);CΓS,L3(Sε¯))\phi\in C_{0}^{\infty}((0,T);C_{\Gamma_{S,L_{3}}}^{\infty}(\overline{S^{\varepsilon}})) and ψC0((0,T);CΓP,L3(Pε¯))\psi\in C_{0}^{\infty}((0,T);C_{\Gamma_{P,L_{3}}}^{\infty}(\overline{P^{\varepsilon}})) as test functions in (3.5) and (3.6), respectively, and considering the limits as m,Nm,N\to\infty, we obtain that (hS,hP)(h_{S},h_{P}) is a weak solution of (2.1), (2.3) (2.2), and (2.4). ∎

Remark 3.6.

In some models no water flux at root tips has been considered, i.e. KP(hP)(hP+e3)ν=0-K_{P}(h_{P})(\nabla h_{P}+{e}_{3})\cdot\nu=0 on (0,T]×ΓP,L3ε(0,T]\times\Gamma^{\varepsilon}_{P,L_{3}}. Theorem 3.5 holds also for such boundary conditions and the only difference is in the estimation of the boundary integral. From (3.6), applying the trace theorem and Poincaré inequality, we obtain

ΘP(hP,nm)L1(Pε)+i=1nτ[hP,imL2(Pε)2+εhP,imL2(ΓPε)2]\displaystyle\big{\|}{\Theta}_{P}\big{(}h_{P,n}^{m}\big{)}\big{\|}_{L^{1}(P^{\varepsilon})}+\sum_{i=1}^{n}\tau\Big{[}\|\nabla h_{P,i}^{m}\|^{2}_{L^{2}(P^{\varepsilon})}+\varepsilon\|h_{P,i}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}\Big{]}
C[1+i=1nτεhS,i1mL2(ΓPε)2+ΘP(hP,0m)L1(Pε)].\displaystyle\qquad\leq C\Big{[}1+\sum_{i=1}^{n}\tau\varepsilon\|h_{S,i-1}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}+\big{\|}{\Theta}_{P}\big{(}h_{P,0}^{m}\big{)}\big{\|}_{L^{1}(P^{\varepsilon})}\Big{]}.

Thus in the equation for hS,imh_{S,i}^{m} the boundary integral involving hP,imh_{P,i}^{m} can be estimated

i=1nτεhP,imL2(ΓPε)2C[1+i=1nτεhS,i1mL2(ΓPε)2]C[1+i=1nτhS,imL2(Sε)2].\sum_{i=1}^{n}\tau\varepsilon\|h_{P,i}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}\leq C\Big{[}1+\sum_{i=1}^{n}\tau\varepsilon\|h_{S,i-1}^{m}\|^{2}_{L^{2}(\Gamma_{P}^{\varepsilon})}\Big{]}\leq C\Big{[}1+\sum_{i=1}^{n}\tau\|\nabla h_{S,i}^{m}\|^{2}_{L^{2}(S^{\varepsilon})}\Big{]}.
Theorem 3.7.

Under Assumption 3.1 and, additionally, if ff is non-decreasing and KRK_{R}, KBK_{B}, and KPK_{P} are Lipschitz continuous, the weak solution (hSε,hPε)(h_{S}^{\varepsilon},h_{P}^{\varepsilon}) of model (2.1), (2.3), (2.2), and (2.4) is unique.

The proof of Theorem 3.7 employs the same method as in [38]. For this we first prove two inequalities. Consider σδ+,σδ:[0,1]\sigma_{\delta}^{+},\sigma_{\delta}^{-}:\mathbb{R}\to[0,1] and ηJ,δ+\eta_{J,\delta}^{+}ηJ,δ:[0,)\eta_{J,\delta}^{-}:\mathbb{R}\to[0,\infty), for δ>0\delta>0, given by

σδ+(z)={1for zδ,zδfor 0<z<δ,0for z0,σδ(z)=σδ+(z),ηJ,δ±(h,v0)=v0hσδ±(zv0)θJ(z)𝑑z,\displaystyle\sigma_{\delta}^{+}(z)=\begin{cases}1&\text{for }z\geq\delta,\\ \frac{z}{\delta}&\text{for }0<z<\delta,\\ 0&\text{for }z\leq 0,\end{cases}\qquad\quad\begin{aligned} &\sigma_{\delta}^{-}(z)=-\sigma_{\delta}^{+}(-z),\\ &\eta_{J,\delta}^{\pm}\big{(}h,v^{0}\big{)}=\int_{v^{0}}^{h}\sigma_{\delta}^{\pm}(z-v^{0}){\theta}_{J}^{{}^{\prime}}(z)dz,\end{aligned} (3.23)

for J=P,R,BJ=P,R,B, where v0V(Sε)v^{0}\in V(S^{\varepsilon}) if J=R,BJ=R,B and v0aV(Pε)v^{0}-a\in V(P^{\varepsilon}) if J=PJ=P. Then for κC0(,T)\kappa\in C^{\infty}_{0}(-\infty,T) and for v0V(Sε)v^{0}\in V(S^{\varepsilon}) and w0aV(Pε)w^{0}-a\in V(P^{\varepsilon}) we have σδ±(hSv0)κL2(0,T;V(Sε))\sigma_{\delta}^{\pm}(h_{S}-v^{0})\kappa\in L^{2}(0,T;V(S^{\varepsilon})) and σδ±(hPw0)κL2(0,T;V(Pε))\sigma_{\delta}^{\pm}(h_{P}-w^{0})\kappa\in L^{2}(0,T;V(P^{\varepsilon})).

Similar as in the proof of Theorem 3.5, we omit the superscript ε\varepsilon in hSεh_{S}^{\varepsilon} and hPεh^{\varepsilon}_{P}.

Lemma 3.8.

Let (hS,hP)(h_{S},h_{P}) be a solution to  (2.1), (2.3), (2.2), and (2.4) and let κC0(,T)\kappa\in C^{\infty}_{0}(-\infty,T) be non-negative. Then for v0V(Sε)v^{0}\in V(S^{\varepsilon}), w0aV(Pε)w^{0}-a\in V(P^{\varepsilon}), and any δ>0\delta>0, we have

0T[ηS,δ+(hS,0,v0)ηS,δ+(hS,v0),dκdtSε+ηP,δ+(hP,0,w0)ηP,δ+(hP,w0),dκdtPε\displaystyle\int_{0}^{T}\!\!\Big{[}\Big{\langle}\eta_{S,\delta}^{+}\big{(}h_{S,0},v^{0}\big{)}-\eta_{S,\delta}^{+}(h_{S},v^{0}),\frac{d\kappa}{dt}\Big{\rangle}_{S^{\varepsilon}}+\Big{\langle}\eta_{P,\delta}^{+}\big{(}h_{P,0},w^{0}\big{)}-\eta_{P,\delta}^{+}(h_{P},w^{0}),\frac{d\kappa}{dt}\Big{\rangle}_{P^{\varepsilon}} (3.24)
+[KS(hS)(hS+e3),σδ+(hSv0)Sε+KP(hP)(hP+e3),σδ+(hPw0)Pε\displaystyle+\Big{[}\big{\langle}K_{S}(h_{S})(\nabla h_{S}+{e}_{3}),\!\nabla\sigma_{\delta}^{+}(h_{S}-v^{0})\big{\rangle}_{S^{\varepsilon}}\!\!+\big{\langle}K_{P}(h_{P})(\nabla h_{P}+{e}_{3}),\!\nabla\sigma_{\delta}^{+}(h_{P}-w^{0})\big{\rangle}_{P^{\varepsilon}}
+εkΓhShP,σδ+(hSv0)σδ+(hPw0)ΓPε\displaystyle\qquad+\varepsilon k_{\Gamma}\big{\langle}h_{S}-h_{P},\sigma_{\delta}^{+}(h_{S}-v^{0})-\sigma_{\delta}^{+}(h_{P}-w^{0})\big{\rangle}_{\Gamma_{P}^{\varepsilon}}
+f(hS),σδ+(hSv0)ΓS,0ε+𝒯pot,σδ+(hPw0)ΓP,0ε]κ(t)]dt0.\displaystyle+\big{\langle}f(h_{S}),\sigma_{\delta}^{+}(h_{S}-v^{0})\big{\rangle}_{\Gamma_{S,0}^{\varepsilon}}+\big{\langle}\mathcal{T}_{\text{pot}},\sigma_{\delta}^{+}(h_{P}-w^{0})\big{\rangle}_{\Gamma_{P,0}^{\varepsilon}}\Big{]}\kappa(t)\Big{]}dt\leq 0.

The same inequality holds with ηS,δ(hS,v0)\eta_{S,\delta}^{-}(h_{S},v^{0}) and ηP,δ(hP,w0)\eta_{P,\delta}^{-}(h_{P},w^{0}) instead of ηS,δ+(hS,v0)\eta_{S,\delta}^{+}(h_{S},v^{0}) and ηP,δ+(hP,w0)\eta_{P,\delta}^{+}(h_{P},w^{0}), and with σδ(hSv0)\sigma_{\delta}^{-}(h_{S}-v^{0}) and σδ(hPw0)\sigma_{\delta}^{-}(h_{P}-w^{0}) instead of σδ+(hSv0)\sigma_{\delta}^{+}(h_{S}-v^{0}) and σδ+(hPw0)\sigma_{\delta}^{+}(h_{P}-w^{0}).

Proof.

We shall prove (3.24) and the proof for ηS,δ(hS,v0)\eta_{S,\delta}^{-}(h_{S},v^{0}), ηP,δ(hP,w0)\eta_{P,\delta}^{-}(h_{P},w^{0}), σδ(hSv0)\sigma_{\delta}^{-}(h_{S}-v^{0}), σδ(hPw0)\sigma_{\delta}^{-}(h_{P}-w^{0}) follows the same lines. Considering

ζτ(t,x)=1τtt+τζ(s,x)𝑑s, where ζ=σδ+(hSv0)κL2(0,T;V(Sε)),\zeta_{\tau}(t,x)=\frac{1}{\tau}\int_{t}^{t+\tau}\zeta(s,x)ds,\;\;\text{ where }\;\zeta=\sigma_{\delta}^{+}(h_{S}-v^{0})\kappa\in L^{2}(0,T;V(S^{\varepsilon})), (3.25)

for t[0,T]t\in[0,T] and τ>0\tau>0, and using κ(T)=0\kappa(T)=0 we can write

0TtθS(hS),ζτV(Sε)𝑑t=0Tτ1τθS(hS,0)θS(hS(t)),ζ(t+τ)Sε𝑑t\displaystyle\int_{0}^{T}\!\!\big{\langle}\partial_{t}{\theta}_{S}(h_{S}),\zeta_{\tau}\big{\rangle}_{V^{\prime}(S^{\varepsilon})}dt=\int_{0}^{T-\tau}\!\!\frac{1}{\tau}\big{\langle}{\theta}_{S}(h_{S,0})-{\theta}_{S}(h_{S}(t)),\zeta(t+\tau)\big{\rangle}_{S^{\varepsilon}}dt (3.26)
0T1τθS(hS,0)θS(hS(t)),ζ(t)Sε𝑑t=0T1τθS(hS(t))θS(hS(tτ)),ζ(t)Sε𝑑t,\displaystyle\quad-\int_{0}^{T}\!\!\frac{1}{\tau}\big{\langle}{\theta}_{S}(h_{S,0})-{\theta}_{S}(h_{S}(t)),\zeta(t)\big{\rangle}_{S^{\varepsilon}}dt=\int_{0}^{T}\!\!\frac{1}{\tau}\big{\langle}{\theta}_{S}(h_{S}(t))-{\theta}_{S}(h_{S}(t-\tau)),\zeta(t)\rangle_{S^{\varepsilon}}dt,

where θS(hS){\theta}_{S}(h_{S}) is extended by θS(hS)=θS(hS,0){\theta}_{S}(h_{S})={\theta}_{S}(h_{S,0}) for t<0t<0. Since θR,θB\theta_{R},\theta_{B} are Lipschitz continuous and θR,θB\theta_{R},\theta_{B} and σδ+\sigma_{\delta}^{+} are monotone increasing, it follows that

ηS,δ+(hS(t),v0)ηS,δ+(hS(tτ),v0)=hS(tτ)hS(t)σδ+(zv0)θS(z)𝑑z\displaystyle\eta_{S,\delta}^{+}\big{(}h_{S}(t),v^{0}\big{)}-\eta_{S,\delta}^{+}\big{(}h_{S}(t-\tau),v^{0}\big{)}=\int_{h_{S}(t-\tau)}^{h_{S}(t)}\sigma_{\delta}^{+}(z-v^{0}){\theta}_{S}^{{}^{\prime}}(z)dz (3.27)
σδ+(hS(t)v0)[θS(hS(t))θS(hS(tτ))].\displaystyle\leq\sigma_{\delta}^{+}\big{(}h_{S}(t)-v^{0}\big{)}\big{[}{\theta}_{S}(h_{S}(t))-{\theta}_{S}(h_{S}(t-\tau))\big{]}.

Using (3.27) in (3.26), together with the non-negativity of κC0(,T)\kappa\in C^{\infty}_{0}(-\infty,T), yields

lim infτ00TtθS(hS),ζτV(Sε)𝑑t\displaystyle\liminf_{\tau\to 0}\int_{0}^{T}\!\!\!\big{\langle}\partial_{t}{\theta}_{S}(h_{S}),\zeta_{\tau}\big{\rangle}_{V^{\prime}(S^{\varepsilon})}dt (3.28)
lim infτ00T1τηS,δ+(hS(t),v0)ηS,δ+(hS(tτ),v0),κ(t)Sε𝑑t\displaystyle\geq\liminf_{\tau\to 0}\int_{0}^{T}\frac{1}{\tau}\big{\langle}\eta_{S,\delta}^{+}(h_{S}(t),v^{0})-\eta_{S,\delta}^{+}(h_{S}(t-\tau),v^{0}),\kappa(t)\big{\rangle}_{S^{\varepsilon}}dt
=lim infτ00TηS,δ+(hS,0,v0)ηS,δ+(hS(t),v0),1τ(κ(t+τ)κ(t))Sε𝑑t\displaystyle=\liminf_{\tau\to 0}\int_{0}^{T}\big{\langle}\eta_{S,\delta}^{+}(h_{S,0},v^{0})-\eta_{S,\delta}^{+}(h_{S}(t),v^{0}),\frac{1}{\tau}\big{(}\kappa(t+\tau)-\kappa(t)\big{)}\big{\rangle}_{S^{\varepsilon}}dt
=0TηS,δ+(hS,0,v0)ηS,δ+(hS(t),v0),dκdtSε𝑑t.\displaystyle=\int_{0}^{T}\!\!\!\Big{\langle}\eta_{S,\delta}^{+}(h_{S,0},v^{0})-\eta_{S,\delta}^{+}(h_{S}(t),v^{0}),\frac{d\kappa}{dt}\Big{\rangle}_{S^{\varepsilon}}dt.

The same calculations hold for tθP(hP)\partial_{t}\theta_{P}(h_{P}) and ηP,δ+(hP,w0)\eta_{P,\delta}^{+}(h-P,w^{0}), considering ζ=σδ+(hPw0)κ\zeta=\sigma_{\delta}^{+}(h_{P}-w^{0})\kappa in the definition of ζτ\zeta_{\tau} in (3.25). Using the Cauchy-Schwarz inequality, change in the order of integration and that κC0(,T)\kappa\in C^{\infty}_{0}(-\infty,T), we obtain ζτ\zeta_{\tau}, ζτ,tζτL2((0,T)×Sε)\nabla\zeta_{\tau},\partial_{t}\zeta_{\tau}\in L^{2}((0,T)\times S^{\varepsilon}). The Lebesgue differentiation theorem and the regularity of hSL2(0,T;V(Sε))h_{S}\in L^{2}(0,T;V(S^{\varepsilon})) and v0V(Sε)v^{0}\in V(S^{\varepsilon}), imply ζτζ\zeta_{\tau}\to\zeta in L2(0,T;V(Sε))L^{2}(0,T;V(S^{\varepsilon})) and, by applying the trace theorem, also in L2((0,T)×ΓPε)L^{2}((0,T)\times\Gamma^{\varepsilon}_{P}) and L2((0,T)×ΓS,0ε)L^{2}((0,T)\times\Gamma^{\varepsilon}_{S,0}), as τ0\tau\to 0. Hence, testing (3.1) and (3.2) with the corresponding ζτ\zeta_{\tau} and taking the limit as τ0\tau\to 0 yield the inequality stated in the lemma. ∎

Proof of Theorem 3.7.

To prove the uniqueness result, assume there are two solutions (hS,1,hP,1)(h_{S,1},h_{P,1}) and (hS,2,hP,2)(h_{S,2},h_{P,2}), consider a doubling of the time variable (t1,t2)(0,T)2(t_{1},t_{2})\in(0,T)^{2} and define functions hJ,1,hJ,2:(0,T)2×Jεh_{J,1},h_{J,2}:(0,T)^{2}\times J^{\varepsilon}\to\mathbb{R} such that

hJ,1(x,t1,t2)=hJ,1(x,t1)andhJ,2(x,t1,t2)=hJ,2(x,t2), for J=S,P.h_{J,1}(x,t_{1},t_{2})=h_{J,1}(x,t_{1})\quad\text{and}\quad h_{J,2}(x,t_{1},t_{2})=h_{J,2}(x,t_{2}),\quad\text{ for }\;J=S,P.

