\documentclass[reqno]{amsart} \usepackage{epic} % For drawing figures \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2003(2003), No. 44, pp. 1--27.\newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu (login: ftp)} \thanks{\copyright 2003 Southwest Texas State University.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE--2001/44\hfil Interface separating fresh and salt groundwater] {On the behavior of the interface separating fresh and salt groundwater in a heterogeneous coastal aquifer} \author[S. Challal \& A. Lyaghfouri\hfil EJDE--2001/44\hfilneg] {Samia Challal \& Abdeslem Lyaghfouri} \address{Samia Challal \hfill\break King Fahd University of Petroleum and Minerals\\ P. O. Box 728, Dhahran 31261, Saudi Arabia} \address{Abdeslem Lyaghfouri \hfill\break King Fahd University of Petroleum and Minerals\\ P. O. Box 728, Dhahran 31261, Saudi Arabia} \email{lyaghfo@kfupm.edu.sa} \date{} \thanks{Submitted December 17, 2002. Published April 17, 2003.} \subjclass[2000]{35Q35, 35R35, 76S05} \keywords{Fresh-salt water, heterogeneous aquifer, nonlinear Darcy's law, \hfill\break\indent monotone solution, comparison and uniqueness, continuity of the free boundary, limit behavior} \begin{abstract} We consider a flow of fresh and salt groundwater in a two-dimensional heterogeneous horizontal aquifer. Assuming the flow governed by a nonlinear Darcy law and the permeability depending only on the vertical coordinate, we show the existence of a unique monotone solution that increases (resp. decreases) with respect to the salt (resp. fresh) water discharge. For this solution we prove that the free boundary is represented by the graph $x=g(z)$ of a continuous function. Finally we prove a limit behavior at the end points of the interval of definition of $g$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{definition}{Definition} \newtheorem{lemma}{Lemma} \newtheorem{proposition}{Proposition} \newtheorem{corollary}{Corollary} \newtheorem{remark}{Remark} \allowdisplaybreaks \section*{Introduction}\label{S:intro} Fresh water and sea water are actually miscible fluids and therefore the zone of contact between them takes the form of a transition zone caused by hydrodynamics dispersion. Across this zone the density of the mixed water varies from that of fresh water to that of sea water. Under certain conditions the width of this zone is relatively small (when compared with the thickness of the aquifer) so that we assume that each liquid is confined to a well defined portion of the flow domain with an abrupt interface separating the two domains. We consider here a two-dimensional model for fresh-salt water in a horizontally heterogeneous extended aquifer in the $xz$-plane. We suppose that the scale of the problem is sufficiently large so that the abrupt interface approximation is applicable. Moreover we consider the model of a flow obeying to a nonlinear Darcy law. In section 1, we indicate briefly how to obtain the weak formulation and the existence and some properties of the solutions, the definition of the free boundary $\Gamma=[x=g(z)]$ and the continuity of $g$ on its interval of definition $(-h^*,0) $, $h^*>0$. All these results generalize previous works in \cite{CL1} and \cite{CCL} (resp. \cite{CL2}) where the aquifer was supposed homogeneous (resp. heterogeneous) and the flow governed by a nonlinear (resp. linear) Darcy law. The aim of this paper is to study the behavior of the free boundary when $z\to -h^*$ and $z\to 0$ under the assumption that the permeability of the porous medium depends only on $z$ including the case of horizontal layers. Actually we establish in section 3 that $\lim_{z\to -h^*}g(z)=+\infty$ by generalizing the proof given in \cite{CL2}. We also give a second simple proof which works only for a constant permeability. Then we prove that $\lim_{z\to 0}g(z)=g(0-)$ exists. We recall that in \cite{AD} the authors first proved (in the linear and homogeneous case) that $\liminf_{z\to 0}g(z)>-\infty$ by using blow up arguments. They also proved that $\limsup_{z\to 0}g(z)<0$ and used this result to prove the existence of the limit $g(0-)$. Our proof does not assume $\limsup_{z\to 0}g(z)<0$ and is valid in the general case. Moreover we prove that $g(0-)$ is finite in more general cases. Our proofs are systematically based on comparing the solution locally or globally with explicit functions satisfying similar equations. This method of comparison is developed in section 2 to show that the solution increases with respect to the salt water discharge $Q_s$ and decreases with respect to the fresh water discharge $Q_f$. The uniqueness of the solution is obtained as a corollary of this monotonicity result. We also deduce a limit behavior of the solution when $Q_f$ or $Q_s$ goes to zero. Also by a comparison argument, when the permeability is constant, we give a simple proof of the fact that the set filled by fresh water is star shaped with the origin and the free boundary is non increasing in the region $[x>0]$. These two last results were proved in the linear case in \cite{AD} by using blow up arguments. Our proofs based on monotonicity arguments are much simpler. \section{Description of the model}\label{1} In this paper we are interested with the study of a stationary flow of fresh and salt water in a heterogeneous coastal aquifer $\Omega=\mathbb{R}\times(-h,0),\, h>0$, with permeability $A(X)$, $X=(x,z)\in \mathbb{R}^2$. The velocity and the pressure of the fluid are related by the following nonlinear Darcy law \begin{equation}\label{1.1} {\bf v}=-\left(\langle A(\nabla p+\gamma e_{z}),\nabla p+\gamma e_{z} \rangle \right)^{\frac{r-2}{2}}A(\nabla p+\gamma e_{z}) \end{equation} with $r>1$, $e_z=(0,1)$ and $\gamma$ is given by \begin{equation}\label{1.2} \gamma=\gamma_f\chi(\Omega_f)+\gamma_s\chi(\Omega_s) \quad\mbox{with}\quad 0<\gamma_f<\gamma_s \end{equation} where $\gamma_f$ (resp. $\gamma_s$) represents the specific weight of the fresh (resp. salt) water occupying the region $\Omega_f$ (resp. $\Omega_s$) of $\Omega$, $\chi(E)$ denotes the characteristic function of the set $E$. \begin{figure}[t] \begin{center} \setlength{\unitlength}{1mm} \begin{picture}(100,52)(0,0) \put(0,35){\line(1,0){40}} \put(55,35){\line(1,0){45}} \put(99,34.15){$\to$} \put(0,0){\line(1,0){100}} \put(40,35){\line(0,1){15}} \put(39.1,50){$\uparrow$} \put(35,50){$z$} \put(47,46){$Q_f$} \put(47,40){$\Downarrow$} \qbezier(20,35)(40,33)(60,25) \qbezier(60,25)(80,18)(100,15) \put(20,40){$S$} \put(55,40){$A$} \put(35,40){$O$} \put(100,37){$x$} \put(0,20){Salt