Considering (3.24) with v0=hS,2(t2)v^{0}=h_{S,2}(t_{2})w0=hP,2(t2)w^{0}=h_{P,2}(t_{2}), and κ(t2):t1κ(t1,t2)\kappa(t_{2}):t_{1}\to\kappa(t_{1},t_{2}), for non-negative κC0((0,T)2)\kappa\in C_{0}^{\infty}((0,T)^{2}), we obtain

0T[[SεηS,δ+(hS,1,hS,2(t2))dx+PεηP,δ+(hP,1,hP,2(t2))dx]dκ(t2)dt1\displaystyle\int_{0}^{T}\!\!\Big{[}-\Big{[}\int_{S^{\varepsilon}}\!\!\eta_{S,\delta}^{+}(h_{S,1},h_{S,2}(t_{2}))dx+\int_{P^{\varepsilon}}\!\!\eta_{P,\delta}^{+}(h_{P,1},h_{P,2}(t_{2}))dx\Big{]}\frac{d\kappa(t_{2})}{dt_{1}} (3.29)
+J=S,PKJ(hJ,1)(hJ,1+e3),σδ+(hJ,1hJ,2(t2))Jεκ(t2)\displaystyle\qquad+\sum_{J=S,P}\big{\langle}K_{J}(h_{J,1})\big{(}\nabla h_{J,1}+{e}_{3}\big{)},\nabla\sigma_{\delta}^{+}(h_{J,1}-h_{J,2}(t_{2}))\big{\rangle}_{J^{\varepsilon}}\kappa(t_{2})
+εkΓhS,1hP,1,σδ+(hS,1hS,2(t2))σδ+(hP,1hP,2(t2))ΓPεκ(t2)\displaystyle\qquad+\varepsilon k_{\Gamma}\big{\langle}h_{S,1}-h_{P,1},\sigma_{\delta}^{+}(h_{S,1}-h_{S,2}(t_{2}))-\sigma_{\delta}^{+}(h_{P,1}-h_{P,2}(t_{2}))\big{\rangle}_{\Gamma_{P}^{\varepsilon}}\kappa(t_{2})
+[f(hS,1),σδ+(hS,1hS,2(t2))ΓS,0ε+𝒯pot,σδ+(hP,1hP,2(t2))ΓP,0ε]κ(t2)]dt10,\displaystyle+\big{[}\big{\langle}f(h_{S,1}),\sigma_{\delta}^{+}(h_{S,1}-h_{S,2}(t_{2}))\big{\rangle}_{\Gamma^{\varepsilon}_{S,0}}+\big{\langle}\mathcal{T}_{\text{pot}},\sigma_{\delta}^{+}(h_{P,1}-h_{P,2}(t_{2}))\big{\rangle}_{\Gamma^{\varepsilon}_{P,0}}\big{]}\kappa(t_{2})\Big{]}dt_{1}\leq 0,

for a.e. t2(0,T)t_{2}\in(0,T), whereas considering v0=hS,1(t1)v^{0}=h_{S,1}(t_{1}), w0=hP,1(t1)w^{0}=~{}h_{P,1}(t_{1}), and κ(t1):t2κ(t1,t2)\kappa(t_{1}):t_{2}\to\kappa(t_{1},t_{2}) in the inequality with σδ\sigma_{\delta}^{-} and ηJ,δ\eta^{-}_{J,\delta}, for J=S,PJ=S,P, yields

0T[[SεηS,δ(hS,2,hS,1(t1))dx+PεηP,δ(hP,2,hP,1(t1))dx]dκ(t1)dt2\displaystyle\int_{0}^{T}\!\!\Big{[}-\Big{[}\int_{S^{\varepsilon}}\!\!\eta_{S,\delta}^{-}(h_{S,2},h_{S,1}(t_{1}))dx+\int_{P^{\varepsilon}}\!\!\eta_{P,\delta}^{-}(h_{P,2},h_{P,1}(t_{1}))dx\Big{]}\frac{d\kappa(t_{1})}{dt_{2}} (3.30)
+J=S,PKJ(hJ,2)(hJ,2+e3),σδ(hJ,2hJ,1(t1))Jεκ(t1)\displaystyle\qquad+\sum_{J=S,P}\big{\langle}K_{J}(h_{J,2})\big{(}\nabla h_{J,2}+{e}_{3}\big{)},\nabla\sigma_{\delta}^{-}(h_{J,2}-h_{J,1}(t_{1}))\big{\rangle}_{J^{\varepsilon}}\,\kappa(t_{1})
+εkΓhS,2hP,2,σδ(hS,2hS,1(t1))σδ(hP,2hP,1(t1))ΓPεκ(t1)\displaystyle\qquad+\varepsilon k_{\Gamma}\big{\langle}h_{S,2}-h_{P,2},\sigma_{\delta}^{-}(h_{S,2}-h_{S,1}(t_{1}))-\sigma_{\delta}^{-}(h_{P,2}-h_{P,1}(t_{1}))\big{\rangle}_{\Gamma_{P}^{\varepsilon}}\kappa(t_{1})
+[f(hS,2),σδ(hS,2hS,1(t1))ΓS,0ε+𝒯pot,σδ(hP,2hP,1(t1))ΓP,0ε]κ(t1)]dt20,\displaystyle+\big{[}\big{\langle}f(h_{S,2}),\sigma_{\delta}^{-}(h_{S,2}-h_{S,1}(t_{1}))\big{\rangle}_{\Gamma^{\varepsilon}_{S,0}}+\big{\langle}\mathcal{T}_{\text{pot}},\sigma_{\delta}^{-}(h_{P,2}-h_{P,1}(t_{1}))\big{\rangle}_{\Gamma^{\varepsilon}_{P,0}}\big{]}\kappa(t_{1})\Big{]}dt_{2}\leq 0,

for a.e. t1(0,T)t_{1}\in(0,T). Integrating (3.29) with respect to t2t_{2} and (3.30) with respect to t1t_{1}, adding the resultant inequalities and using σδ(z)=σδ+(z)\sigma_{\delta}^{-}(z)=-\sigma_{\delta}^{+}(-z), we obtain

J=S,PηJ,δ+(hJ,1,hJ,2)t1κ+ηJ,δ(hJ,2,hJ,1)t2κ,1QJ\displaystyle-\!\!\sum_{J=S,P}\!\!\!\big{\langle}\eta_{J,\delta}^{+}(h_{J,1},h_{J,2})\partial_{t_{1}}\kappa+\eta_{J,\delta}^{-}(h_{J,2},h_{J,1})\partial_{t_{2}}\kappa,1\big{\rangle}_{Q_{J}} (3.31)
+J=S,PKJ(hJ,1)(hJ,1+e3)KJ(hJ,2)(hJ,2+e3),σδ+(hJ,1hJ,2)κQJ\displaystyle+\!\!\sum_{J=S,P}\!\!\!\big{\langle}K_{J}(h_{J,1})\big{(}\nabla h_{J,1}+{e}_{3}\big{)}-K_{J}(h_{J,2})\big{(}\nabla h_{J,2}+{e}_{3}\big{)},\nabla\sigma_{\delta}^{+}(h_{J,1}-h_{J,2})\kappa\big{\rangle}_{Q_{J}}
+εkΓ(hS,1hS,2)(hP,1hP,2),[σδ+(hS,1hS,2)σδ+(hP,1hP,2)]κΓ~Pε\displaystyle+\varepsilon k_{\Gamma}\big{\langle}(h_{S,1}-h_{S,2})-(h_{P,1}-h_{P,2}),\big{[}\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})-\sigma_{\delta}^{+}(h_{P,1}-h_{P,2})\big{]}\kappa\big{\rangle}_{\tilde{\Gamma}_{P}^{\varepsilon}}
+f(hS,1)f(hS,2),σδ+(hS,1hS,2)κ(0,T)2×ΓS,0ε0,\displaystyle+\big{\langle}f(h_{S,1})-f(h_{S,2}),\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})\kappa\big{\rangle}_{(0,T)^{2}\times\Gamma^{\varepsilon}_{S,0}}\leq 0,

where QS=(0,T)2×SεQ_{S}=(0,T)^{2}\times S^{\varepsilon}, QP=(0,T)2×PεQ_{P}=(0,T)^{2}\times P^{\varepsilon}, and Γ~Pε=(0,T)2×ΓPε\tilde{\Gamma}_{P}^{\varepsilon}=(0,T)^{2}\times\Gamma_{P}^{\varepsilon}. Using the definition of σδ+\sigma^{+}_{\delta} we have

[KS(hS,1)(hS,1+e3)KS(hS,2)(hS,2+e3)]σδ+(hS,1hS,2)\displaystyle\big{[}K_{S}(h_{S,1})(\nabla h_{S,1}+{e}_{3})-K_{S}(h_{S,2})(\nabla h_{S,2}+{e}_{3})\big{]}\cdot\nabla\sigma_{\delta}^{+}(h_{S,1}-h_{S,2}) (3.32)
(KS(hS,1)ς)|(hS,1hS,2)|2(σδ+)(hS,1hS,2)\displaystyle\quad\geq(K_{S}(h_{S,1})-\varsigma)|\nabla(h_{S,1}-h_{S,2})|^{2}\big{(}\sigma_{\delta}^{+}\big{)}^{\prime}(h_{S,1}-h_{S,2})
12ς[1+|hS,2|2]|KS(hS,1)KS(hS,2)|2(σδ+)(hS,1hS,2),\displaystyle\qquad-\frac{1}{2\varsigma}\big{[}1+|\nabla h_{S,2}|^{2}\big{]}|K_{S}(h_{S,1})-K_{S}(h_{S,2})|^{2}\ \big{(}\sigma_{\delta}^{+}\big{)}^{\prime}(h_{S,1}-h_{S,2}),

where (σδ+)\big{(}\sigma_{\delta}^{+}\big{)}^{\prime} is non-negative and singular as δ0\delta\downarrow 0. The first term on the right-hand side of (3.32) is non-negative for 0<ςmin{KR,0,KB,0}0<\varsigma\leq\min\{K_{R,0},K_{B,0}\} and the second term is non-zero only if 0<hS,1hS,2<δ0<h_{S,1}-h_{S,2}<\delta with

|KS(hS,1)KS(hS,2)|2(σδ+)(hS,1hS,2)=|KS(hS,1)KS(hS,2)|2δLKS2δ,|K_{S}(h_{S,1})-K_{S}(h_{S,2})|^{2}\big{(}\sigma_{\delta}^{+}\big{)}^{\prime}(h_{S,1}-h_{S,2})=\frac{|K_{S}(h_{S,1})-K_{S}(h_{S,2})|^{2}}{\delta}\leq L^{2}_{K_{S}}\delta,

where LKS=max{LKR,LKB}L_{K_{S}}=\max\{L_{K_{R}},L_{K_{B}}\} and LKJL_{K_{J}} is the Lipschitz constant for KJK_{J}, with J=R,BJ=R,B. A similar estimate holds for [KP(hP,1)(hP,1+e3)KP(hP,2)(hP,2+e3)]σδ+(hP,1hP,2)[K_{P}(h_{P,1})(\nabla h_{P,1}+{e}_{3})-K_{P}(h_{P,2})(\nabla h_{P,2}+{e}_{3})]\cdot\nabla\sigma_{\delta}^{+}(h_{P,1}-h_{P,2}). The definition of σδ+\sigma^{+}_{\delta} yields

[(hS,1hS,2)(hP,1hP,2)][σδ+(hS,1hS,2)σδ+(hP,1hP,2)]\displaystyle\big{[}(h_{S,1}-h_{S,2})-(h_{P,1}-h_{P,2})\big{]}\big{[}\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})-\sigma_{\delta}^{+}(h_{P,1}-h_{P,2})\big{]}
[(hS,1hS,2)+(hP,1hP,2)+][σδ+(hS,1hS,2)σδ+(hP,1hP,2)]=I,\displaystyle\quad\geq\big{[}(h_{S,1}-h_{S,2})^{+}-(h_{P,1}-h_{P,2})^{+}\big{]}\big{[}\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})-\sigma_{\delta}^{+}(h_{P,1}-h_{P,2})\big{]}=I,

where u+=max{u,0}u^{+}=\max\{u,0\}. When hS,1hS,20h_{S,1}-h_{S,2}\leq 0, hP,1hP,20h_{P,1}-h_{P,2}\leq 0, hS,1hS,2δh_{S,1}-h_{S,2}\geq\delta, and hP,1hP,2δh_{P,1}-h_{P,2}\geq\delta we directly have I0I\geq 0. If hS,1hS,2δh_{S,1}-h_{S,2}\geq\delta and 0<hP,1hP,2<δ0<h_{P,1}-h_{P,2}<\delta, then σδ+(hS,1hS,2)=1\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})=1 and σδ+(hP,1hP,2)<1\sigma_{\delta}^{+}(h_{P,1}-h_{P,2})<1, resulting in

I[(hS,1hS,2)+(hP,1hP,2)+][(hS,1hS,2)+(hP,1hP,2)+]=0.I\geq\big{[}(h_{S,1}-h_{S,2})^{+}-(h_{P,1}-h_{P,2})^{+}\big{]}-\big{[}(h_{S,1}-h_{S,2})^{+}-(h_{P,1}-h_{P,2})^{+}\big{]}=0.

In the case where δ>hS,1hS,2>hP,1hP,2>0\delta>h_{S,1}-h_{S,2}>h_{P,1}-h_{P,2}>0 we have σδ+(hS,1hS,2)>σδ+(hP,1hP,2)\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})>\sigma_{\delta}^{+}(h_{P,1}-h_{P,2}) and hence I0I\geq 0. Analogous arguments imply I0I\geq 0 when hP,1hP,2>hS,1hS,2>0h_{P,1}-h_{P,2}>h_{S,1}-h_{S,2}>0.

Using (3.31), the estimates above, and the fact that ff is non-decreasing, yields

J=S,P[ηJ,δ+(hJ,1,hJ,2)t1κ+ηJ,δ(hJ,2,hJ,1)t2κ,1QJ+δC(1+hJ,2κL2(QJ)2)]0.-\!\!\!\sum_{J=S,P}\!\!\Big{[}\big{\langle}\eta_{J,\delta}^{+}(h_{J,1},h_{J,2})\partial_{t_{1}}\kappa+\eta_{J,\delta}^{-}(h_{J,2},h_{J,1})\partial_{t_{2}}\kappa,1\big{\rangle}_{Q_{J}}+\delta C\big{(}1+\|\nabla h_{J,2}\,\kappa\|^{2}_{L^{2}(Q_{J})}\big{)}\Big{]}\leq 0. (3.33)

The monotonicity of θS{\theta}_{S} and σδ+\sigma_{\delta}^{+} and the non-negativity of σδ+\sigma_{\delta}^{+} ensure

hS,2hS,1σδ+(zhS,2)θS(z)𝑑zLp(QS)pσδ+(hS,1hS,2)(θS(hS,1)θS(hS,2))Lp(QS)p,\Big{\|}\int_{h_{S,2}}^{h_{S,1}}\!\!\!\!\!\sigma_{\delta}^{+}(z-h_{S,2}){\theta}_{S}^{{}^{\prime}}(z)dz\Big{\|}^{p}_{L^{p}(Q_{S})}\!\!\!\leq\big{\|}\sigma_{\delta}^{+}(h_{S,1}-h_{S,2})({\theta}_{S}(h_{S,1})-{\theta}_{S}(h_{S,2}))\big{\|}^{p}_{L^{p}(Q_{S})},

for 1<p<1<p<\infty. A similar estimate is obtained for σδ\sigma^{-}_{\delta} by using that σδ(z)=σδ+(z)\sigma^{-}_{\delta}(z)=-\sigma^{+}_{\delta}(-z). The boundedness of θS\theta_{S} and σδ+\sigma^{+}_{\delta} implies ηS,δ+(hS,1,hS,2)\eta_{S,\delta}^{+}(h_{S,1},h_{S,2}) and ηS,δ(hS,2,hS,1)\eta_{S,\delta}^{-}(h_{S,2},h_{S,1}) are uniformly bounded in Lp(QS)L^{p}(Q_{S}) and, since 0σδ+10\leq\sigma_{\delta}^{+}\leq 1, we have

|ηS,δ+(hS,1,hS,2)|(θS(hS,1)θS(hS,2))+,|ηS,δ(hS,2,hS,1)|(θS(hS,1)θS(hS,2))+.|\eta_{S,\delta}^{+}(h_{S,1},h_{S,2})|\leq({\theta}_{S}(h_{S,1})-{\theta}_{S}(h_{S,2}))^{+},\;\;|\eta_{S,\delta}^{-}(h_{S,2},h_{S,1})|\leq({\theta}_{S}(h_{S,1})-{\theta}_{S}(h_{S,2}))^{+}.

Additionally, (θS(hS,1(t1,x))θS(hS,2(t2,x)))+=0({\theta}_{S}(h_{S,1}(t_{1},x))-{\theta}_{S}(h_{S,2}(t_{2},x)))^{+}\!=\!0 if hS,1(t1,x)hS,2(t2,x)h_{S,1}(t_{1},x)\leq h_{S,2}(t_{2},x), and σδ+(zhS,2(t2,x))=0\sigma_{\delta}^{+}(z-h_{S,2}(t_{2},x))=0 for z(hS,1(t1,x),hS,2(t2,x))z\in\big{(}h_{S,1}(t_{1},x),h_{S,2}(t_{2},x)\big{)}, σδ+(zhS,2(t2,x))=1\sigma_{\delta}^{+}(z-h_{S,2}(t_{2},x))=1 for zhS,2+δ(t2,z)z\geq h_{S,2}+\delta(t_{2},z), for (t1,t2,x)QS(t_{1},t_{2},x)\in Q_{S}. Then the Lipschitz continuity of θS{\theta}_{S} implies

|ηS,δ+(hS,1(t1,x),hS,2(t2,x))(θS(hS,1(t1,x))θS(hS,2(t2,x)))+|\displaystyle\big{|}\eta_{S,\delta}^{+}(h_{S,1}(t_{1},x),h_{S,2}(t_{2},x))-\big{(}{\theta}_{S}(h_{S,1}(t_{1},x))-{\theta}_{S}(h_{S,2}(t_{2},x))\big{)}^{+}\big{|}
LθS|hS,2(t2,x)hS,1(t1,x)[σδ+(zhS,2(t2,x))1]𝑑z|\displaystyle\quad\leq L_{{\theta}_{S}}\Big{|}\int_{h_{S,2}(t_{2},x)}^{h_{S,1}(t_{1},x)}\big{[}\sigma_{\delta}^{+}(z-h_{S,2}(t_{2},x))-1\big{]}dz\Big{|} LθSδ.\displaystyle\leq L_{{\theta}_{S}}\delta.

Using similar arguments for hP,1h_{P,1} and hP,2h_{P,2} and the fact that σ(z)=σ+(z)\sigma^{-}(z)=-\sigma^{+}(-z) yields

ηJ,δ+(hJ,1,hJ,2)(θJ(hJ,1)θJ(hJ,2))+,\displaystyle\eta_{J,\delta}^{+}(h_{J,1},h_{J,2})\to\big{(}{\theta}_{J}(h_{J,1})-{\theta}_{J}(h_{J,2})\big{)}^{+}, (3.34)
ηJ,δ(hJ,2,hJ,1)(θJ(hJ,1)θJ(hJ,2))+,\displaystyle\eta_{J,\delta}^{-}(h_{J,2},h_{J,1})\to\big{(}{\theta}_{J}(h_{J,1})-{\theta}_{J}(h_{J,2})\big{)}^{+},

pointwise a.e. in QJQ_{J} as δ0\delta\to 0, for J=S,PJ=S,P. The dominated convergence theorem implies the strong convergence of ηJ,δ+(hJ,1,hJ,2)\eta_{J,\delta}^{+}(h_{J,1},h_{J,2}) and ηJ,δ(hJ,2,hJ,1)\eta_{J,\delta}^{-}(h_{J,2},h_{J,1}) in Lp((0,T)2×Jε)L^{p}((0,T)^{2}\times J^{\varepsilon}), as δ0\delta\to 0, for J=S,PJ=S,P and 1<p<1<p<\infty. Taking in (3.33) the limits as δ0\delta\to 0 we obtain

J=S,PJε(θJ(hJ,1)θJ(hJ,2))+𝑑x,t1κ+t2κ(0,T)20.-\Big{\langle}\!\sum_{J=S,P}\int_{J^{\varepsilon}}\!\!\big{(}{\theta}_{J}(h_{J,1})-{\theta}_{J}(h_{J,2})\big{)}^{+}dx,\partial_{t_{1}}\kappa+\partial_{t_{2}}\kappa\Big{\rangle}_{(0,T)^{2}}\leq 0. (3.35)

For any non-negative κC0(0,T)\kappa\in C_{0}^{\infty}(0,T), there exists ϱ(0,T/2)\varrho^{*}\in(0,T/2) such that κ(t)=0\kappa(t)=0 if 0<t<ϱ0<t<\varrho^{*} or t>Tϱt>T-\varrho^{*}. For a non-negative ϑC0()\vartheta\in C_{0}^{\infty}(\mathbb{R}) of unit mass we have that ϑ(r/ϱ)=0\vartheta(r/{\varrho})=0, for |r|ϱ\lvert r\rvert\geq\varrho and some ϱ>0\varrho>0, and the function

κϱ(t1,t2):=1ϱϑ(t1t2ϱ)κ(t1+t22),\kappa_{\varrho}(t_{1},t_{2}):=\frac{1}{\varrho}\vartheta\Big{(}\frac{t_{1}-t_{2}}{\varrho}\Big{)}\kappa\Big{(}\frac{t_{1}+t_{2}}{2}\Big{)}, (3.36)

is admissible in (3.35), for ϱ<2ϱ\varrho<2\varrho^{*}. Applying the change of variables τ=t1t2\tau=t_{1}-t_{2} and denoting t=t1t=t_{1}, yield

1ϱϑ(τϱ)0TJ=S,PJε(θJ(hJ,1)θJ(hJ,2τ))+𝑑xtκτ2dtdτ0,-\int_{\mathbb{R}}\!\frac{1}{\varrho}\vartheta\Big{(}\frac{\tau}{\varrho}\Big{)}\int_{0}^{T}\!\!\!\sum_{J=S,P}\int_{J^{\varepsilon}}\!\!\!\big{(}{\theta}_{J}(h_{J,1})-{\theta}_{J}(h_{J,2}^{\tau})\big{)}^{+}dx\,\partial_{t}\kappa^{\frac{\tau}{2}}\,dtd\tau\leq 0, (3.37)

with the abbreviation hJτ(t)=hJ(tτ)h_{J}^{\tau}(t)=h_{J}(t-\tau) for J=S,PJ=S,P. To take the limit as τ0\tau\to 0, we first show θJ(hJ,2τ)θJ(hJ,2){\theta}_{J}(h_{J,2}^{\tau})\to{\theta}_{J}(h_{J,2}) strongly in L2((0,T)×Jε)L^{2}((0,T)\times J^{\varepsilon}), where J=S,PJ=S,P. Assume τ>0\tau>0 and consider

ζ(t):=(hS,2(t)hS,2(tτ))χ[0,T)(t), and ζτ(t):=1τtt+τζ(s)𝑑s.\zeta(t):=(h_{S,2}(t)-h_{S,2}(t-\tau))\chi_{[0,T)}(t),\quad\text{ and }\quad\zeta_{\tau}(t):=\frac{1}{\tau}\int_{t}^{t+\tau}\zeta(s)ds.