water} \put(30,20){$\Omega_s$} \put(15,10){$Q_s \quad \Rightarrow$} \put(80,14){$\Gamma$} \put(80,27){Fresh water} \put(95,22){$\Omega_f$ } \put(43.5,1.5){$z=-h$} \end{picture} \end{center} \caption{} \end{figure} Fresh water is injected over the segment $[OA]$ ($A=(0,a)$ with $a>0$) uniformly (see Figure 1) with a total amount of $Q_f$. From infinity at the left of the aquifer, salt water arrives with a total discharge of $Q_s$. We assume that the two fluids are unmixed and separated by an interface $\Gamma$. The part of the boundary $\partial\Omega\setminus[OA]$ is assumed to be impervious and the flow incompressible. So the velocity satisfies \begin{equation}\label{1.3} \begin{gathered}\mathop{\rm div} {\bf v} =0\quad \text{in } \Omega,\quad {\bf v}=-\frac{Q_f}{a}e_z \quad\text{on } [OA],\\ {\bf v}\cdot\nu =0 \quad\text{on } \partial\Omega\setminus [OA],\quad {\bf v_i}\cdot\nu =0 \quad \text{on } \Gamma\quad(i=s,f), \end{gathered} \end{equation} where ${\bf v_i}$ is the restriction of ${\bf v}$ to $\Omega_i$ and $\nu$ is the outward unit normal to $\partial\Omega$ or $\Gamma$. We deduce from (1.3) that there exists a stream function $\psi$ satisfying \begin{equation}\label{1.4} \begin{gathered} {\bf v} =\mathop{\rm Rot}\psi=\Big(-\frac{\partial\psi} {\partial z},\frac{\partial\psi}{\partial x}\Big)\quad \text{in }\Omega,\quad\ \psi =0 \quad\text{on}\quad\Gamma,\\ \psi(x,-h) =Q_s \quad\text{and}\quad \psi(x,0)=\phi_0 (x)\quad\text{for } x\in\mathbb{R} \end{gathered} \end{equation} with \begin{equation}\label{1.5} \phi_0(x)=-Q_f\min\big(\frac{x^+}{a},1\big). \end{equation} We suppose the permeability matrix $A(X)$ such that \begin{equation}\label{1.6} \begin{gathered} A(X)\in\big[L^\infty(\Omega)\big]^4,\\ \exists m_0>0 :\; \langle A(X)\xi,\xi\rangle \geq m_0\vert\xi\vert^2,\; \forall\xi\in\mathbb{R}^2,\quad\text{a.e. } X\in\Omega,\\ ^tA=A.\end{gathered} \end{equation} Then there exists a unique symmetric and strictly elliptic matrix $\mathcal{A}$ satisfying $A(X)=\mathcal{A} ^t\mathcal{A}= \mathcal{A}^2$ (see \cite{Ci}). Then (1.1) becomes \begin{align*} {\bf v} &=-\langle \mathcal{A}(X)(\nabla p+\gamma e_{z}),\mathcal{A}(X) (\nabla p+\gamma e_{z} )\rangle ^{(r-2)/2} \mathcal{A}(X)\mathcal{A}(X)(\nabla p+\gamma e_{z})\\ &=-\vert \mathcal{A}(X)(\nabla p+\gamma e_{z})\vert^{r-2} \mathcal{A}(X)\mathcal{A}(X)(\nabla p+\gamma e_{z}). \end{align*} This leads to \[ \vert\mathcal{A}^{-1}(X){\bf v}\vert =\vert\mathcal{A}(X)(\nabla p+\gamma e_{z})\vert^{r-1} \] and \begin{equation}\label{1.7} \nabla p+\gamma e_{z}=-\vert\mathcal{A}^{-1}(X){\bf v}\vert^{\frac{2- r}{r-1}} \mathcal{A}^{-1}(X)(\mathcal{A}^{-1}(X){\bf v}). \end{equation} Now for $\zeta\in\mathcal{D}(\Omega)$ we get by \eqref{1.4} and \eqref{1.7} \[ \int_\Omega (\nabla p+\gamma e_z)\mathop{\rm Rot}\zeta=-\int_\Omega \vert\mathcal{A}^{-1}(X)\mathop{\rm Rot}\psi\vert^{\frac{2-r}{r-1}} \mathcal{A}^{-1}(X)(\mathcal{A}^{-1}(X)\mathop{\rm Rot}\psi)\cdot \mathop{\rm Rot}\zeta. \] If we set $\overline{A}(X)={\frac{1}{det A}A}$, then there exists a unique symmetric and strictly elliptic matrix $B(X)$ such that $\overline{A}(X)=B^2=B\cdot B$ and for which we have \[ ^t\mathcal{A}^{-1}(X)(\mathcal{A}^{-1}(X)\mathop{\rm Rot}\psi)\cdot \mathop{\rm Rot}\zeta=\langle\overline{A}(X) \nabla\psi,\nabla\zeta\rangle=B(X)\nabla\psi\cdot B(X)\nabla\zeta. \] Therefore \begin{equation}\label{1.8} \int_\Omega \vert B(X)\nabla\psi\vert^{\frac{2-r}{r-1}}B(X) \nabla\psi\cdot B(X)\nabla\zeta\ +\gamma e_x\nabla\zeta=0 \quad\forall \zeta\in \mathcal{D}(\Omega). \end{equation} Setting $q=r/(r-1)$, we deduce from \eqref{1.8} by taking $\zeta\in \mathcal{D} (\Omega_i)$ ($i=f,s$) \begin{equation}\label{1.9} \mathop{\rm div}\big(\mathcal{B}(X,\psi)\big)=0\quad\text{in }\mathcal{D}' (\Omega_i)\quad i=f,s \end{equation} with $$ \mathcal{B}(X,\xi)=\langle\overline{A}(X)\xi,\xi\rangle ^{\frac{q-2}{2}} \overline{A}(X) \xi. $$ Moreover if we assume that \begin{equation}\label{1.10} \lim_{x\to +\infty}\psi(x,z)\leq 0\quad\text{for }(x,z)\in \Omega_f\quad\text{and} \quad \lim_{x\to \pm\infty}\psi(x,z)\geq 0\quad\text{for }(x,z)\in \Omega_s \end{equation} and since \begin{equation}\label{1.11} \psi\leq 0\quad\text{on }\partial\Omega_f\quad\text{and} \quad\psi\geq 0\quad\text{on }\partial\Omega_s, \end{equation} we deduce by \eqref{1.9}-\eqref{1.11} and the maximum principle for $\mathcal{B}$-harmonic functions in unbounded domains (see \cite{DL}, \cite{K}) that \begin{equation}\label{1.12} \psi<0\quad\text{in }\Omega_f\quad\text{and}\quad\psi> 0\quad\text{in }\Omega_s. \end{equation} It follows from (1.2) and (1.12) that $\gamma\in H(\psi)$, where $H$ is the maximal monotone graph defined by \begin{equation*} H(t)=\gamma_f \chi([t<0])+[ \gamma_f ,\gamma_s]\chi([t=0]) +\gamma_s\chi([t>0]), \end{equation*} $[ \gamma_f ,\gamma_s]$ being the closed interval of $\mathbb{R}$ with endpoints $\gamma_f$ and $\gamma_s$. Then we are led to the following question:\\ \noindent\textbf{problem (P)} Find $(\psi,\gamma)\in W_{\rm loc}^{1,q}(\Omega)\times L^ {\infty}(\Omega)$ such that \begin{itemize} \item[(i)] $\int_\Omega\vert B(X)\nabla\psi\vert^{q-2}B(X) \nabla\psi\cdot B(X)\nabla\zeta+\gamma e_x\cdot\nabla\zeta=0$ for all $\zeta\in W_0^{1,q}(\Omega)$ with compact support in $\bar{\Omega}$ \item[(ii)] $\gamma\in H(\psi)$ a.e. in $\Omega$ \item[(iii)] $-Q_f\leq\psi\leq Q_s$ a.e. in $\Omega$ \item[(iv)] $\psi(x,-h)=Q_s$, $\psi(x,0)=\phi_0(x)$ for all $x\in\mathbb{R}$. \end{itemize} \smallskip Adapting technics in \cite{CCL}, \cite{CL1} and \cite{CL2} we prove the following theorems. \begin{theorem}\label{thm1.1} \begin{itemize} \item[(i)] There exists a solution $(\psi,\gamma)$ of $(P)$ that satisfies $\psi\in C_{\rm loc}^{0,\alpha}(\bar{\Omega})$ for some $\alpha\in(0,1)$. Moreover if $A(X)\in C_{\rm loc}^ {0,\sigma}(\Omega)$, then $\psi\in C_{\rm loc}^{1,\beta} (\Omega\setminus[\psi=0])$, $\sigma, \beta\in (0,1)$. \item[(ii)] If $B(X)=B(z)$, then there exists a monotone solution $(\psi,\gamma)$ of $(P)$ in the following sense: \begin{equation}\label{1.13} \partial_x\psi\leq 0\quad\text{and} \quad\partial_x\gamma\leq 0\quad\text{in }\mathcal{D}'(\Omega). \end{equation} \end{itemize} \end{theorem} For the rest of this article, we assume that \begin{equation}\label{1.14} B(X)=B(z)=\big(B_{ij}(z)\big)_{1\leq i,j\leq 2}\quad\hbox{a.e. in } \Omega \end{equation} and will consider only monotone