The regularity of hS,2h_{S,2} ensures that ζτ\zeta_{\tau} is an admissible test function in (3.1). Using an integration by parts and the regularity of ζτ\zeta_{\tau}, we obtain

tθS(hS,2),ζτV(Sε),T=1τ(θ(hS,2)θ(hS,2(τ))),hS,2hS,2(τ)STε,\big{\langle}\partial_{t}{\theta}_{S}(h_{S,2}),\zeta_{\tau}\big{\rangle}_{V(S^{\varepsilon})^{\prime},T}=\Big{\langle}\frac{1}{\tau}\big{(}{\theta}(h_{S,2})-{\theta}(h_{S,2}(\cdot-\tau))\big{)},h_{S,2}-h_{S,2}(\cdot-\tau)\Big{\rangle}_{S^{\varepsilon}_{T}},

where we used that θS(hS,2(t))=θ(hS,0){\theta}_{S}(h_{S,2}(t))={\theta}(h_{S,0}) for t<0t<0 and χ[0,T)(t)=0\chi_{[0,T)}(t)=0 for tTt\geq T. An analogous arguments hold also for τ<0\tau<0. Thus from equation (3.1) we have

θ(hS,2)θ(hS,2τ),hS,2hS,2τSTετ[|KS(hS,2)(hS,2+e3),ζτSTε|\displaystyle\big{\langle}{\theta}(h_{S,2})-{\theta}(h_{S,2}^{\tau}),h_{S,2}-h_{S,2}^{\tau}\rangle_{S^{\varepsilon}_{T}}\leq\tau\Big{[}\big{|}\big{\langle}K_{S}(h_{S,2})(\nabla h_{S,2}+{e}_{3}),\nabla\zeta_{\tau}\big{\rangle}_{S^{\varepsilon}_{T}}\big{|} (3.38)
+εkΓ(|hS,2,ζτΓP,Tε|+|hP,2,ζτΓP,Tε|)+|f(hS,2),ζτ(0,T)×ΓS,0ε|].\displaystyle+\varepsilon\,k_{\Gamma}\,\big{(}\big{|}\langle h_{S,2},\zeta_{\tau}\rangle_{\Gamma^{\varepsilon}_{P,T}}\big{|}+\big{|}\langle h_{P,2},\zeta_{\tau}\rangle_{\Gamma^{\varepsilon}_{P,T}}\big{|}\big{)}+\big{|}\big{\langle}f(h_{S,2}),\zeta_{\tau}\big{\rangle}_{(0,T)\times\Gamma^{\varepsilon}_{S,0}}\big{|}\Big{]}.

By the Lebesgue differentiation theorem, we have ζτζ\zeta_{\tau}\to\zeta strongly in L2(0,T;V(Sε))L^{2}(0,T;V(S^{\varepsilon})), and by the trace theorem, also in L2((0,T)×ΓPε)L^{2}((0,T)\times\Gamma^{\varepsilon}_{P}) and L2((0,T)×ΓS,0ε)L^{2}((0,T)\times\Gamma^{\varepsilon}_{S,0}). Hence the right hand side of (3.38) converges to 0 as τ0\tau\to 0 and from the Lipschitz continuity and monotonicity of θS{\theta}_{S} we obtain

θS(hS,2)θS(hS,2τ)L2((0,T)×Sε)2CθS(hS,2)θS(hS,2τ),hS,2hS,2τSTε0,\big{\|}{\theta}_{S}(h_{S,2})-{\theta}_{S}(h_{S,2}^{\tau})\big{\|}^{2}_{L^{2}((0,T)\times S^{\varepsilon})}\!\!\leq C\big{\langle}\theta_{S}(h_{S,2})-\theta_{S}(h_{S,2}^{\tau}),h_{S,2}-h_{S,2}^{\tau}\big{\rangle}_{S^{\varepsilon}_{T}}\!\!\to 0, (3.39)

as τ0\tau\to 0. An identical argument is employed to show

θP(hP,2)θP(hP,2τ)L2((0,T)×Pε)0 as τ0.\big{\|}{\theta}_{P}(h_{P,2})-{\theta}_{P}(h_{P,2}^{\tau})\big{\|}_{L^{2}((0,T)\times P^{\varepsilon})}\to 0\quad\text{ as }\tau\to 0. (3.40)

Combining (3.39) and (3.40) with the continuity of tκ\partial_{t}\kappa and taking τ0\tau\to 0 in (3.37) imply

0TJ=S,PJε(θJ(hJ,1)θJ(hJ,2))+𝑑xtκ(t)dt0.-\int_{0}^{T}\sum_{J=S,P}\int_{J^{\varepsilon}}\big{(}{\theta}_{J}(h_{J,1})-{\theta}_{J}(h_{J,2})\big{)}^{+}dx\,\partial_{t}\kappa(t)\,dt\leq 0. (3.41)

Applying integration by parts and using that κ\kappa is compactly supported yield

0Tκ(t)t[Sε(θS(hS,1)θS(hS,2))+𝑑x+Pε(θP(hP,1)θP(hP,2))+𝑑x]dt0,\int_{0}^{T}\!\!\!\kappa(t)\,\partial_{t}\Big{[}\int_{S^{\varepsilon}}\!\!\big{(}{\theta}_{S}(h_{S,1})-{\theta}_{S}(h_{S,2})\big{)}^{+}dx+\int_{P^{\varepsilon}}\!\!\big{(}{\theta}_{P}(h_{P,1})-{\theta}_{P}(h_{P,2})\big{)}^{+}dx\Big{]}dt\leq 0,

and, since this holds for all non-negative κC0(0,T)\kappa\in C_{0}^{\infty}(0,T), it follows that

J=S,PJε(θJ(hJ,1(t,x))θJ(hJ,2(t,x)))+𝑑xJ=S,PJε(θJ(hJ,1(0,x))θJ(hJ,2(0,x)))+𝑑x=0,\displaystyle\sum_{J=S,P}\int_{J^{\varepsilon}}\!\!\!\big{(}\theta_{J}(h_{J,1}(t,x))-\theta_{J}(h_{J,2}(t,x))\big{)}^{+}dx\leq\!\!\sum_{J=S,P}\int_{J^{\varepsilon}}\!\!\!\big{(}\theta_{J}(h_{J,1}(0,x))-\theta_{J}(h_{J,2}(0,x))\big{)}^{+}dx=0,

for a.e. t(0,T)t\in(0,T) and (θJ(hJ,1)θJ(hJ,2))+=0\big{(}{\theta}_{J}(h_{J,1})-{\theta}_{J}(h_{J,2})\big{)}^{+}=0 a.e. in (0,T)×Jε(0,T)\times J^{\varepsilon}, for J=S,PJ=S,P.

Using (3.24) with v0=hS,1(t1),w0=hP,1(t1)v^{0}=h_{S,1}(t_{1}),w^{0}=h_{P,1}(t_{1}), and κ(t1)\kappa(t_{1}), and the corresponding inequality for σδ\sigma_{\delta}^{-} and ηJ,δ\eta^{-}_{J,\delta} with v0=hS,2(t2),w0=hP,2(t2)v^{0}=h_{S,2}(t_{2}),w^{0}=h_{P,2}(t_{2}) and κ(t2)\kappa(t_{2}) and performing the same calculations as above yield (θJ(hJ,2)θJ(hJ,1))+=0\big{(}\theta_{J}(h_{J,2})-\theta_{J}(h_{J,1})\big{)}^{+}=0 a.e. in (0,T)×Jε(0,T)\times J^{\varepsilon}, for J=S,PJ=S,P. Thus, since θS{\theta}_{S} and θP{\theta}_{P} are strictly increasing, we obtain the uniqueness of hSh_{S} and hPh_{P}. ∎

Proposition 3.9.

Under Assumption 3.1 and additionally if KJ(z)=KJ,satK_{J}(z)=K_{J,\text{sat}} for z0z\geq 0, where J=R,B,PJ=R,B,P, and f(z)0f(z)\geq 0 for z0z\geq 0, solutions of (2.1), (2.3), (2.2), and (2.4) are non-positive.

Proof.

Using hS+h_{S}^{+} and hP+h_{P}^{+} as test functions in (3.1) and (3.2) and adding the resulting equations yield

J=S,P[tθJ(hJ),hJ+V(Jε),T+KJ(hJ)(hJ+e3),hJ+JTε]\displaystyle\sum_{J=S,P}\Big{[}\big{\langle}\partial_{t}{\theta}_{J}\big{(}h_{J}\big{)},h_{J}^{+}\big{\rangle}_{V(J^{\varepsilon})^{\prime},T}+\langle K_{J}(h_{J})\big{(}\nabla h_{J}+{e}_{3}\big{)},\nabla{h_{J}^{+}}\rangle_{J^{\varepsilon}_{T}}\Big{]} (3.42)
+εkΓhShP,hS+hP+ΓP,Tε+f(hS),hS+ΓS,0,Tε+𝒯pot,hP+ΓP,0,Tε=0.\displaystyle+\varepsilon k_{\Gamma}\langle h_{S}-h_{P},h_{S}^{+}-h_{P}^{+}\rangle_{\Gamma_{P,T}^{\varepsilon}}+\langle f(h_{S}),h_{S}^{+}\rangle_{\Gamma_{S,0,T}^{\varepsilon}}+\langle\mathcal{T}_{\text{pot}},h_{P}^{+}\rangle_{\Gamma_{P,0,T}^{\varepsilon}}=0.

Notice that hP+h_{P}^{+}\in V(Pε)V(P^{\varepsilon}), since hP=a0h_{P}=a\leq 0 on ΓP,L3\Gamma_{P,L_{3}}. We first show

ΨJ(hJ(s))ΨJ(hJ(sτ))[θJ(hJ(s))θJ(hJ(sτ))]hJ+(s),\Psi_{J}(h_{J}(s))-\Psi_{J}(h_{J}(s-\tau))\leq\big{[}\theta_{J}\big{(}h_{J}(s)\big{)}-\theta_{J}\big{(}h_{J}(s-\tau)\big{)}\big{]}h_{J}^{+}(s), (3.43)

for all s(0,T]s\in(0,T] and J=S,PJ=S,P, where ΨJ:[0,)\Psi_{J}:\mathbb{R}\to[0,\infty) is given by ΨJ(hJ)=0hJθJ(z)z+𝑑z\Psi_{J}(h_{J})=\int_{0}^{h_{J}}{\theta}^{\prime}_{J}(z)z^{+}dz. Integrating by parts and using θJ(0)=0{\theta}_{J}(0)=0 yield

ΨJ(hJ)={θJ(hJ)hJ0hJθJ(z)𝑑z if hJ>0,0 otherwise.\displaystyle\Psi_{J}(h_{J})=\begin{cases}{\theta}_{J}(h_{J})h_{J}-\int_{0}^{h_{J}}{\theta}_{J}(z)dz&\text{ if }h_{J}>0,\\ 0&\text{ otherwise}.\end{cases}

For τ>0\tau>0 and s(0,T]s\in(0,T], if hJ(s)>hJ(sτ)>0h_{J}(s)>h_{J}(s-\tau)>0, where hJ(sτ)=hJ,0h_{J}(s-\tau)=h_{J,0} for τ>s\tau>s, we have

ΨJ(hJ(s))ΨJ(hJ(sτ))θJ(hJ(s))hJ(s)θJ(hJ(sτ))hJ(sτ)\displaystyle\Psi_{J}(h_{J}(s))-\Psi_{J}(h_{J}(s-\tau))\leq\theta_{J}(h_{J}(s))h_{J}(s)-\theta_{J}(h_{J}(s-\tau))h_{J}(s-\tau) (3.44)
(hJ(s)hJ(sτ))θJ(hJ(sτ))(θJ(hJ(s))θJ(hJ(sτ)))hJ+(s).\displaystyle-(h_{J}(s)-h_{J}(s-\tau))\theta_{J}(h_{J}(s-\tau))\leq\big{(}\theta_{J}(h_{J}(s))-\theta_{J}(h_{J}(s-\tau))\big{)}h_{J}^{+}(s).

Following a similar line of argument, result (3.44) is also obtained for hJ(sτ)>hJ(s)>0h_{J}(s-\tau)>h_{J}(s)>0. If hJ(s)>0>hJ(sτ)h_{J}(s)>0>h_{J}(s-\tau), then since θJ(hJ(sτ))<0\theta_{J}(h_{J}(s-\tau))<0 it follows

ΨJ(hJ(s))ΨJ(hJ(sτ))θJ(hJ(s))hJ(s)(θJ(hJ(s))θJ(hJ(sτ)))hJ+(s).\Psi_{J}(h_{J}(s))-\Psi_{J}(h_{J}(s-\tau))\leq\theta_{J}\big{(}h_{J}(s)\big{)}h_{J}(s)\leq\big{(}{\theta}_{J}(h_{J}(s))-{\theta}_{J}(h_{J}(s-\tau))\big{)}h_{J}^{+}(s).

If hJ(sτ)>0>hJ(s)h_{J}(s-\tau)>0>h_{J}(s), then since ΨJ(hJ(s))=0\Psi_{J}(h_{J}(s))=0 and hJ+(s)=0h_{J}^{+}(s)=0 we have

ΨJ(hJ(s))ΨJ(hJ(sτ))0=[θJ(hJ(s))θJ(hJ(sτ))]hJ+(s).\Psi_{J}(h_{J}(s))-\Psi_{J}(h_{J}(s-\tau))\leq 0=\big{[}\theta_{J}(h_{J}(s))-\theta_{J}(h_{J}(s-\tau))\big{]}h_{J}^{+}(s).

Combining estimates above yields (3.43). Multiplying each side of (3.43) by 1/τ1/\tau, using that tθJ(hJ)L2(0,T;V(Jε))\partial_{t}{\theta}_{J}(h_{J})\in L^{2}(0,T;V(J^{\varepsilon})^{{}^{\prime}}) and hJ+L2(0,T;V(Jε))h_{J}^{+}\in L^{2}\big{(}0,T;V(J^{\varepsilon})\big{)} and the initial condition hJ,00h_{J,0}\leq 0 implies ΨJ(hJ,0)=0\Psi_{J}(h_{J,0})=0, and taking the limit as τ0\tau\to 0 yields

0TtθJ(hJ),hJ+V(Jε)𝑑tJεΨJ(hJ(T))𝑑x0, for J=S,P.\int_{0}^{T}\big{\langle}\partial_{t}{\theta}_{J}(h_{J}),h_{J}^{+}\big{\rangle}_{V(J^{\varepsilon})^{\prime}}dt\geq\int_{J^{\varepsilon}}\Psi_{J}(h_{J}(T))dx\geq 0,\qquad\text{ for }\;\;J=S,P. (3.45)

The definition of hS+,hS,hP+h_{S}^{+},h_{S}^{-},h_{P}^{+} and hPh_{P}^{-}, implies

hShS++hPhP+hPhS+hShP+=(hS+hP+)2hPhS+hShP+0.h_{S}h_{S}^{+}+h_{P}h_{P}^{+}-h_{P}h_{S}^{+}-h_{S}h_{P}^{+}=\big{(}h_{S}^{+}-h_{P}^{+}\big{)}^{2}-h_{P}^{-}h_{S}^{+}-h_{S}^{-}h_{P}^{+}\geq 0.

Combining the results above, together with 𝒯pot>0\mathcal{T}_{\text{pot}}>0 and f(hS)hS+0f\big{(}h_{S}\big{)}h_{S}^{+}\geq 0, it follows, from (3.42),

0T[SεKS(hS)[|hS+|2+x3hS+]𝑑x+PεKP(hP)[|hP+|2+x3hP+]𝑑x]𝑑t0.\int_{0}^{T}\!\!\Big{[}\int_{S^{\varepsilon}}\!\!\!K_{S}(h_{S})\big{[}|\nabla h_{S}^{+}|^{2}+\partial_{x_{3}}h_{S}^{+}\big{]}dx+\int_{P^{\varepsilon}}\!\!\!K_{P}(h_{P})\big{[}|\nabla h_{P}^{+}|^{2}+\partial_{x_{3}}h_{P}^{+}\big{]}dx\Big{]}dt\leq 0. (3.46)

Using KJ(hJ)=KJ,sat>0K_{J}(h_{J})=K_{J,\text{sat}}>0 for hJ>0h_{J}>0 and the boundary conditions on ΓJ,L3ε\Gamma^{\varepsilon}_{J,L_{3}}, gives

0TJεKJ(hJ)x3hJ+dxdt=KJ,sat0TJεx3hJ+dxdt=KJ,sat0TΓJ,0εhJ+𝑑x𝑑t,\int_{0}^{T}\!\!\!\int_{J^{\varepsilon}}\!\!\!K_{J}(h_{J})\partial_{x_{3}}h_{J}^{+}dxdt=K_{J,\text{sat}}\int_{0}^{T}\!\!\!\int_{J^{\varepsilon}}\!\!\!\partial_{x_{3}}h_{J}^{+}dxdt=K_{J,\text{sat}}\int_{0}^{T}\!\!\!\int_{\Gamma_{J,0}^{\varepsilon}}\!\!\!h_{J}^{+}dxdt, (3.47)

for J=S,PJ=S,P. Thus from (3.46) and (3.47) we obtain

0TJεKJ(hJ)|hJ+|2𝑑x𝑑t=0,KJ,sat0TΓJ,0εhJ+𝑑x𝑑t=0,J=S,P,\int_{0}^{T}\!\!\!\int_{J^{\varepsilon}}\!\!K_{J}\big{(}h_{J}\big{)}\big{\lvert}\nabla h_{J}^{+}\big{\rvert}^{2}dxdt=0,\qquad K_{J,\text{sat}}\int_{0}^{T}\!\!\!\int_{\Gamma_{J,0}^{\varepsilon}}\!\!\!h_{J}^{+}dxdt=0,\qquad J=S,P,

and that hSh_{S} and hPh_{P} are non-positive over (0,T)×Sε(0,T)\times S^{\varepsilon} and (0,T)×Pε(0,T)\times P^{\varepsilon} respectively. ∎

Remark 3.10.

Theorems 3.5 and 3.7 were proven for hS:(0,T)×Sεh_{S}:(0,T)\times S^{\varepsilon}\to\mathbb{R} and hP:(0,T)×Pεh_{P}:(0,T)\times P^{\varepsilon}\to\mathbb{R}. Physically realistic functions for water content θR\theta_{R}, θB\theta_{B}, and θP\theta_{P} and hydraulic conductivity KRK_{R}, KBK_{B}, and KPK_{P} are usually not defined for positive values of pressure head and, in the cases where they are, they often fail to satisfy Assumption 3.1, see [43, 25, 9]. In Appendix, we provide functions for θS\theta_{S}KSK_{S}θP\theta_{P} and KPK_{P}, which extend the expressions used in [43, 25, 9], to positive values of pressure head, in a way that the criteria of Assumption 3.1, Theorem 3.7 and Proposition 3.9 are satisfied. These extensions can therefore be assumed throughout the proofs of Theorems 3.5 and 3.7. Moreover, since Proposition 3.9 shows that hSh_{S} and hPh_{P} will remain non-positive, for any non-positive initial condition, the question of how realistic these extensions are for positive values of pressure head is not a concern.

4 Derivation of macroscopic model

To derive macroscopic equations from the microscopic model for the water transport in vegetated soil, we apply the two-scale convergence and periodic unfolding method, see e.g. [1, 11, 36, 37]. To pass to the limit as ε0\varepsilon\to 0 we first derive a priori estimates uniform in ε>0\varepsilon>0.

Lemma 4.1.

Under Assumption 3.1, solutions (hSε,hPε)(h_{S}^{\varepsilon},h_{P}^{\varepsilon}) of (2.1)–(2.4) satisfy the following estimates

hSεL2(0,T;V(Sε))2+hPεL2(PTε)2+x3hPεL2(PTε)2+εx^hPεL2(PTε)2\displaystyle\big{\|}h_{S}^{\varepsilon}\big{\|}^{2}_{L^{2}(0,T;V(S^{\varepsilon}))}+\big{\|}h_{P}^{\varepsilon}\big{\|}^{2}_{L^{2}(P^{\varepsilon}_{T})}+\big{\|}\partial_{x_{3}}h_{P}^{\varepsilon}\big{\|}^{2}_{L^{2}(P^{\varepsilon}_{T})}+\varepsilon\big{\|}\nabla_{\hat{x}}h_{P}^{\varepsilon}\big{\|}^{2}_{L^{2}(P^{\varepsilon}_{T})} C,\displaystyle\leq C, (4.1)
εhSεL2(ΓP,Tε)2+εhPεL2(ΓP,Tε)2+hSεL2((0,T)×ΓS,0ε)2+hPεL2((0,T)×ΓP,0ε)2\displaystyle\varepsilon\|h_{S}^{\varepsilon}\|^{2}_{L^{2}(\Gamma_{P,T}^{\varepsilon})}+\varepsilon\|h_{P}^{\varepsilon}\|^{2}_{L^{2}(\Gamma_{P,T}^{\varepsilon})}+\|h_{S}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times\Gamma_{S,0}^{\varepsilon})}+\|h_{P}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times\Gamma_{P,0}^{\varepsilon})} C,\displaystyle\leq C,
sup0tTΘSε(,hSε(t))L1(Sε)+sup0tTΘP(hPε(t))L1(Pε)\displaystyle\sup_{0\leq t\leq T}\|{\Theta}_{S}^{\varepsilon}(\cdot,{h}^{\varepsilon}_{S}(t))\|_{L^{1}(S^{\varepsilon})}+\sup_{0\leq t\leq T}\|{\Theta}_{P}(h^{\varepsilon}_{P}(t))\|_{L^{1}(P^{\varepsilon})} C,\displaystyle\leq C,

where x^hPε=(x1hPε,x2hPε)\nabla_{\hat{x}}h_{P}^{\varepsilon}=(\partial_{x_{1}}h_{P}^{\varepsilon},\partial_{x_{2}}h_{P}^{\varepsilon})^{\top}ΘSε{\Theta}_{S}^{\varepsilon} and ΘP{\Theta}_{P} as in (3.3), and C>0C>0 is independent of ε\varepsilon.

Proof.

Considering hSεh_{S}^{\varepsilon} and hPεah_{P}^{\varepsilon}-a as test functions in (3.1) and (3.2), respectively, adding the resulting equations, using Assumption 3.1 on ff and KJK_{J}, for J=B,R,PJ=B,R,P, and a0a\leq 0, yields

tθSε(x,hSε),hSεV(Sε),τ+tθP(hPε),hPεaV(Pε),τ\displaystyle\langle\partial_{t}{\theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon}),h_{S}^{\varepsilon}\rangle_{V(S^{\varepsilon})^{\prime},\tau}+\langle\partial_{t}{\theta}_{P}(h_{P}^{\varepsilon}),h_{P}^{\varepsilon}-a\rangle_{V(P^{\varepsilon})^{\prime},\tau} (4.2)
+KS,02hSεL2(Sτε)2+KP,02IεhPεL2(Pτε)2+εkΓhSεhPεL2(ΓP,τε)2\displaystyle+\frac{K_{S,0}}{2}\|\nabla h_{S}^{\varepsilon}\|^{2}_{L^{2}(S^{\varepsilon}_{\tau})}+\frac{K_{P,0}}{2}\|I_{\sqrt{\varepsilon}}\nabla h_{P}^{\varepsilon}\|^{2}_{L^{2}(P^{\varepsilon}_{\tau})}+\varepsilon k_{\Gamma}\|h_{S}^{\varepsilon}-h_{P}^{\varepsilon}\|^{2}_{L^{2}(\Gamma^{\varepsilon}_{P,\tau})}
δ[hSεL2(ΓS,0,τε)2+hPεL2(ΓP,0,τε)2]+14δ[KSε(x,hSε)L2(Sτε)2\displaystyle\leq\delta\big{[}\|h_{S}^{\varepsilon}\|^{2}_{L^{2}(\Gamma_{S,0,\tau}^{\varepsilon})}+\|h_{P}^{\varepsilon}\|^{2}_{L^{2}(\Gamma_{P,0,\tau}^{\varepsilon})}\big{]}+\frac{1}{4\delta}\big{[}\|K_{S}^{\varepsilon}(x,h_{S}^{\varepsilon})\|^{2}_{L^{2}(S^{\varepsilon}_{\tau})}
+IεKP(hPε)L2(Pτε)2+f(hSε)L2(ΓS,0,τε)2+𝒯potL2(ΓP,0,τε)2],\displaystyle\qquad+\|I_{\sqrt{\varepsilon}}K_{P}(h_{P}^{\varepsilon})\|^{2}_{L^{2}(P^{\varepsilon}_{\tau})}+\|f(h_{S}^{\varepsilon})\|^{2}_{L^{2}(\Gamma_{S,0,\tau}^{\varepsilon})}\!\!+\|\mathcal{T}_{\rm pot}\|^{2}_{L^{2}(\Gamma_{P,0,\tau}^{\varepsilon})}\big{]},

for τ(0,T]\tau\in(0,T] and 0<δmin{KS,0,KP,0}/20<\delta\leq\min\{K_{S,0},K_{P,0}\}/2. The definition of ΘSε\Theta_{S}^{\varepsilon} and ΘP\Theta_{P}, implies

0τtθSε(x,hSε),hSεV(Sε)𝑑t\displaystyle\int_{0}^{\tau}\!\!\big{\langle}\partial_{t}{\theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon}),h_{S}^{\varepsilon}\big{\rangle}_{V(S^{\varepsilon})^{\prime}}dt\geq SεΘSε(x,hSε(τ))𝑑xSεΘSε(x,hS,0)𝑑x,\displaystyle\int_{S^{\varepsilon}}\!\!{\Theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon}(\tau))dx-\int_{S^{\varepsilon}}\!\!{\Theta}_{S}^{\varepsilon}(x,h_{S,0})dx,\qquad (4.3)
0τtθP(hPε),hPεaV(Pε)𝑑t\displaystyle\int_{0}^{\tau}\!\!\big{\langle}\partial_{t}{\theta}_{P}(h_{P}^{\varepsilon}),h_{P}^{\varepsilon}-a\big{\rangle}_{V(P^{\varepsilon})^{\prime}}dt PεΘP(hPε(τ))𝑑xPεΘP(hP,0)𝑑x\displaystyle\geq\int_{P^{\varepsilon}}\!\!{\Theta}_{P}(h_{P}^{\varepsilon}(\tau))dx-\int_{P^{\varepsilon}}\!\!{\Theta}_{P}(h_{P,0})dx
aPεθP(hPε(τ))𝑑x+aPεθP(hP,0)𝑑x,\displaystyle-a\int_{P^{\varepsilon}}\!\!{\theta}_{P}(h_{P}^{\varepsilon}(\tau))dx+a\int_{P^{\varepsilon}}\!\!{\theta}_{P}(h_{P,0})dx,

for τ(0,T]\tau\in(0,T]. To estimate the third integral on the right hand-side of the second inequality in (4.3) we can use the continuity of θP\theta_{P} and estimate

|θP(hPε)|δΘP(hPε)+sup|σ|1/2|θP(σ)|,|{\theta}_{P}(h_{P}^{\varepsilon})|\leq\delta{\Theta}_{P}(h_{P}^{\varepsilon})+\sup\limits_{|\sigma|\leq 1/2}|{\theta}_{P}(\sigma)|,

which follows from the definition of ΘP{\Theta}_{P} or, as in our case, using the boundedness of θP\theta_{P}. The assumption on the microscopic structure of SεS^{\varepsilon} ensures existence of an extension h^SεL2(0,T;V(Ω))\hat{h}_{S}^{\varepsilon}\in L^{2}(0,T;V(\Omega)) of hSεh_{S}^{\varepsilon} from (0,T)×Sε(0,T)\times S^{\varepsilon} to (0,T)×Ω(0,T)\times\Omega such that

h^SεL2(ΩT)2ChSεL2(STε)2,h^SεL2(ΩT)2ChSεL2(STε)2,\|\hat{h}_{S}^{\varepsilon}\|^{2}_{L^{2}(\Omega_{T})}\leq C\|h_{S}^{\varepsilon}\|^{2}_{L^{2}(S^{\varepsilon}_{T})},\quad\|\nabla\hat{h}_{S}^{\varepsilon}\|^{2}_{L^{2}(\Omega_{T})}\leq C\|\nabla h_{S}^{\varepsilon}\|^{2}_{L^{2}(S^{\varepsilon}_{T})}, (4.4)

where C>C> is independent of ε\varepsilon, see e.g. [12, Theorem 2.10]. By the trace theorem and (4.4), we have

hSεL2((0,T)×ΓS,0ε)2h^SεL2((0,T)×Γ0)2C[h^SεL2(ΩT)2+h^SεL2(ΩT)2]\displaystyle\|h_{S}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times\Gamma_{S,0}^{\varepsilon})}\leq\|\hat{h}_{S}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times\Gamma_{0})}\leq C\big{[}\|\hat{h}_{S}^{\varepsilon}\|^{2}_{L^{2}(\Omega_{T})}+\|\nabla\hat{h}_{S}^{\varepsilon}\|^{2}_{L^{2}(\Omega_{T})}\big{]} (4.5)
C[hSεL2(STε)2+hSεL2(STε)2].\displaystyle\leq C\big{[}\|h_{S}^{\varepsilon}\|^{2}_{L^{2}(S^{\varepsilon}_{T})}+\|\nabla h_{S}^{\varepsilon}\|^{2}_{L^{2}(S^{\varepsilon}_{T})}\big{]}.

Applying the trace theorem over the unit cell P=YP×(L3,0)P=Y_{P}\times(-L_{3},0) and the ε\varepsilon-scaling in (y1,y2)(y_{1},y_{2})-variables, together with the definition of the domain PεP^{\varepsilon}, yields

hPεL2((0,T)×ΓP,0ε)2C(hPεL2((0,T)×Pε)2+IεhPεL2((0,T)×Pε)2).\|h_{P}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times\Gamma^{\varepsilon}_{P,0})}\leq C\big{(}\|h_{P}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times P^{\varepsilon})}+\|I_{\varepsilon}\nabla h_{P}^{\varepsilon}\|^{2}_{L^{2}((0,T)\times P^{\varepsilon})}\big{)}. (4.6)

Thus, using (4.3), (4.5), and (4.6) in inequality (4.2) implies

ΘSε(,hSε(τ))L1(Sε)\displaystyle\big{\|}{\Theta}_{S}^{\varepsilon}\big{(}\cdot,h_{S}^{\varepsilon}(\tau)\big{)}\big{\|}_{L^{1}(S^{\varepsilon})} +ΘP(hPε(τ))L1(Pε)+hSεL2(Sτε)2\displaystyle+\big{\|}{\Theta}_{P}\big{(}h_{P}^{\varepsilon}(\tau)\big{)}\big{\|}_{L^{1}(P^{\varepsilon})}+\|\nabla h_{S}^{\varepsilon}\|^{2}_{L^{2}(S^{\varepsilon}_{\tau})} (4.7)
+IεhPεL2(Pτε)2δ[hPεL2(Pτε)2+hSεL2(Sτε)2]+Cδ,\displaystyle+\|I_{\sqrt{\varepsilon}}\nabla h_{P}^{\varepsilon}\|^{2}_{L^{2}(P^{\varepsilon}_{\tau})}\leq\delta\Big{[}\|h_{P}^{\varepsilon}\|_{L^{2}(P^{\varepsilon}_{\tau})}^{2}+\|h_{S}^{\varepsilon}\|_{L^{2}(S^{\varepsilon}_{\tau})}^{2}\Big{]}+C_{\delta},

for τ(0,T]\tau\in(0,T] and 0<δmin{KS,0,KP,0}/(4C)0<\delta\leq\min\{K_{S,0},K_{P,0}\}/(4C). Applying the generalised Poincaré inequality over PP and using the boundary condition on ΓP,L3\Gamma_{P,L_{3}} and the standard scaling argument, for the first term on the right hand side of (4.7) we have

hPεL2(Pτε)2C[IεhPεL2(Pτε)2+T|a|2|ΓP,L3ε|],\displaystyle\|h_{P}^{\varepsilon}\|^{2}_{L^{2}(P^{\varepsilon}_{\tau})}\leq C\big{[}\|I_{\varepsilon}\nabla h_{P}^{\varepsilon}\|^{2}_{L^{2}(P^{\varepsilon}_{\tau})}+T|a|^{2}|\Gamma^{\varepsilon}_{P,L_{3}}|\big{]}, (4.8)

where C>0C>0 is independent of ε\varepsilon. Similarly for the second term on the right hand side of (4.7), the generalised Poincaré inequality, the extension properties, and hSε=0h_{S}^{\varepsilon}=0 at ΓS,L3ε\Gamma^{\varepsilon}_{S,L_{3}}, imply

hSεL2(Sτε)h^SεL2(Ωτ)C1h^SεL2(Ωτ)C2hSεL2(Sτε),\displaystyle\|h_{S}^{\varepsilon}\|_{L^{2}(S^{\varepsilon}_{\tau})}\leq\|\hat{h}_{S}^{\varepsilon}\|_{L^{2}(\Omega_{\tau})}\leq C_{1}\|\nabla\hat{h}_{S}^{\varepsilon}\|_{L^{2}(\Omega_{\tau})}\leq C_{2}\|\nabla h_{S}^{\varepsilon}\|_{L^{2}(S^{\varepsilon}_{\tau})}, (4.9)

where C1,C2>0C_{1},C_{2}>0 are independent of ε\varepsilon. Using now (4.8) and (4.9) in (4.7), and considering an appropriate δ\delta, yields the first and the last estimates stated in the lemma.

The trace theorem over YY, together with the standard scaling argument, implies

εhJεL2(ΓP,Tε)2C(hJεL2(JTε)2+ε2x^hJεL2(JTε)2), for J=S,P,\varepsilon\|h_{J}^{\varepsilon}\|^{2}_{L^{2}(\Gamma_{P,T}^{\varepsilon})}\leq C\big{(}\|h_{J}^{\varepsilon}\|^{2}_{L^{2}(J_{T}^{\varepsilon})}+\varepsilon^{2}\|\nabla_{\hat{x}}h_{J}^{\varepsilon}\|^{2}_{L^{2}(J_{T}^{\varepsilon})}\big{)},\quad\text{ for }J=S,P, (4.10)

with constant C>0C>0 independent of ε\varepsilon. This, together with the first estimate in (4.1), yields the estimates for the L2((0,T)×ΓPε)L^{2}((0,T)\times\Gamma_{P}^{\varepsilon})-norm. Estimates (4.5) and (4.6) ensure the uniform in ε\varepsilon boundedness of hSεh_{S}^{\varepsilon} in L2((0,T)×ΓS,0ε)L^{2}((0,T)\times\Gamma_{S,0}^{\varepsilon}) and hPεh_{P}^{\varepsilon} in L2((0,T)×ΓP,0ε)L^{2}((0,T)\times\Gamma_{P,0}^{\varepsilon}) respectively. ∎

Estimates in (4.1), together with the properties of the extension (4.4) and of the two-scale convergence, see e.g. [1, 37, 36], imply the following convergence results

Lemma 4.2.

There exist hSL2(0,T;V(Ω))h_{S}\in L^{2}(0,T;V(\Omega)) and hS,1L2((0,T)×Ω;Hper1(Y))h_{S,1}\in L^{2}((0,T)\times\Omega;H^{1}_{\rm per}(Y)) such that, upto a subsequence,

h^SεhS\displaystyle\hat{h}_{S}^{\varepsilon}\rightharpoonup h_{S} weakly in L2(0,T;V(Ω)),\displaystyle\text{ weakly in }L^{2}(0,T;V(\Omega)), (4.11)
h^SεhS,h^SεhS+y,0hS,1\displaystyle\hat{h}_{S}^{\varepsilon}\rightharpoonup h_{S},\;\;\nabla\hat{h}_{S}^{\varepsilon}\rightharpoonup\nabla h_{S}+\nabla_{y,0}h_{S,1} two-scale,\displaystyle\text{ two-scale},
hSεhS\displaystyle h_{S}^{\varepsilon}\rightharpoonup h_{S} two-scale on (0,T)×ΓPε,\displaystyle\text{ two-scale on }(0,T)\times\Gamma_{P}^{\varepsilon},

as ε0\varepsilon\to 0, where y,0u=(y1u,y2u,0)T\nabla_{y,0}u=(\partial_{y_{1}}u,\partial_{y_{2}}u,0)^{T}.

The next lemma provides the convergence result for the sequence {hPε}\{h_{P}^{\varepsilon}\}.

Lemma 4.3.

For an extension hPε~\widetilde{h_{P}^{\varepsilon}} of hPεh_{P}^{\varepsilon} by zero from (0,T)×Pε(0,T)\times P^{\varepsilon} into (0,T)×Ω(0,T)\times\Omega, there exists hPL2((0,T)×Ω)h_{P}\in L^{2}((0,T)\times\Omega), with x3hPL2((0,T)×Ω)\partial_{x_{3}}h_{P}\in L^{2}((0,T)\times\Omega), such that

hPε~hPχYP\displaystyle\widetilde{h_{P}^{\varepsilon}}\rightharpoonup h_{P}\,\chi_{Y_{P}} two-scale,\displaystyle\text{ two-scale}, (4.12)
hPεhP\displaystyle h_{P}^{\varepsilon}\rightharpoonup h_{P} two-scale on (0,T)×ΓPε,\displaystyle\text{ two-scale on }\;(0,T)\times\Gamma_{P}^{\varepsilon},
hPε~|YP||Y|hP,x3hPε~|YP||Y|x3hP\displaystyle\widetilde{h_{P}^{\varepsilon}}\rightharpoonup\frac{|Y_{P}|}{\lvert Y\rvert}h_{P},\quad\widetilde{\partial_{x_{3}}h_{P}^{\varepsilon}}\rightharpoonup\frac{|Y_{P}|}{\lvert Y\rvert}\partial_{x_{3}}h_{P} weakly in L2((0,T)×Ω),\displaystyle\text{ weakly in }\;L^{2}((0,T)\times\Omega),
εx^hPε~0\displaystyle\varepsilon\widetilde{\nabla_{\hat{x}}h_{P}^{\varepsilon}}\rightharpoonup 0 weakly in L2((0,T)×Ω)2.\displaystyle\text{ weakly in }\;L^{2}((0,T)\times\Omega)^{2}.
Proof.

From Lemma 4.1 we have that the sequence hPε~\widetilde{h_{P}^{\varepsilon}} is bounded in L2((0,T)×Ω)L^{2}((0,T)\times\Omega) and there exists a subsequence, denoted again by  hPε~\widetilde{h_{P}^{\varepsilon}}, which converges two-scale to hPL2((0,T)×Ω×Y)h_{P}\in L^{2}((0,T)\times\Omega\times Y) with hP(t,x,y)=0h_{P}(t,x,y)=0 for yYY¯Py\in Y\setminus\overline{Y}_{P}, see e.g. [1]. From Lemma 4.1 we also have existence of a subsequence of ε1/2x^hPε~\varepsilon^{1/2}\widetilde{\nabla_{\hat{x}}h_{P}^{\varepsilon}} that converges two-scale and weakly in L2((0,T)×Ω)L^{2}((0,T)\times\Omega), and

ε0TPεx^hPεψ(t,x,x^ε)𝑑x𝑑t=ε0TΩx^hPε~ψ(t,x,x^ε)𝑑x𝑑t0as ε0,\displaystyle\varepsilon\int_{0}^{T}\!\!\!\!\int_{P^{\varepsilon}}\nabla_{\hat{x}}h_{P}^{\varepsilon}\,\psi\Big{(}t,x,\frac{\hat{x}}{\varepsilon}\Big{)}dxdt=\varepsilon\int_{0}^{T}\!\!\!\!\int_{\Omega}\widetilde{\nabla_{\hat{x}}h_{P}^{\varepsilon}}\,\psi\Big{(}t,x,\frac{\hat{x}}{\varepsilon}\Big{)}dxdt\to 0~{}\text{as }\varepsilon\to 0, (4.13)

for ψL2(0,T;C0(Ω×YP))2\psi\in L^{2}(0,T;C_{0}^{\infty}(\Omega\times Y_{P}))^{2}. This, together with the two-scale convergence of hPε~\widetilde{h_{P}^{\varepsilon}} and with

ε0TPεx^hPεψ(t,x,xε)𝑑x𝑑t=0TPεhPε[εdivx^ψ+divyψ]𝑑x𝑑t\displaystyle\varepsilon\int_{0}^{T}\!\!\!\!\int_{P^{\varepsilon}}\nabla_{\hat{x}}h_{P}^{\varepsilon}\,\psi\Big{(}t,x,\frac{x}{\varepsilon}\Big{)}dxdt=-\int_{0}^{T}\!\!\!\!\int_{P^{\varepsilon}}h_{P}^{\varepsilon}\,\big{[}\varepsilon{\rm div}_{\hat{x}}\psi+{\rm div}_{y}\psi\big{]}dxdt (4.14)
=0TΩhPε~[εdivx^ψ+divyψ]𝑑x𝑑t1|Y|0TΩYPhPdivyψ𝑑y𝑑x𝑑t,\displaystyle=-\int_{0}^{T}\!\!\!\!\int_{\Omega}\widetilde{h_{P}^{\varepsilon}}\big{[}\varepsilon{\rm div}_{\hat{x}}\psi+{\rm div}_{y}\psi\big{]}dxdt\to-\frac{1}{\lvert Y\rvert}\int_{0}^{T}\!\!\!\!\int_{\Omega}\int_{Y_{P}}h_{P}\,{\rm div}_{y}\psi\,dydxdt,

as ε0\varepsilon\to 0, implies hP(t,x,y)=hP(t,x)h_{P}(t,x,y)=h_{P}(t,x) for (t,x,y)(0,T)×Ω×YP(t,x,y)\in(0,T)\times\Omega\times Y_{P}. Choosing ψ(t,x,y)=ψ(t,x)\psi(t,x,y)=\psi(t,x), with ψC0((0,T)×Ω)\psi\in C_{0}((0,T)\times\Omega), in the definition of the two-scale convergence of hPε~\widetilde{h_{P}^{\varepsilon}} gives

limε00TΩhPε~(t,x)ψ(t,x)𝑑x𝑑t=0TΩ|YP||Y|hP(t,x)ψ(t,x)𝑑x𝑑t\displaystyle\lim_{\varepsilon\to 0}\int_{0}^{T}\!\!\!\!\int_{\Omega}\widetilde{h_{P}^{\varepsilon}}(t,x)\psi(t,x)dxdt=\int_{0}^{T}\!\!\!\!\int_{\Omega}\frac{|Y_{P}|}{\lvert Y\rvert}h_{P}(t,x)\psi(t,x)dxdt

and hence the third convergence in (4.12). The microscopic structure of PεP^{\varepsilon} implies x3hPε~=x3hPε~\widetilde{\partial_{x_{3}}h_{P}^{\varepsilon}}=\partial_{x_{3}}\widetilde{h_{P}^{\varepsilon}}. Then the estimate for x3hPε\partial_{x_{3}}h_{P}^{\varepsilon}, see Lemma 4.1, ensures existence of a subsequence of x3hPε~\widetilde{\partial_{x_{3}}h_{P}^{\varepsilon}} converging weakly in L2((0,T)×Ω)L^{2}((0,T)\times\Omega) and, using the weak convergence of hPε~\widetilde{h^{\varepsilon}_{P}}, we obtain the fourth convergence in (4.12). A priori estimates in (4.1) and properties of the two-scale convergence on oscillating boundaries, see e.g. [2, 36] ensure the second convergence result in (4.12). The uniform in ε\varepsilon estimate for ε12x^hPε~\varepsilon^{\frac{1}{2}}\widetilde{\nabla_{\hat{x}}h_{P}^{\varepsilon}}, see (4.1), ensures weak convergence, up to a subsequence, of ε12x^hPε~\varepsilon^{\frac{1}{2}}\widetilde{\nabla_{\hat{x}}h_{P}^{\varepsilon}} in L2((0,T)×Ω)L^{2}((0,T)\times\Omega), and hence the last convergence result in (4.12). ∎

To pass to the limit in the nonlinear terms in (3.1) and (3.2), we show the strong two-scale convergence of θSε(,hSε){\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}) and θP(hPε){\theta}_{P}(h_{P}^{\varepsilon}), by using the Aubin-Lions-Simon compactness lemma [41] and showing the equicontinity of the corresponding sequences.

Lemma 4.4.

Under Assumption 3.1, for solutions of (2.1)-(2.4) we have

0TλθSε(,hSε(t+λ))θSε(,hSε(t)),hSε(t+λ)hSε(t)Sε𝑑tCλ,\displaystyle\int_{0}^{T-\lambda}\big{\langle}\theta_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}(t+\lambda))-\theta^{\varepsilon}_{S}(\cdot,h_{S}^{\varepsilon}(t)),h_{S}^{\varepsilon}(t+\lambda)-h_{S}^{\varepsilon}(t)\big{\rangle}_{S^{\varepsilon}}dt\leq C\lambda, (4.15)
0TλθP(hPε(t+λ))θP(hPε(t)),hPε(t+λ)hPε(t)Pε𝑑tCλ,\displaystyle\int_{0}^{T-\lambda}\big{\langle}{\theta}_{P}(h_{P}^{\varepsilon}(t+\lambda))-{\theta}_{P}(h_{P}^{\varepsilon}(t)),h_{P}^{\varepsilon}(t+\lambda)-h_{P}^{\varepsilon}(t)\big{\rangle}_{P^{\varepsilon}}dt\leq C\lambda,

where 0<λ<T0<\lambda<T and C>0C>0 is independent of ε\varepsilon.

Proof.

Considering as a test function in (3.1) the following function

ψλε(hSε):=1λtλt(hSε(s+λ)hSε(s))χ(0,Tλ](s)𝑑s,\psi^{\varepsilon}_{\lambda}(h_{S}^{\varepsilon}):=\frac{1}{\lambda}\int_{t-\lambda}^{t}\!\!\big{(}h_{S}^{\varepsilon}(s+\lambda)-h_{S}^{\varepsilon}(s)\big{)}\chi_{(0,T-\lambda]}(s)ds,

using an integration by parts, and applying the Cauchy-Schwarz inequality yield

θSε(,hSε(+λ))θSε(,hSε),hSε(+λ)hSεSTλε=λtθSε(,hSε),ψλεV(Sε),Tλ\displaystyle\big{\langle}{\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}(\cdot+\lambda))-{\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}),h_{S}^{\varepsilon}(\cdot+\lambda)-h_{S}^{\varepsilon}\big{\rangle}_{S^{\varepsilon}_{T-\lambda}}=\lambda\big{\langle}\partial_{t}{\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}),\psi^{\varepsilon}_{\lambda}\big{\rangle}_{V^{\prime}(S^{\varepsilon}),T-\lambda}
λ[KSε(,hSε)hSεL2(STλε)+KSε(,hSε)L2(STλε)]ψλεL2(STλε)\displaystyle\leq\lambda\big{[}\|K_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon})\nabla h_{S}^{\varepsilon}\|_{L^{2}(S^{\varepsilon}_{T-\lambda})}+\|K_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon})\|_{L^{2}(S^{\varepsilon}_{T-\lambda})}\big{]}\|\nabla\psi^{\varepsilon}_{\lambda}\|_{L^{2}(S^{\varepsilon}_{T-\lambda})}
+λεkΓ[hSεL2(ΓP,Tλε)+hPεL2(ΓP,Tλε)]ψλεL2(ΓP,Tλε)\displaystyle\qquad+\lambda\varepsilon k_{\Gamma}\big{[}\|h_{S}^{\varepsilon}\|_{L^{2}(\Gamma_{P,T-\lambda}^{\varepsilon})}+\|h_{P}^{\varepsilon}\|_{L^{2}(\Gamma_{P,T-\lambda}^{\varepsilon})}\big{]}\|\psi^{\varepsilon}_{\lambda}\|_{L^{2}(\Gamma_{P,T-\lambda}^{\varepsilon})}
+λf(hSε)L2((0,Tλ)×ΓS,0ε)ψλεL2((0,Tλ)×ΓS,0ε).\displaystyle\qquad+\lambda\|f(h_{S}^{\varepsilon})\|_{L^{2}((0,T-\lambda)\times\Gamma_{S,0}^{\varepsilon})}\|\psi^{\varepsilon}_{\lambda}\|_{L^{2}((0,T-\lambda)\times\Gamma_{S,0}^{\varepsilon})}.

A priori estimates in Lemma 4.1 and the uniform boundedness of KSεK_{S}^{\varepsilon} and ff imply the first estimate in (4.15). Analogous calculations, with ψλε(hPε)\psi_{\lambda}^{\varepsilon}(h_{P}^{\varepsilon}) as a test function in (3.2), yield the second estimate in (4.15). ∎

To show the strong two-scale convergence of θSε(x,hSε){\theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon}) we use the unfolding operator 𝒯ε:L2((0,T)×Sε)L2((0,T)×Ω×S^)\mathcal{T}^{\varepsilon}:L^{2}((0,T)\times S^{\varepsilon})\to L^{2}((0,T)\times\Omega\times\hat{S}) given by

𝒯ε(hSε)(t,x,y)=hSε(t,ε[x^ε]Y+εy,x3) for t(0,T),xΩ,yS^,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})(t,x,y)=h_{S}^{\varepsilon}\Big{(}t,\varepsilon\Big{[}\frac{\hat{x}}{\varepsilon}\Big{]}_{Y}+\varepsilon y,x_{3}\Big{)}\qquad\text{ for }\;t\in(0,T),\;x\in\Omega,\;y\in\hat{S},

where S^=YY¯P\hat{S}=Y\setminus\overline{Y}_{P}, x^=(x1,x2)\hat{x}=(x_{1},x_{2})y=(y1,y2)y=(y_{1},y_{2}), and [x^ε]Y[\frac{\hat{x}}{\varepsilon}]_{Y} is the unique integer combination, see e.g. [11]. Similar definition we have for 𝒯ε:L2((0,T)×Pε)L2((0,T)×Ω×YP)\mathcal{T}^{\varepsilon}:L^{2}((0,T)\times P^{\varepsilon})\to L^{2}((0,T)\times\Omega\times Y_{P}).

Lemma 4.5.

Under Assumption 3.1, for solutions of (2.1), (2.3) we have, up to a subsequence,

𝒯ε(hSε)hS\displaystyle\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\to h_{S} a.e. in (0,T)×Ω×S^,\displaystyle\text{ a.e.~{}in }(0,T)\times\Omega\times\hat{S},
θSε(x,hSε)θS(y,hS)=χR^(y)θR(hS)+χB^(y)θB(hS)\displaystyle{\theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon})\to{\theta}_{S}(y,h_{S})=\chi_{\hat{R}}(y){\theta}_{R}(h_{S})+\chi_{\hat{B}}(y){\theta}_{B}(h_{S})\quad strongly two-scale.\displaystyle\text{strongly two-scale}.
Proof.

First we show the strong convergence of θR(𝒯ε(hSε)){\theta}_{R}(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})) in L2((0,T)×Ω×R^)L^{2}((0,T)\times\Omega\times\hat{R}) and of θB(𝒯ε(hSε)){\theta}_{B}(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})) in L2((0,T)×Ω×B^)L^{2}((0,T)\times\Omega\times\hat{B}). The Lipschitz continuity of θJ\theta_{J}, with J=R,BJ=R,B, properties of the unfolding operator, see e.g. [11], and Lemma 4.4 imply

t1t2θJ(𝒯ε(hSε)(t,+r,))dtt1t2θJ(𝒯ε(hSε))dtL2(Ω×J^)2\displaystyle\Big{\|}\int_{t_{1}}^{t_{2}}\!\!{\theta}_{J}\big{(}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})(t,\cdot+r,\cdot)\big{)}dt-\int_{t_{1}}^{t_{2}}\!\!{\theta}_{J}\big{(}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\big{)}dt\Big{\|}_{L^{2}(\Omega\times\hat{J})}^{2}
C1θJ(hSε(,+r))θJ(hSε)L2((t1,t2)×Jε)2C2|r|2hSεL2((t1,t2)×Jε)2C|r|2,\displaystyle\leq C_{1}\big{\|}{\theta}_{J}\big{(}h_{S}^{\varepsilon}(\cdot,\cdot+r)\big{)}-{\theta}_{J}\big{(}h_{S}^{\varepsilon}\big{)}\big{\|}^{2}_{L^{2}((t_{1},t_{2})\times J^{\varepsilon})}\leq C_{2}|r|^{2}\|\nabla{h_{S}^{\varepsilon}}\|_{L^{2}((t_{1},t_{2})\times J^{\varepsilon})}^{2}\leq C|r|^{2},
t1t2θJ(𝒯ε(hSε)(t,,+r1))dtt1t2θJ(𝒯ε(hSε))dtL2(Ω×J^)2\displaystyle\Big{\|}\int_{t_{1}}^{t_{2}}\!\!{\theta}_{J}\big{(}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})(t,\cdot,\cdot+r_{1})\big{)}dt-\int_{t_{1}}^{t_{2}}\!\!{\theta}_{J}\big{(}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\big{)}dt\Big{\|}_{L^{2}(\Omega\times\hat{J})}^{2}
C1|r1|2y𝒯ε(θJ(hSε))L2((t1,t2)×Ω×J^)2C2|r1|2ε2hSεL2((t1,t2)×Jε)2C|r1|2,\displaystyle\leq C_{1}|r_{1}|^{2}\|\nabla_{y}\mathcal{T}^{\varepsilon}\big{(}{\theta}_{J}(h_{S}^{\varepsilon})\big{)}\big{\|}_{L^{2}((t_{1},t_{2})\times\Omega\times\hat{J})}^{2}\leq C_{2}|r_{1}|^{2}\varepsilon^{2}\|\nabla{h_{S}^{\varepsilon}}\|_{L^{2}((t_{1},t_{2})\times J^{\varepsilon})}^{2}\leq C|r_{1}|^{2},
θJ(𝒯ε(hSε)(+λ,,))θJ(𝒯ε(hSε))L2(ΩTλ×J^)2\displaystyle\big{\|}{\theta}_{J}(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})(\cdot+\lambda,\cdot,\cdot))-{\theta}_{J}(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}))\big{\|}^{2}_{L^{2}(\Omega_{T-\lambda}\times\hat{J})}
C10TλθSε(,hSε(t+λ))θSε(,hSε(t)),hSε(t+λ)hSε(t)Sε𝑑tCλ,\displaystyle\quad\leq C_{1}\int_{0}^{T-\lambda}\!\!\!\big{\langle}{\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}(t+\lambda))-{\theta}_{S}^{\varepsilon}(\cdot,h_{S}^{\varepsilon}(t)),h_{S}^{\varepsilon}(t+\lambda)-h_{S}^{\varepsilon}(t)\big{\rangle}_{S^{\varepsilon}}dt\leq C\lambda,

for r3r\in\mathbb{R}^{3}, r12r_{1}\in\mathbb{R}^{2}, and λ>0\lambda>0, where J=R,BJ=R,B and C1,C2,C>0C_{1},C_{2},C>0 are independent of ε\varepsilon. Combining the estimates from above and using the compactness result in [41], yield the existence of zRL2((0,T)×Ω×R^)z_{R}\in L^{2}((0,T)\times\Omega\times\hat{R}) and zBL2((0,T)×Ω×B^)z_{B}\in L^{2}((0,T)\times\Omega\times\hat{B}) such that, up to a subsequence,

θJ(𝒯ε(hSε))zJstrongly in L2((0,T)×Ω×J^), for J=R,B.{\theta}_{J}\big{(}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\big{)}\to z_{J}\quad\text{strongly in }\;\;L^{2}((0,T)\times\Omega\times\hat{J}),\quad\text{ for }\;\;J=R,B. (4.16)

Since θB{\theta}_{B} and θR{\theta}_{R} are strictly increasing and continuous, we have

𝒯ε(hSε)=θJ1(θJ(𝒯ε(hSε)))θJ1(zJ)a.e. in (0,T)×Ω×J^,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})={\theta}_{J}^{-1}\big{(}{\theta}_{J}\big{(}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\big{)}\big{)}\to{\theta}_{J}^{-1}(z_{J})\quad\text{a.e.~{}in }(0,T)\times\Omega\times\hat{J}, (4.17)

for J=R,BJ=R,B. The two-scale convergence of hSεh_{S}^{\varepsilon} implies 𝒯ε(hSε)hS\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\rightharpoonup h_{S} in L2((0,T)×Ω×S^)L^{2}((0,T)\times\Omega\times\hat{S}), see e.g. [11]. Using (4.16) and (4.17), together with 𝒯ε(θJ(hSε))=θJ(𝒯ε(hSε))\mathcal{T}^{\varepsilon}({\theta}_{J}(h_{S}^{\varepsilon}))={\theta}_{J}(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})), ensures 𝒯ε(θJ(hSε))θJ(hS)\mathcal{T}^{\varepsilon}\big{(}{\theta}_{J}(h_{S}^{\varepsilon})\big{)}\to{\theta}_{J}(h_{S}) strongly in L2((0,T)×Ω×J^)L^{2}\big{(}(0,T)\times\Omega\times\hat{J}\big{)} and θJ(hSε)θJ(hS){\theta}_{J}(h_{S}^{\varepsilon})\to{\theta}_{J}(h_{S}) strongly two-scale, for J=R,BJ=R,B. Then θSε(x,hSε)=χR^(x^/ε)θR(hSε)+χB^(x^/ε)θB(hSε){\theta}_{S}^{\varepsilon}(x,h_{S}^{\varepsilon})=\chi_{\hat{R}}(\hat{x}/\varepsilon){\theta}_{R}(h_{S}^{\varepsilon})+\chi_{\hat{B}}(\hat{x}/\varepsilon){\theta}_{B}(h_{S}^{\varepsilon}) converges strongly two-scale to θS(,hS)L2((0,T)×Ω×S^){\theta}_{S}(\cdot,h_{S})\in L^{2}((0,T)\times\Omega\times\hat{S}). ∎

Lemma 4.6.

Under Assumption 3.1, for solutions of (2.2), (2.4) we have, up to a subsequence,

𝒯ε(hPε)hP\displaystyle\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})\to h_{P} a.e. in (0,T)×Ω×YP,\displaystyle\text{ a.e.~{}in }\,(0,T)\times\Omega\times Y_{P}, (4.18)
θP(hPε)θP(hP)χYP\displaystyle{\theta}_{P}(h_{P}^{\varepsilon})\to{\theta}_{P}(h_{P})\chi_{Y_{P}} strongly two-scale.\displaystyle\text{ strongly two-scale}.
Proof.

To prove strong convergence of θP(hPε){\theta}_{P}(h_{P}^{\varepsilon}) we first show the equicontinuity of 𝒯ε(θP(hPε))\mathcal{T}^{\varepsilon}(\theta_{P}(h^{\varepsilon}_{P})) by using similar arguments as in the proof of the uniqueness result in Theorem 3.7. Consider Ω^=Ω{x3=0}\hat{\Omega}=\Omega\cap\{x_{3}=0\} and Ως={xΩ:dist(x^,Ω^)ς}\Omega_{\varsigma}=\{x\in\Omega:{\rm dist}(\hat{x},\partial\hat{\Omega})\geq\varsigma\}, with ς>0\varsigma>0 and x^=(x1,x2)\hat{x}=(x_{1},x_{2}), and define

P^ςε=ξΞςεε(YP+ξ),Γ^Pςε=ξΞςεε(ΓP+ξ),\displaystyle\hat{P}^{\varepsilon}_{\varsigma}=\bigcup\limits_{\xi\in\Xi_{\varsigma}^{\varepsilon}}\varepsilon(Y_{P}+\xi),\quad\hat{\Gamma}^{\varepsilon}_{P_{\varsigma}}=\bigcup\limits_{\xi\in\Xi_{\varsigma}^{\varepsilon}}\varepsilon(\Gamma_{P}+\xi),

and Pςε=P^ςε×(L3,0)P^{\varepsilon}_{\varsigma}=\hat{P}^{\varepsilon}_{\varsigma}\times(-L_{3},0), Γςε=Γ^Pςε×(L3,0)\Gamma^{\varepsilon}_{\varsigma}=\hat{\Gamma}^{\varepsilon}_{P_{\varsigma}}\times(-L_{3},0), where Ξςε={ξΞε:ε(Y+ξ)×(L3,0)Ως}\Xi^{\varepsilon}_{\varsigma}=\{\xi\in\Xi^{\varepsilon}:\varepsilon(Y+\xi)\times(-L_{3},0)\subset\Omega_{\varsigma}\}. For l2l\in\mathbb{Z}^{2}, such that |lε|<ς|l\varepsilon|<\varsigma, and ψl(t,x)=ψ(x^lε,x3,t)\psi^{-l}(t,x)=\psi(\hat{x}-l\varepsilon,x_{3},t), where ψC01(0,T;V(Pςε))\psi\in C^{1}_{0}(0,T;V(P_{\varsigma}^{\varepsilon})), we have ψlC01(0,T;V(Pς,lε))\psi^{-l}\in C^{1}_{0}(0,T;V(P_{\varsigma,l}^{\varepsilon})), with Pς,lε={x+(εl1,εl2,0)T:xPςε}P_{\varsigma,l}^{\varepsilon}=\{x+(\varepsilon l_{1},\varepsilon l_{2},0)^{T}:x\in P_{\varsigma}^{\varepsilon}\}. Using ψl\psi^{-l} in the weak formulation of (2.2) and (2.4) over Pς,lεP_{\varsigma,l}^{\varepsilon}, integrating by parts in the time derivative and changing variables from xx to x(εl1,εl2,0)x-(\varepsilon l_{1},\varepsilon l_{2},0)^{\top}, for hPε,l=hPε(t,x^+εl,x3)h_{P}^{\varepsilon,l}=h_{P}^{\varepsilon}(t,\hat{x}+\varepsilon l,x_{3}), yield

tθP(hPε,l),ψV(Pςε),T+IεKP(hPε,l)(hPε,l+e3),ψPς,Tε\displaystyle\langle\partial_{t}{\theta}_{P}(h_{P}^{\varepsilon,l}),\psi\rangle_{V(P^{\varepsilon}_{\varsigma})^{\prime},T}+\langle I_{\varepsilon}K_{P}(h_{P}^{\varepsilon,l})(\nabla h_{P}^{\varepsilon,l}+e_{3}),\nabla\psi\rangle_{P^{\varepsilon}_{\varsigma,T}} (4.19)
+εkΓhPε,lhSε,l,ψΓPς,Tε+𝒯pot,ψ(0,T)×ΓPς,0ε\displaystyle+\varepsilon k_{\Gamma}\langle h_{P}^{\varepsilon,l}-h_{S}^{\varepsilon,l},\psi\rangle_{\Gamma^{\varepsilon}_{P_{\varsigma},T}}+\langle\mathcal{T}_{\text{pot}},\psi\rangle_{(0,T)\times\Gamma^{\varepsilon}_{P_{\varsigma},0}} =0.\displaystyle=0.

Consider now the functions σδ+\sigma_{\delta}^{+}, σδ\sigma_{\delta}^{-}, ηP,δ+(hPε,l,w0)\eta_{P,\delta}^{+}(h_{P}^{\varepsilon,l},w^{0}), and ηP,δ(hPε,w0)\eta_{P,\delta}^{-}(h_{P}^{\varepsilon},w^{0}), with w0aV(Pςε)w^{0}-a\in V(P_{\varsigma}^{\varepsilon}), as in (3.23). The arguments similar to those in the proof of Theorem 3.7 and Lemma 3.8 yield that ζτ+=tt+τσδ+(hPε,lw0)κ(s)𝑑s\zeta^{+}_{\tau}=\int_{t}^{t+\tau}\!\!\sigma_{\delta}^{+}(h_{P}^{\varepsilon,l}-w^{0})\kappa(s)ds and ζτ=tt+τσδ(hPεw0)κ(s)𝑑s\zeta^{-}_{\tau}=\int_{t}^{t+\tau}\!\!\sigma_{\delta}^{-}(h_{P}^{\varepsilon}-w^{0})\kappa(s)ds, for τ>0\tau>0 and κC0(0,T)\kappa\in C_{0}^{\infty}(0,T), are admissible test functions in (4.19) and in the weak formulation of (2.2) over PςεP_{\varsigma}^{\varepsilon} respectively, and we obtain the following inequalities

ηP,δ+(hPε,l,w0),dκdtPς,Tε+IεKP(hPε,l)(hPε,l+e3),σδ+(hPε,lw0)κPς,Tε\displaystyle-\big{\langle}\eta_{P,\delta}^{+}(h_{P}^{\varepsilon,l},w^{0}),\frac{d\kappa}{dt}\big{\rangle}_{P_{\varsigma,T}^{\varepsilon}}+\big{\langle}I_{\varepsilon}K_{P}(h_{P}^{\varepsilon,l})(\nabla h_{P}^{\varepsilon,l}+e_{3}),\nabla\sigma_{\delta}^{+}(h_{P}^{\varepsilon,l}-w^{0})\kappa\big{\rangle}_{P^{\varepsilon}_{\varsigma,T}} (4.20)
+εkΓhPε,lhSε,l,σδ+(hPε,lw0)κΓPς,Tε+𝒯pot,σδ+(hPε,lw0)κΓPς,0,Tε0,\displaystyle+\varepsilon\,k_{\Gamma}\big{\langle}h_{P}^{\varepsilon,l}-h_{S}^{\varepsilon,l},\sigma_{\delta}^{+}(h_{P}^{\varepsilon,l}-w^{0})\kappa\big{\rangle}_{\Gamma^{\varepsilon}_{P_{\varsigma},T}}+\big{\langle}\mathcal{T}_{\text{pot}},\sigma_{\delta}^{+}(h_{P}^{\varepsilon,l}-w^{0})\kappa\big{\rangle}_{\Gamma_{P_{\varsigma},0,T}^{\varepsilon}}\leq 0,
ηP,δ(hPε,w0),dκdtPς,Tε+IεKP(hPε)(hPε+e3),σδ(hPεw0)κPς,Tε\displaystyle-\big{\langle}\eta_{P,\delta}^{-}(h_{P}^{\varepsilon},w^{0}),\frac{d\kappa}{dt}\big{\rangle}_{P_{\varsigma,T}^{\varepsilon}}+\big{\langle}I_{\varepsilon}K_{P}(h_{P}^{\varepsilon})(\nabla h_{P}^{\varepsilon}+e_{3}),\nabla\sigma_{\delta}^{-}(h_{P}^{\varepsilon}-w^{0})\kappa\big{\rangle}_{P^{\varepsilon}_{\varsigma,T}}
+εkΓhPεhSε,σδ(hPεw0)κΓPς,Tε+𝒯pot,σδ(hPεw0)κΓPς,0,Tε0.\displaystyle+\varepsilon\,k_{\Gamma}\big{\langle}h_{P}^{\varepsilon}-h_{S}^{\varepsilon},\sigma_{\delta}^{-}(h_{P}^{\varepsilon}-w^{0})\kappa\big{\rangle}_{\Gamma^{\varepsilon}_{P_{\varsigma},T}}+\big{\langle}\mathcal{T}_{\text{pot}},\sigma_{\delta}^{-}(h_{P}^{\varepsilon}-w^{0})\kappa\big{\rangle}_{\Gamma^{\varepsilon}_{P_{\varsigma},0,T}}\leq 0.

Considering a doubling of the time variable (t1,t2)(0,T)2(t_{1},t_{2})\in(0,T)^{2}, with hPε,l(x,t1,t2)=hPε,l(x,t1)h_{P}^{\varepsilon,l}(x,t_{1},t_{2})=h_{P}^{\varepsilon,l}(x,t_{1}), w0=hPε(x,t2)w^{0}=h_{P}^{\varepsilon}(x,t_{2}), and κ=κ(t2):t1κ(t1,t2)\kappa=\kappa(t_{2}):t_{1}\to\kappa(t_{1},t_{2}) in the first inequality, and hPε(x,t1,t2)=hPε(x,t2)h_{P}^{\varepsilon}(x,t_{1},t_{2})=h_{P}^{\varepsilon}(x,t_{2}), w0=hPε,l(x,t1)w^{0}=h_{P}^{\varepsilon,l}(x,t_{1}), and κ=κ(t1):t2κ(t1,t2)\kappa=\kappa(t_{1}):t_{2}\to\kappa(t_{1},t_{2}) in the second inequality in (4.20), with non-negative κC0((0,T)2)\kappa\in C_{0}^{\infty}((0,T)^{2}), and adding the resulting inequalities yield

(0,T)2[Pςε(ηP,δ+(hPε,l,hPε)t1κ+ηP,δ(hPε,hPε,l)t2κ)dx\displaystyle\int_{(0,T)^{2}}\!\Big{[}-\int_{P_{\varsigma}^{\varepsilon}}\!\Big{(}\eta_{P,\delta}^{+}(h_{P}^{\varepsilon,l},h_{P}^{\varepsilon})\partial_{t_{1}}\kappa+\eta_{P,\delta}^{-}(h_{P}^{\varepsilon},h_{P}^{\varepsilon,l})\partial_{t_{2}}\kappa\Big{)}dx
+IεKP(hPε,l)(hPε,l+e3)IεKP(hPε)(hPε+e3),σδ+(hPε,lhPε)κPςε\displaystyle\qquad+\big{\langle}I_{\varepsilon}K_{P}(h_{P}^{\varepsilon,l})(\nabla h_{P}^{\varepsilon,l}+e_{3})-I_{\varepsilon}K_{P}\big{(}h_{P}^{\varepsilon}\big{)}(\nabla h_{P}^{\varepsilon}+e_{3}),\nabla\sigma_{\delta}^{+}(h_{P}^{\varepsilon,l}-h_{P}^{\varepsilon})\kappa\big{\rangle}_{P^{\varepsilon}_{\varsigma}}
+εkΓ(hPε,lhPε)(hSε,lhSε),σδ+(hPε,lhPε)κΓPςε]dt1dt20.\displaystyle\qquad+\varepsilon\,k_{\Gamma}\big{\langle}(h_{P}^{\varepsilon,l}-h_{P}^{\varepsilon})-(h_{S}^{\varepsilon,l}-h_{S}^{\varepsilon}),\sigma_{\delta}^{+}(h_{P}^{\varepsilon,l}-h_{P}^{\varepsilon})\kappa\big{\rangle}_{\Gamma_{P_{\varsigma}}^{\varepsilon}}\Big{]}dt_{1}dt_{2}\leq 0.

Taking δ0\delta\to 0 in the above inequality and applying the arguments similar to the one used in the proof of (3.35), give

(θP(hPε,l)θP(hPε))+,t1κ+t2κ(0,T)2×Pςε+εkΓ(hPε,lhPε)+,κ(0,T)2×ΓPςε\displaystyle-\big{\langle}(\theta_{P}(h_{P}^{\varepsilon,l})-\theta_{P}(h_{P}^{\varepsilon}))^{+},\partial_{t_{1}}\kappa+\partial_{t_{2}}\kappa\big{\rangle}_{(0,T)^{2}\times P^{\varepsilon}_{\varsigma}}+\varepsilon k_{\Gamma}\big{\langle}(h_{P}^{\varepsilon,l}-h_{P}^{\varepsilon})^{+},\kappa\big{\rangle}_{(0,T)^{2}\times\Gamma^{\varepsilon}_{P_{\varsigma}}}
εkΓ(hSε,lhSε)+sign+(hPε,lhPε),κ(0,T)2×ΓPςε.\displaystyle\leq\varepsilon k_{\Gamma}\big{\langle}(h_{S}^{\varepsilon,l}-h_{S}^{\varepsilon})^{+}\text{sign}^{+}(h_{P}^{\varepsilon,l}-h_{P}^{\varepsilon}),\kappa\big{\rangle}_{(0,T)^{2}\times\Gamma^{\varepsilon}_{P_{\varsigma}}}. (4.21)

The function κϱ\kappa_{\varrho} defined as in (3.36) is admissible in (4), in place of κ\kappa, and applying the change of variables τ=t1t2\tau=t_{1}-t_{2} and denoting t=t1t=t_{1} we obtain

1ϱϑ(τϱ)0T[\displaystyle\int_{\mathbb{R}}\!\frac{1}{\varrho}\vartheta\big{(}\frac{\tau}{\varrho}\big{)}\!\int_{0}^{T}\!\!\Big{[} Pςε(θP(hPε,l(t))θP(hPε(tτ)))+tκ(tτ2)dx\displaystyle-\int_{P_{\varsigma}^{\varepsilon}}\!\!\big{(}\theta_{P}(h_{P}^{\varepsilon,l}(t))-\theta_{P}(h_{P}^{\varepsilon}(t-\tau))\big{)}^{+}\partial_{t}\kappa(t-\frac{\tau}{2})dx (4.22)
+εkΓΓP,ςε(hPε,l(t)hPε(tτ))+κ(tτ2)dγ]dtdτ\displaystyle+\varepsilon k_{\Gamma}\int_{\Gamma_{P,\varsigma}^{\varepsilon}}\!\!\!(h_{P}^{\varepsilon,l}(t)-h_{P}^{\varepsilon}(t-\tau))^{+}\kappa(t-\frac{\tau}{2})d\gamma\Big{]}dtd\tau
εkΓ1ϱϑ(τϱ)\displaystyle\leq\varepsilon k_{\Gamma}\int_{\mathbb{R}}\!\frac{1}{\varrho}\vartheta\big{(}\frac{\tau}{\varrho}\big{)} 0TΓP,ςε(hSε,l(t)hSε(tτ))+sign+(hPε,lhPε)κ(tτ2)𝑑γ𝑑t𝑑τ.\displaystyle\int_{0}^{T}\!\!\!\int_{\Gamma_{P,\varsigma}^{\varepsilon}}\!\!\!\!\!(h_{S}^{\varepsilon,l}(t)-h_{S}^{\varepsilon}(t-\tau))^{+}\text{sign}^{+}(h_{P}^{\varepsilon,l}-h_{P}^{\varepsilon})\kappa(t-\frac{\tau}{2})d\gamma dtd\tau.

Taking ϱ0\varrho\to 0, applying the integration by parts, and using the compact support of κ\kappa, imply

0Tκ(t)tPςε(θP(hPε,l(t))θP(hPε(t)))+𝑑x𝑑tεkΓ0Tκ(t)ΓP,ςε|hSε,l(t)hSε(t)|𝑑γ𝑑t.\displaystyle\int_{0}^{T}\!\!\!\!\!\kappa(t)\partial_{t}\!\!\int_{P_{\varsigma}^{\varepsilon}}\!\!\!(\theta_{P}(h_{P}^{\varepsilon,l}(t))-\theta_{P}(h_{P}^{\varepsilon}(t)))^{+}dxdt\leq\varepsilon k_{\Gamma}\!\!\int_{0}^{T}\!\!\!\!\!\kappa(t)\!\!\int_{\Gamma_{P,\varsigma}^{\varepsilon}}\!\!\!\!\!\!\!|h_{S}^{\varepsilon,l}(t)-h_{S}^{\varepsilon}(t)|d\gamma dt. (4.23)

Exchanging hPε,lh_{P}^{\varepsilon,l} and hPεh_{P}^{\varepsilon} in the calculations above yields (4.23) for (θP(hPε)θP(hPε,l))+(\theta_{P}(h_{P}^{\varepsilon})-\theta_{P}(h_{P}^{\varepsilon,l}))^{+}. Applying the trace theorem over the unit cell YY, together with the standard scaling argument, and using estimates in Lemma 4.1, yield

0Tκ(t)tPςε|θP(hPε,l(t))θP(hPε(t))|𝑑x𝑑tCεl\int_{0}^{T}\!\!\!\!\kappa(t)\,\partial_{t}\int_{P_{\varsigma}^{\varepsilon}}|\theta_{P}(h_{P}^{\varepsilon,l}(t))-\theta_{P}(h_{P}^{\varepsilon}(t))|dxdt\leq C\varepsilon l (4.24)

for any non-negative κC0(0,T)\kappa\in C^{\infty}_{0}(0,T). Using the regularity of initial conditions and Lipschitz continuity of θP\theta_{P}, from (4.24) we obtain

θP(hPε,l(t))θP(hPε(t))L1(Pςε)Cεl, for t(0,T].\|\theta_{P}(h_{P}^{\varepsilon,l}(t))-\theta_{P}(h_{P}^{\varepsilon}(t))\|_{L^{1}(P^{\varepsilon}_{\varsigma})}\leq C\varepsilon l,\quad\text{ for }\;t\in(0,T].

Considering |z|ς|z|\leq\varsigma and using the definition of the unfolding operator imply

ΩTYP|θP(𝒯ε(hPε)(t,x^+z,x3,y))θP(𝒯ε(hPε)(t,x,y))|𝑑y𝑑x𝑑t\displaystyle\int_{\Omega_{T}}\int_{Y_{P}}|\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})(t,\hat{x}+z,x_{3},y))-\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})(t,x,y))|dydxdt (4.25)
k{0,1}20TPςε|θP(hPε,lk)θP(hPε)|𝑑x𝑑tκ(ς)0 as ς0,\displaystyle\leq\sum_{k\in\{0,1\}^{2}}\int_{0}^{T}\int_{P^{\varepsilon}_{\varsigma}}|\theta_{P}(h^{\varepsilon,l_{k}}_{P})-\theta_{P}(h^{\varepsilon}_{P})|dxdt\leq\kappa(\varsigma)\to 0\;\;\text{ as }\;\;\varsigma\to 0,

where lk=k+[z/ε]l_{k}=k+[z/\varepsilon] and |εlk|ς|\varepsilon l_{k}|\leq\varsigma for all ε>0\varepsilon>0, such that εε0\varepsilon\leq\varepsilon_{0} for some ε0>0\varepsilon_{0}>0. For the finite number of ε>ε0\varepsilon>\varepsilon_{0} estimate (4.25) follows from the continuity of the L2L^{2}-norm. The estimates for εx^hPε\varepsilon\nabla_{\hat{x}}h^{\varepsilon}_{P} and x3hPε\partial_{x_{3}}h^{\varepsilon}_{P} ensure, for r2r\in\mathbb{R}^{2} and r1r_{1}\in\mathbb{R},

θP(𝒯ε(hPε)(,,,y+r))θP(𝒯ε(hPε))L2(ΩT×YP)2\displaystyle\|\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})(\cdot,\cdot,\cdot,y+r))-\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P}))\|^{2}_{L^{2}(\Omega_{T}\times Y_{P})} (4.26)
C|r|2y𝒯ε(θP(hPε))L2(ΩT×YP)2C|r|2ε2x^hPεL2(PTε)2C|r|2,\displaystyle\qquad\qquad\leq C|r|^{2}\|\nabla_{y}\mathcal{T}^{\varepsilon}(\theta_{P}(h^{\varepsilon}_{P}))\|^{2}_{L^{2}(\Omega_{T}\times Y_{P})}\leq C|r|^{2}\varepsilon^{2}\|\nabla_{\hat{x}}h^{\varepsilon}_{P}\|^{2}_{L^{2}(P^{\varepsilon}_{T})}\leq C|r|^{2},
θP(𝒯ε(hPε)(,,x3+r1,))θP(𝒯ε(hPε))L2(ΩT×YP)2Cr12x3hPεL2(PTε)2Cr12.\displaystyle\|\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})(\cdot,\cdot,x_{3}+r_{1},\cdot))-\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P}))\|^{2}_{L^{2}(\Omega_{T}\times Y_{P})}\leq Cr_{1}^{2}\|\partial_{x_{3}}h^{\varepsilon}_{P}\|^{2}_{L^{2}(P^{\varepsilon}_{T})}\leq Cr_{1}^{2}.

Using (4.25), (4.26), and the second estimate in (4.15), and applying the compactness theorem, see [41] and boundedness of θP\theta_{P}, yields the strong convergence of θP(𝒯ε(hPε))\theta_{P}(\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})) in L2((0,T)×Ω×YP)L^{2}((0,T)\times\Omega\times Y_{P}). This, together with the monotonicity and continuity of θP\theta_{P} and two-scale convergence of hPεh^{\varepsilon}_{P} to hPχYPh_{P}\chi_{Y_{P}}, implies 𝒯ε(hPε)hP\mathcal{T}^{\varepsilon}(h^{\varepsilon}_{P})\to h_{P} a.e. in (0,T)×Ω×YP(0,T)\times\Omega\times Y_{P}. Hence 𝒯ε(θP(hPε))θP(hP)\mathcal{T}^{\varepsilon}(\theta_{P}(h^{\varepsilon}_{P}))\to\theta_{P}(h_{P}) strongly in L2((0,T)×Ω×YP)L^{2}((0,T)\times\Omega\times Y_{P}), which, applying the properties of the unfolding operator, implies the strong two-scale convergence of θP(hPε)\theta_{P}(h^{\varepsilon}_{P}), stated in the lemma. ∎

Theorem 4.7.

A sequence of solutions (hSε,hPε)(h_{S}^{\varepsilon},h_{P}^{\varepsilon}) of microscopic model (2.1)–(2.4) converges to solution hSL2(0,T;V(Ω))h_{S}\in L^{2}(0,T;V(\Omega)), hPaL2(0,T;U(Ω))h_{P}-a\in L^{2}(0,T;U(\Omega)) of the macroscopic problem

tθS(hS)(KS,hom(hS)(hS+e3))=kΓϑΓ(hPhS)\displaystyle\partial_{t}\theta_{S}^{\ast}(h_{S})-\nabla\cdot(K_{S,\rm hom}(h_{S})(\nabla h_{S}+e_{3}))=k_{\Gamma}\vartheta_{\Gamma}(h_{P}-h_{S}) in (0,T)×Ω,\displaystyle\text{in }(0,T)\times\Omega, (4.27)
tθP(hP)x3(KP(hP)(x3hP+1))=kΓϑΓ,P(hShP)\displaystyle\partial_{t}\theta_{P}(h_{P})-\partial_{x_{3}}(K_{P}(h_{P})(\partial_{x_{3}}h_{P}+1))=k_{\Gamma}\vartheta_{\Gamma,P}(h_{S}-h_{P}) in (0,T)×Ω,\displaystyle\text{in }(0,T)\times\Omega,
KS,hom(hS)(hS+e3)ν=0\displaystyle K_{S,\rm hom}(h_{S})(\nabla h_{S}+e_{3})\cdot\nu=0\quad on (0,T)×ΓN,\displaystyle\text{on }(0,T)\times\Gamma_{N},
KS,hom(hS)(hS+e3)ν=ϑSf(hS)\displaystyle K_{S,\rm hom}(h_{S})(\nabla h_{S}+e_{3})\cdot\nu=\vartheta_{S}f(h_{S})\quad on (0,T)×Γ0,\displaystyle\text{on }(0,T)\times\Gamma_{0},
KP(hP)(x3hP+1)=𝒯pot\displaystyle K_{P}(h_{P})(\partial_{x_{3}}h_{P}+1)=\mathcal{T}_{\rm pot} on (0,T)×Γ0,\displaystyle\text{on }(0,T)\times\Gamma_{0},
hS(0)=hS,0,hP(0)=hP,0\displaystyle h_{S}(0)=h_{S,0},\qquad h_{P}(0)=h_{P,0} in Ω,\displaystyle\text{in }\Omega,

where θS=ϑRθR(hS)+ϑBθB(hS)\theta_{S}^{\ast}=\vartheta_{R}\theta_{R}(h_{S})+\vartheta_{B}\theta_{B}(h_{S}), ϑJ=|J^|/|Y|\vartheta_{J}=|\hat{J}|/|Y|, for J=R,B,SJ=R,B,S, ϑΓ,P=|ΓP|/|YP|\vartheta_{\Gamma,P}=|\Gamma_{P}|/|Y_{P}|, ϑΓ=|ΓP|/|Y|\vartheta_{\Gamma}=|\Gamma_{P}|/|Y|, and ΓN=Ω^×(L3,0)\Gamma_{N}=\partial\hat{\Omega}\times(-L_{3},0),

V(Ω)={vH1(Ω):v=0 on ΓL3},U(Ω)={w,x3wL2(Ω):w=0 on ΓL3},V(\Omega)=\{v\in H^{1}(\Omega):v=0\text{ on }\Gamma_{L_{3}}\},\quad U(\Omega)=\{w,\partial_{x_{3}}w\in L^{2}(\Omega):w=0\text{ on }\Gamma_{L_{3}}\},

and

KS,hom,ij(hS)=1|Y|YS[KS(y,hS)δij+KS(y,hS)yiwj]𝑑y,i,j=1,2,\displaystyle K_{S,\rm hom,ij}(h_{S})=\frac{1}{|Y|}\int_{Y_{S}}\Big{[}K_{S}(y,h_{S})\delta_{ij}+K_{S}(y,h_{S})\partial_{y_{i}}w^{j}\big{]}dy,\quad i,j=1,2,
KS,hom,3j(hS)=KS,hom,j3(hS)=1|Y|YSKS(y,hS)𝑑yδ3j,j=1,2,3,\displaystyle K_{S,\rm hom,3j}(h_{S})=K_{S,\rm hom,j3}(h_{S})=\frac{1}{|Y|}\int_{Y_{S}}K_{S}(y,h_{S})dy\,\delta_{3j},\quad j=1,2,3,

with wjw^{j}, for j=1,2j=1,2, being the solution of the unit cell problems

y(KS(y,hS)(ywj+ej))\displaystyle\nabla_{y}\cdot\big{(}K_{S}(y,h_{S})(\nabla_{y}w^{j}+e_{j})\big{)} =0 in YS,\displaystyle=0\;\;\;\text{ in }Y_{S}, (4.28)
KS(y,hS)(ywj+ej)ν\displaystyle K_{S}(y,h_{S})(\nabla_{y}w^{j}+e_{j})\cdot\nu =0 on ΓP,\displaystyle=0\;\;\;\text{ on }\Gamma_{P},\; wjY periodic.\displaystyle w^{j}\;\;Y-\text{ periodic}.
Proof.

Considering ϕ(t,x)=ϕ1(t,x)+εϕ2(t,x,x^/ε)\phi(t,x)=\phi_{1}(t,x)+\varepsilon\phi_{2}(t,x,\hat{x}/\varepsilon), with ϕ1C0(0,T;CΓL3(Ω))\phi_{1}\in C^{\infty}_{0}(0,T;C^{\infty}_{\Gamma_{L_{3}}}\!\!(\Omega)) and ϕ2C0((0,T)×Ω;Cper(Y))\phi_{2}\in C^{\infty}_{0}((0,T)\times\Omega;C^{\infty}_{\rm per}(Y)), and ψC0(0,T;CΓL3(Ω))\psi\in C^{\infty}_{0}(0,T;C^{\infty}_{\Gamma_{L_{3}}}(\Omega)) as test functions in (3.1) and (3.2) and applying the unfolding operator we have

θS(y,𝒯ε(hSε)),t𝒯ε(ϕ1+εϕ2)ΩT×S^+kΓ𝒯ε(hSε)𝒯ε(hPε),𝒯ε(ϕ1+εϕ2)ΩT×ΓP\displaystyle-\big{\langle}{\theta}_{S}(y,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})),\partial_{t}\mathcal{T}^{\varepsilon}(\phi_{1}+\varepsilon\phi_{2})\big{\rangle}_{\Omega_{T}\times\hat{S}}+k_{\Gamma}\big{\langle}\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})-\mathcal{T}^{\varepsilon}(h_{P}^{\varepsilon}),\mathcal{T}^{\varepsilon}(\phi_{1}+\varepsilon\phi_{2})\big{\rangle}_{\Omega_{T}\times\Gamma_{P}} (4.29)
+KS(y,𝒯ε(hSε))(𝒯ε(hSε)+e3),𝒯ε(ϕ1)+ε𝒯ε(ϕ2)ΩT×S^\displaystyle+\big{\langle}K_{S}(y,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}))(\mathcal{T}^{\varepsilon}(\nabla h_{S}^{\varepsilon})+{e}_{3}),\mathcal{T}^{\varepsilon}(\nabla\phi_{1})+\varepsilon\mathcal{T}^{\varepsilon}(\nabla\phi_{2})\big{\rangle}_{\Omega_{T}\times\hat{S}}
=f(𝒯ε(hSε)),𝒯ε(ϕ1)+ε𝒯ε(ϕ2)Γ0,T×S^,\displaystyle\qquad=-\big{\langle}f(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})),\mathcal{T}^{\varepsilon}(\phi_{1})+\varepsilon\mathcal{T}^{\varepsilon}(\phi_{2})\big{\rangle}_{\Gamma_{0,T}\times\hat{S}},
θP(𝒯ε(hPε)),t𝒯ε(ψ)ΩT×YP+kΓ𝒯ε(hPε)𝒯ε(hSε),𝒯ε(ψ)ΩT×ΓP\displaystyle-\big{\langle}{\theta}_{P}(\mathcal{T}^{\varepsilon}(h_{P}^{\varepsilon})),\partial_{t}\mathcal{T}^{\varepsilon}(\psi)\big{\rangle}_{\Omega_{T}\times Y_{P}}+k_{\Gamma}\big{\langle}\mathcal{T}^{\varepsilon}(h_{P}^{\varepsilon})-\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}),\mathcal{T}^{\varepsilon}(\psi)\big{\rangle}_{\Omega_{T}\times\Gamma_{P}} (4.30)
+IεKP(𝒯ε(hPε))(𝒯ε(hPε)+e3),𝒯ε(ψ)ΩT×YP=𝒯pot,𝒯ε(ψ)Γ0,T×YP.\displaystyle+\big{\langle}I_{\varepsilon}K_{P}(\mathcal{T}^{\varepsilon}(h_{P}^{\varepsilon}))(\mathcal{T}^{\varepsilon}(\nabla h_{P}^{\varepsilon})+{e}_{3}),\mathcal{T}^{\varepsilon}(\nabla\psi)\big{\rangle}_{\Omega_{T}\times Y_{P}}=-\big{\langle}\mathcal{T}_{\text{pot}},\mathcal{T}^{\varepsilon}(\psi)\big{\rangle}_{\Gamma_{0,T}\times Y_{P}}.

Using the properties of the unfolding operator 𝒯ε\mathcal{T}^{\varepsilon}, i.e. 𝒯ε(ϕ)ϕ\mathcal{T}^{\varepsilon}(\phi)\to\phi for ϕLp((0,T)×A)\phi\in L^{p}((0,T)\times A), with 1<p<1<p<\infty and A=ΩA=\Omega or A=ΩA=\partial\Omega, 𝒯ε(ϕ(,,/ε))ϕ\mathcal{T}^{\varepsilon}(\phi(\cdot,\cdot,\cdot/\varepsilon))\to\phi for ϕLp((0,T)×A;Cper(Y))\phi\in L^{p}((0,T)\times A;C_{\rm per}(Y)), and ε𝒯ε(ϕ(,,/ε))yϕ\varepsilon\mathcal{T}^{\varepsilon}(\nabla\phi(\cdot,\cdot,\cdot/\varepsilon))\to\nabla_{y}\phi for ϕLp(0,T;W1,p(Ω);Cper1(Y))\phi\in L^{p}(0,T;W^{1,p}(\Omega);C^{1}_{\rm per}(Y)), and the relations between the two-scale (strong two-scale) convergence of a sequence and weak (strong) convergence of the corresponding unfolded sequence, see e.g. [11], together with the convergence results in Lemmas 4.24.34.5, and 4.6, and taking in (4.29) and (4.30) the limit as ε0\varepsilon\to 0 we obtain

θS(y,hS),tϕ1ΩT×S^+kΓhShP,ϕ1ΩT×ΓP+f(hS),ϕ1Γ0,T×S^\displaystyle-\big{\langle}{\theta}_{S}(y,h_{S}),\partial_{t}\phi_{1}\big{\rangle}_{\Omega_{T}\times\hat{S}}+k_{\Gamma}\big{\langle}h_{S}-h_{P},\phi_{1}\big{\rangle}_{\Omega_{T}\times\Gamma_{P}}+\big{\langle}f(h_{S}),\phi_{1}\big{\rangle}_{\Gamma_{0,T}\times\hat{S}} (4.31)
+KS(y,hS)(hS+y,0hS,1+e3),ϕ1+y,0ϕ2ΩT×S^=0,\displaystyle+\big{\langle}K_{S}(y,h_{S})(\nabla h_{S}+\nabla_{y,0}h_{S,1}+{e}_{3}),\nabla\phi_{1}+\nabla_{y,0}\phi_{2}\big{\rangle}_{\Omega_{T}\times\hat{S}}=0,
θP(hP),tψΩT×YP+KP(hP)(x3hP+1),x3ψΩT×YP\displaystyle-\big{\langle}{\theta}_{P}(h_{P}),\partial_{t}\psi\big{\rangle}_{\Omega_{T}\times Y_{P}}+\big{\langle}K_{P}(h_{P})(\partial_{x_{3}}h_{P}+1),\partial_{x_{3}}\psi\big{\rangle}_{\Omega_{T}\times Y_{P}} (4.32)
+kΓhPhS,ψΩT×ΓP+𝒯pot,ψΓ0,T×YP\displaystyle+k_{\Gamma}\big{\langle}h_{P}-h_{S},\psi\big{\rangle}_{\Omega_{T}\times\Gamma_{P}}+\big{\langle}\mathcal{T}_{\text{pot}},\psi\big{\rangle}_{\Gamma_{0,T}\times Y_{P}} =0.\displaystyle=0.

Notice that uniform boundedness and continuity of KJK_{J}, for J=P,R,BJ=P,R,B, and convergence a.e. of 𝒯ε(hSε)\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}) and 𝒯ε(hPε)\mathcal{T}^{\varepsilon}(h_{P}^{\varepsilon}), together with the Lebesgue dominated convergence theorem, imply strong convergence KS(y,𝒯ε(hSε))KS(y,hS)K_{S}(y,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}))\to K_{S}(y,h_{S}) in Lp((0,T)×Ω×S^)L^{p}((0,T)\times\Omega\times\hat{S}) and KP(𝒯ε(hPε))KP(hP)K_{P}(\mathcal{T}^{\varepsilon}(h_{P}^{\varepsilon}))\to K_{P}(h_{P}) in Lp((0,T)×Ω×YP)L^{p}((0,T)\times\Omega\times Y_{P}), for any 1<p<1<p<\infty. To show the convergence of f(𝒯ε(hSε))f(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})) we consider

θS(,𝒯ε(hSε))θS(,hS)L2(Γ0,T×S^)2CJ=R,B(θJ(hSε)θJ(hS))2L1(ΓJ,0,Tε)\displaystyle\|\theta_{S}(\cdot,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}))-\theta_{S}(\cdot,h_{S})\|^{2}_{L^{2}(\Gamma_{0,T}\times\hat{S})}\leq C\!\!\!\sum_{J=R,B}\!\!\!\|(\theta_{J}(h_{S}^{\varepsilon})-\theta_{J}(h_{S}))^{2}\|_{L^{1}(\Gamma^{\varepsilon}_{J,0,T})} (4.33)
CJ=R,B[(θJ(hSε)θJ(hS))2L1(JTε)+Iε(θJ(hSε)θJ(hS))2L1(JTε)]\displaystyle\quad\leq C\!\!\!\sum_{J=R,B}\!\!\Big{[}\|(\theta_{J}(h_{S}^{\varepsilon})-\theta_{J}(h_{S}))^{2}\|_{L^{1}(J^{\varepsilon}_{T})}+\|I_{\varepsilon}\nabla(\theta_{J}(h_{S}^{\varepsilon})-\theta_{J}(h_{S}))^{2}\|_{L^{1}(J^{\varepsilon}_{T})}\Big{]}
C[θS(,hSε)θS(,hS)L2(STε)2\displaystyle\quad\leq C\Big{[}\|\theta_{S}(\cdot,h_{S}^{\varepsilon})-\theta_{S}(\cdot,h_{S})\|^{2}_{L^{2}(S^{\varepsilon}_{T})}
+θS(,hSε)θS(,hS)L2(STε)(hSεL2(STε)+hSL2(STε))].\displaystyle\qquad+\|\theta_{S}(\cdot,h_{S}^{\varepsilon})-\theta_{S}(\cdot,h_{S})\|_{L^{2}(S^{\varepsilon}_{T})}\big{(}\|\nabla h_{S}^{\varepsilon}\|_{L^{2}(S^{\varepsilon}_{T})}+\|\nabla h_{S}\|_{L^{2}(S^{\varepsilon}_{T})}\big{)}\Big{]}.

Here we used the trace theorem and the standard scaling argument to obtain

uL1(ΓJ,0ε)C(uL1(Jε)+IεuL1(Jε)), for uW1,1(Jε),J=R,B.\|u\|_{L^{1}(\Gamma^{\varepsilon}_{J,0})}\leq C\big{(}\|u\|_{L^{1}(J^{\varepsilon})}+\|I_{\varepsilon}\nabla u\|_{L^{1}(J^{\varepsilon})}\big{)},\quad\text{ for }\;u\in W^{1,1}(J^{\varepsilon}),\;\;J=R,B.

Then using in (4.33) the strong convergence of θS(y,𝒯ε(hSε))\theta_{S}(y,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})) in L2((0,T)×Ω×S^)L^{2}((0,T)\times\Omega\times\hat{S}), see Lemma 4.5, we obtain the strong convergence of θS(y,𝒯ε(hSε))\theta_{S}(y,\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})) in L2((0,T)×Γ0×S^)L^{2}((0,T)\times\Gamma_{0}\times\hat{S}). The monotonicity and continuity of θS(y,)\theta_{S}(y,\cdot) ensure 𝒯ε(hSε)hS\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon})\to h_{S} a.e. in (0,T)×Γ0×S^(0,T)\times\Gamma_{0}\times\hat{S} and then the continuity and boundedness of ff imply f(𝒯ε(hSε))f(hS)f(\mathcal{T}^{\varepsilon}(h_{S}^{\varepsilon}))\to f(h_{S}) in L2((0,T)×Γ0×S^)L^{2}((0,T)\times\Gamma_{0}\times\hat{S}).

Using the standard arguments, see e.g. [1, 12], from (4.31) by considering ϕ1=0\phi_{1}=0 we obtain the unit cell problems (4.28) and the formula for KS,homK_{S,{\rm hom}}. Then, considering (4.31) with ϕ2=0\phi_{2}=0 and first ϕ1C01((0,T)×Ω)\phi_{1}\in C^{1}_{0}((0,T)\times\Omega) and then ϕ1C01(0,T;CΓL31(Ω))\phi_{1}\in C_{0}^{1}(0,T;C^{1}_{\Gamma_{L_{3}}}(\Omega)) and (4.32) first with ψC01((0,T)×Ω)\psi\in C^{1}_{0}((0,T)\times\Omega) and then ψC01(0,T;CΓL31(Ω))\psi\in C_{0}^{1}(0,T;C^{1}_{\Gamma_{L_{3}}}(\Omega)) yields the macroscopic equations (4.27). Considering ϕ2=0\phi_{2}=0 and ϕ1,ψC1([0,T];C01(Ω))\phi_{1},\psi\in C^{1}([0,T];C^{1}_{0}(\Omega)), with ϕ1(T)=ψ(T)=0\phi_{1}(T)=\psi(T)=0, as test functions in the weak formulation of (2.1), (2.3) and (2.2), (2.4) respectively, and using monotonicity of θS\theta_{S}^{\ast} and θP\theta_{P} implies that hSh_{S} and hPh_{P} satisfy the corresponding initial conditions.

Using Assumption 3.1, together with the Lipschitz continuity of KBK_{B}, KRK_{R} and KPK_{P}, and ff being non-decreasing, in the similar way as in the proof of Theorem 3.7, we obtain the uniqueness of solutions of the macroscopic model (4.27). ∎

Acknowledgements

AM was supported by the EPSRC Centre for Doctoral Training in Mathematical Modelling, Analysis & Computation (MAC-MIGS) funded by the UK Engineering and Physical Sciences Research Council (EPSRC) grant EP/S023291/1, Heriot-Watt University, and the University of Edinburgh.

Appendix: Admissible functions for water content, hydraulic conductivity and water flux at the upper soil surface

The function for soil water content θS\theta_{S} can be defined in accordance with the formulation of [43] and extended for positive values of soil water pressure head

θS(hS)={θS,sat1(1+|αhS|n)m+1hS>0,θS,res+θS,satθS,res(1+|αhS|n)mhS0.\theta_{S}(h_{S})=\begin{cases}\frac{\theta_{S,\text{sat}}-1}{(1+\lvert\alpha h_{S}\rvert^{n})^{m}}+1&h_{S}>0,\\ \theta_{S,\text{res}}+\frac{\theta_{S,\text{sat}}-\theta_{S,\text{res}}}{(1+\lvert\alpha h_{S}\rvert^{n})^{m}}&h_{S}\leq 0.\end{cases} (4.34)

Here the residual and saturated soil water contents are θS,res\theta_{S,\text{res}} and θS,sat(L3L3)\theta_{S,\text{sat}}~{}(\text{L}^{3}\text{L}^{-3}) respectively, and constants α(L1)\alpha~{}(\text{L}^{-1})n>1n>1 and m=11/nm=1-1/n are shape parameters, where we have different sets of there parameters for bulk soil and rhizosphere. Water content θP\theta_{P} for root tissue can be defined using the models in [25, 9] and extending for positive root tissue pressure heads in a similar way as (4.34):

θP(hP)={θP,sat1(1+|αhP|n)m+1hP>0,(θP,sat+hPE)haehP<0,(θP,sat+haeE)(hPhae)λPhP<hae.\theta_{P}(h_{P})=\begin{cases}\frac{\theta_{P,\text{sat}}-1}{(1+\lvert\alpha h_{P}\rvert^{n})^{m}}+1&h_{P}>0,\\ \Big{(}\theta_{P,\text{sat}}+\frac{h_{P}}{E}\Big{)}&h_{\text{ae}}\leq h_{P}<0,\\ \Big{(}\theta_{P,\text{sat}}+\frac{h_{\text{ae}}}{E}\Big{)}\Big{(}\frac{h_{P}}{h_{\text{ae}}}\Big{)}^{-\lambda_{P}}&h_{P}<h_{\text{ae}}.\end{cases} (4.35)

Here θP,sat\theta_{P,\text{sat}} (L3L3)\big{(}\text{L}^{3}\text{L}^{-3}\big{)} is the root tissue porosity (or root xylem saturated water content), the air entry pressure head is haeh_{\text{ae}} (L)(\text{L}), the root xylem elastic modulus is EE (L)(L), λP\lambda_{P} is the Brooks and Corey exponent, and α\alphan>1n>1 and m=11/nm=1-1/n are as in (4.34). A suitable expression for KSK_{S} can be defined by taking a regularisation of the Van Genuchten [43] formulation and extending it for positive pressure heads. The first step is to define the regularisation of the water content function

θS,δ(hS)={θS,sat1(1+|αhS|n)m+1δ2hS>0,θS,res+θS,sat(θS,res+δ)(1+|αhS|n)m+δ2hS0,\theta_{S,\delta}(h_{S})=\begin{cases}\frac{\theta_{S,\text{sat}}-1}{(1+\lvert\alpha h_{S}\rvert^{n})^{m}}+1-\frac{\delta}{2}&h_{S}>0,\\ \theta_{S,\text{res}}+\frac{\theta_{S,\text{sat}}-(\theta_{S,\text{res}}+\delta)}{(1+\lvert\alpha h_{S}\rvert^{n})^{m}}+\frac{\delta}{2}&h_{S}\leq 0,\end{cases}

where 0<δ<<θS,satθS,res0<\delta<<\theta_{S,\text{sat}}-\theta_{S,\text{res}}. The regularised soil hydraulic conductivity KS,δ(hS)K_{S,\delta}(h_{S}) satisfying the conditions in Theorems 3.5 and 3.7 is given as

{KS,sat[1δ2(θS,satθS,res)]l[1[1[1δ2(θS,satθS,res)]1m]m]2,hS>0,KS,sat[θS,δ(hS)θS,resθS,satθS,res]l[1[1[θS,δ(hS)θS,resθS,satθS,res]1m]m]2,hS0.\begin{cases}K_{S,\text{sat}}\Big{[}1-\frac{\delta}{2(\theta_{S,\text{sat}}-\theta_{S,\text{res}})}\Big{]}^{l}\Big{[}1-\Big{[}1-\Big{[}1-\frac{\delta}{2(\theta_{S,\text{sat}}-\theta_{S,\text{res}})}\Big{]}^{\frac{1}{m}}\Big{]}^{m}\Big{]}^{2},&h_{S}>0,\\ K_{S,\text{sat}}\Big{[}\frac{\theta_{S,\delta}(h_{S})-\theta_{S,\text{res}}}{\theta_{S,\text{sat}}-\theta_{S,\text{res}}}\Big{]}^{l}\Big{[}1-\Big{[}1-\Big{[}\frac{\theta_{S,\delta}(h_{S})-\theta_{S,\text{res}}}{\theta_{S,\text{sat}}-\theta_{S,\text{res}}}\Big{]}^{\frac{1}{m}}\Big{]}^{m}\Big{]}^{2},&h_{S}\leq 0.\end{cases}

In a similar way, we define a regularised expression for root tissue water content

θP,δ(hP)={θP,sat1(1+|αhP|n)m+1hP>0,(θP,sat+hPE)haehP<0,(θP,sat+haeE)(hPhae)λP+δ(1ehPhae)hP<hae,\theta_{P,\delta}(h_{P})=\begin{cases}\frac{\theta_{P,\text{sat}}-1}{(1+|\alpha h_{P}|^{n})^{m}}+1&h_{P}>0,\\ \Big{(}\theta_{P,\text{sat}}+\frac{h_{P}}{E}\Big{)}&h_{\text{ae}}\leq h_{P}<0,\\ \Big{(}\theta_{P,\text{sat}}+\frac{h_{\text{ae}}}{E}\Big{)}\Big{(}\frac{h_{P}}{h_{\text{ae}}}\Big{)}^{-\lambda_{P}}+\delta(1-e^{h_{P}-h_{\text{ae}}})&h_{P}<h_{\text{ae}},\end{cases}

where δ>0\delta>0, and a regularised version of the function κP,δ(hP)\kappa_{P,\delta}(h_{P}) given as

{1,hP0,(1+haeθP,satE)2λP+1+[1(1+haeθP,satE)2λP+1](hPhaehae)2,haehP<0,((1+haeθP,satE)(hPhae)λP+δθP,sat(1e(hPhae)))2λP+1,hp<hae<0.\begin{cases}1,&h_{P}\geq 0,\\ \big{(}1+\frac{h_{\text{ae}}}{\theta_{P,\text{sat}}E}\big{)}^{\frac{2}{\lambda_{P}}+1}+\Big{[}1-\big{(}1+\frac{h_{\text{ae}}}{\theta_{P,\text{sat}}E}\big{)}^{\frac{2}{\lambda_{P}}+1}\Big{]}\big{(}\frac{h_{P}-h_{\text{ae}}}{h_{\text{ae}}}\big{)}^{2},&h_{\text{ae}}\leq h_{P}<0,\\ \Big{(}\big{(}1+\frac{h_{\text{ae}}}{\theta_{P,\text{sat}}E}\big{)}\big{(}\frac{h_{P}}{h_{\text{ae}}}\big{)}^{-\lambda_{P}}+\frac{\delta}{\theta_{P,\text{sat}}}\big{(}1-e^{(h_{P}-h_{\text{ae}})}\big{)}\Big{)}^{\frac{2}{\lambda_{P}}+1},&h_{p}<h_{\text{ae}}<0.\end{cases}

An admissible function f:f:\mathbb{R}\to\mathbb{R}, for the water flux at the upper soil surface ΓS,0\Gamma_{S,0}, is

f(hS)=Ke(hS)EToP+RO(hS),f(h_{S})=\text{K}_{\text{e}}(h_{S})\text{ET}_{\text{o}}-\text{P}+\text{RO}(h_{S}),

where ETo\text{ET}_{\text{o}} is the reference evapotranspiration (LT1)\big{(}\text{LT}^{-1}\big{)}, see [3], the function Ke:[0,)]\text{K}_{\text{e}}:\mathbb{R}\to[0,\infty)] controls the amount of evaporation from the soil surface, and P (LT1)\big{(}\text{LT}^{-1}\big{)} is the precipitation. The function RO (LT1)\big{(}\text{LT}^{-1}\big{)} incorporates runoff, which occurs when precipitation lands on an already saturated soil surface and cannot infiltrate downwards. The evaporation function Ke\text{K}_{\text{e}} is defined in the same way as in [32] and the runoff function is defined as

RO(hS)=P(1exp(CRO12)+exp(CRO(hS+CRO12))),\text{RO}(h_{S})=\frac{\text{P}}{\bigg{(}1-\exp(-C_{\text{RO}}^{\frac{1}{2}})+\exp(-C_{\text{RO}}\Big{(}h_{S}+C_{\text{RO}}^{-\frac{1}{2}}\Big{)})\bigg{)}},

where CRO>0C_{\text{RO}}>0.

References

  • [1] Allaire, G. Homogenization and two-scale convergence. SIAM J Math Anal 23 (1992), 1482–1518.
  • [2] Allaire, G., Damlamian, A., and Hornung, U. Two-scale convergence on periodic surfaces and applications. Proceedings of Internat Conference Mathematical Modelling of Flow Through Porous Media (1996), 15–25.
  • [3] Allen, R. G., Pereira, L. S., Raes, D., Smith, M., et al. Crop evapotranspiration-guidelines for computing crop water requirements-fao irrigation and drainage paper 56. Fao, Rome 300, 9 (1998), D05109.
  • [4] Alt, H., and Luckhaus, S. Quasilinear elliptic-parabolic differential equations. Math. Z 183, 3 (1983), 311–341.
  • [5] Amaziane, B., Jurak, M., and Vrbaški, A. Homogenization results for a coupled system modelling immiscible compressible two-phase flow in porous media by the concept of global pressure. Applicable Anal 92 (2013), 1417–1433.
  • [6] Arbogast, T., Douglas, J., and Hornung, U. Derivation of the double porosity model of single phase flow via homogenization theory. SIAM J Math. Anal. 21 (1990), 823–836.
  • [7] Beneš, M., and Pažanin, I. Homogenization of degenerate coupled transport processes in porous media with memory terms. Math Meth Appl Sci. 22 (2019), 6227–6258.
  • [8] Benilan, P., and Wittbold, P. On mild and weak solutions of elliptic–parabolic problems. Adv. Differ. Equ. 1 (1996), 1053–1073.
  • [9] Bittner, S., Janott, M., Ritter, D., Köcher, P., Beese, F., and Priesack, E. Functional–structural water flow model reveals differences between diffuse-and ring-porous tree species. Agricult Forest Meteorol 158 (2012), 80–89.
  • [10] Bourgeat, A., Luckhaus, S., and Mikelić, A. Convergence of the homogenization process for a double-porosity model of immiscible two-phase flow. SIAM J Math. Anal. 27 (1996), 1520–1543.
  • [11] Cioranescu, D., Damlamian, A., and Griso, G. The periodic unfolding method: Theory and Applications to Partial Differential Problems. Springer, 2018.
  • [12] Cioranescu, D., and Paulin, J. S. J. Homogenization of reticulated structures. Springer, 2012.
  • [13] D’Apice, C., Maio, U. D., and Mel’nyk, T. Asymptotic analysis of a perturbed parabolic problem in a thick junction of type 3:2:2. Networks Hetergo. Media 2 (2007), 255–277.
  • [14] DiBenedetto, E., and Gariepy, R. Local behavior of solutions of an elliptic–parabolic equation. Arch. Ration. Mech. Anal. 97 (1987), 1–17.
  • [15] El-Hames, A., and Richards, K. An integrated, physically based model for arid region flash flood prediction capable of simulating dynamic transmission loss. Hydrolog Processes 12, 8 (1998), 1219–1232.
  • [16] Endrizzi, S., Gruber, S., Dall’Amico, M., and Rigon, R. GEOtop 2.0: simulating the combined energy and water balance at and below the land surface accounting for soil freezing, snow cover and terrain effects. Geoscie Model Develop 7, 6 (2014), 2831–2857.
  • [17] Evans, L. C. Partial Differential Equations. AMS, Providence, RI, 2010.
  • [18] Frensch, J., and Steudle, E. Axial and radial hydraulic resistance to roots of maize (zea mays l.). Plant Physiology 91, 2 (1989), 719–726.
  • [19] Gaudiello, A., Guibé, O., and Murat, F. Homogenization of the brush problem with a source term in L1L^{1}. Arch. Rational Mech. Anal. 225 (2017), 1–64.
  • [20] Gerke, H. H., and Van Genuchten, M. T. A dual-porosity model for simulating the preferential movement of water and solutes in structured porous media. Water Resources Research 29, 2 (1993), 305–319.
  • [21] Hentschel, R., Bittner, S., Janott, M., Biernath, C., Holst, J., Ferrio, J. P., Gessler, A., and Priesack, E. Simulation of stand transpiration based on a xylem water flow model for individual trees. Agricultural and Forest Meteorology 182 (2013), 31–42.
  • [22] Holzworth, D. P., Huth, N. I., deVoil, P. G., Zurcher, E. J., Herrmann, N. I., McLean, G., Chenu, K., van Oosterom, E. J., Snow, V., Murphy, C., et al. Apsim–evolution towards a new generation of agricultural systems simulation. Environmental Modelling & Software 62 (2014), 327–350.
  • [23] Jäger, W., and Kačur, J. Solution of doubly nonlinear and degenerate parabolic problems by relaxation schemes. RAIRO Modél. Math. Anal. Numér. 29, 5 (1995), 605–627.
  • [24] Jäger, W., and Woukeng, J. Homogenization of richards’ equations in multiscale porous media with soft inclusions. J Diff Equat 281 (2021), 503–549.
  • [25] Janott, M., Gayler, S., Gessler, A., Javaux, M., Klier, C., and Priesack, E. A one-dimensional model of water flow in soil-plant systems based on plant architecture. Plant Soil 341, 1 (2011), 233–256.
  • [26] Kačur, J. Method of Rothe in Evolution Equations. In Proceedings Equadiff 6, 1985 (2006), Springer, pp. 23–34.
  • [27] Kačur, J., Malengier, B., and Keer, R. V. On the mathematical analysis and numerical approximation of a system of nonlinear parabolic PDEs. J Anal. Applic. 28 (2009), 305–332.
  • [28] Kim, I., and Požár, N. Nonlinear elliptic–parabolic problems. Arch. Ration. Mech. Anal. 210 (2013), 975–1020.
  • [29] Knabner, P., and Otto, F. Solute transport in porous media with equilibrium and nonequilibrium multiple-site adsorption: uniqueness of weak solutions. Nonlinear Analysis 42 (2000), 381–403.
  • [30] Kruzkov, S. First order quasilinear equations in several independent variables. Math. USSR Sbornik 10 (1970), 217–243.
  • [31] List, F., Kumar, K., Pop, I., and Radu, F. Rigorous upscaling of unsaturated flow in fractured porous media. SIAM J. Math. Anal. 52 (2020), 239–276.
  • [32] Mair, A., Dupuy, L., and Ptashnyk, M. Can root systems redistribute soil water to mitigate the effects of drought? Field Crops Research 300 (2023), 109006.
  • [33] Merz, W., and Rybka, P. Strong solutions to the richards equation in the unsaturated zone. J. Math. Anal. Appl. 371 (2010), 741–749.
  • [34] Mikelić, A., and Rosier, C. Modeling solute transport through unsaturated porous media using homogenization I. Comput Appl Math 23 (2004), 195–211.
  • [35] Nandakumaran, A., and Rajesh, M. Homogenization of a parabolic equation in perforated domain with dirichlet boundary condition. Proc. Indian Acad. Sci. (Math. Sci.) 122 (2002), 425–439.
  • [36] Neuss-Radu, M. Some extensions of two-scale convergence. C.R. Acad Sci Paris 332 (1996), 899–904.
  • [37] Nguetseng, G. A general convergence results for a functional related to the theory of homogenization. SIAM J Math. Anal. 20 (1989), 608–623.
  • [38] Otto, F. L1-contraction and uniqueness for quasilinear elliptic–parabolic equations. J Diff Eq 131, 1 (1996), 20–38.
  • [39] Ptashnyk, M. Degenerate quasilinear pseudoparabolic equations with memory terms and variational inequalities. Nonlin Anal, Theory, Methods, Appl 66, 12 (2006), 2653–2675.
  • [40] Schweizer, B. Regularization of outflow problems in unsaturated porous media with dry regions. J. Differential Equations 237, 2 (2007), 278–306.
  • [41] Simon, J. Compact sets in the space Lp(0,T;B){L}^{p}(0,{T};{B}). Ann Mat Pura Appl 146 (1986), 65–96.
  • [42] Simunek, J., van Genuchten, M. T., and Sejna, M. The HYDRUS-1D software package for simulating the movement of water, heat, and multiple solutes in variably saturated media, version 3.0, HYDRUS software series 1. Department of Environmental Sciences, University of California Riverside, Riverside (2005).
  • [43] Van Genuchten, M. T. A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil Scie Soc America J 44, 5 (1980), 892–898.
  • [44] Vanderborght, J., Couvreur, V., Meunier, F., Schnepf, A., Vereecken, H., Bouda, M., and Javaux, M. From hydraulic root architecture models to macroscopic representations of root hydraulics in soil water flow and land surface models. Hydrol Earth System Scie 25, 9 (2021), 4835–4860.
  • [45] Zarebanadkouki, M., Meunier, F., Couvreur, V., Cesar, J., Javaux, M., and Carminati, A. Estimation of the hydraulic conductivities of lupine roots by inverse modelling of high-resolution measurements of root water uptake. Annals of botany 118, 4 (2016), 853–864.