solutions. Moreover we need to introduce the following two functions defined for $z\in[- h,0]$. \begin{equation}\label{1.15} v_{+\infty}(z)=-Q_f+(Q_s+Q_f)\phi_1(z), \quad v_{-\infty}(z)=Q_s\phi_1(z) \end{equation} where \[ \phi_1(z)=\frac{{\int_z^0\frac{ds}{(B_{12}^2+B_{22}^2)^ {q'/2}(s)}}} {{\int_{-h}^0\frac{ds}{(B_{12}^2+B_{22}^2)^{q'/2}(s)}}}. \] Then we have the following theorem. \begin{theorem}\label{thm1.2} Let $(\psi,\gamma)$ be a solution of (P) and let $\Omega_{m,n}=(m,n)\times(-h,0)$ for $m,n\in \mathbb{R}$. \begin{itemize} \item[(i)] For $r=\max(q,2)$, we have $\lim_{R\to \pm\infty}\int_{\Omega_{R,R+1}}\vert\nabla(\psi- v_{\pm\infty}\vert^r=0$. \item[(ii)] For all $z\in [-h,0]$, $\psi(x,z)\to v_{\pm\infty}(z)$ as $x\to\pm\infty$. \item[(iii)] $v_{+\infty}\leq\psi\leq v_{-\infty}$ in $\Omega$. \item[(iv)] $\gamma(x+R,z)\rightharpoonup \gamma_{\pm\infty}(z)$ in $L^{q'}(\Omega_{0,1})$ as $R\to\pm\infty$, where $\gamma_{\pm\infty}\in H(v_{\pm\infty})$. \end{itemize} \end{theorem} \begin{remark}\label{rmk1.1}\rm From $(iii)$, one can see that the strip $\mathbb{R} \times(-h,-h^*)$ is contained in $[\psi>0]$, where $h^*\in (0,h)$ is defined by $\phi_1(-h^*)=\frac{Q_f}{Q_s+Q_f}$. Therefore, the free boundary $\Gamma=[\psi=0]$ is contained in $\mathbb{R} \times[-h^*,0)$. \end{remark} Arguing as in \cite{CCL} and \cite{CL2} we can prove the continuity of the free boundary. The proof needs Lemma \ref{lmA.3} (see Appendix) which requires the following regularity of $B$: If $q\neq 2$, then \begin{equation} B(z)\in C^{0,1}_{\rm loc}(-h,0). \end{equation} \begin{remark}\label{rmk1.2} \rm Under assumption (1.16), the critical points of any $\mathcal{B}-$harmonic function in $\Omega$ are isolated (see \cite{AS}). For the rest of this article, we assume that (1.16) is satisfied. \end{remark} \begin{theorem}\label{thm1.3} There exists a continuous function $g:(-h^*,0)\rightarrow \mathbb{R}$ such that \[ [x=g(z)]\subset\Gamma\subset [x=g(z)]\cup[z=-h^*].\] \end{theorem} \begin{corollary}\label{coro1.1} \begin{itemize} \item[(i)] $\gamma=\gamma_s\chi([\psi>0])+\gamma_f\chi ([\psi<0])$ a.e. in $\Omega$. \item[(ii)] The sets $[\psi>0]$ and $[\psi<0]$ are connected by arcs. \end{itemize} \end{corollary} \section{Comparison and Uniqueness}\label{2} In this section we prove that solutions of $(P)$ increase with respect to $Q_s$ and decrease with respect to $Q_f$. As a consequence we obtain the uniqueness of the solution of $(P)$. Let us denote by $(P(Q_s,Q_f))$ the problem $(P)$ corresponding to $Q_s$ and $Q_f$. Then we have the following comparison result \begin{theorem}\label{thm2.1} Let $(\psi_1,\gamma_1)$ and $(\psi_2,\gamma_2)$ be solutions of Problems $(P(Q_{s_1},Q_{f_1}))$ and $(P(Q_{s_2},Q_ {f_2}))$ respectively. If $Q_{s_1}\leq Q_{s_2}$ and $Q_{f_2}\leq Q_ {f_1}$, then we have $\psi_1\leq \psi_2$ and $\gamma_1\leq\gamma_2$ a.e. in $\Omega$. \end{theorem} The proof of this theorem follows an idea in \cite{CCL} and uses a recent result due to Alessandrini and Sigalotti \cite {AS} regarding isolation of critical points of $\mathcal{B}$-harmonic functions in planar domains. First we prove the following lemma \begin{lemma}\label{lm2.1} Under the assumptions of Theorem \ref{thm2.1}, \[ T(\zeta)=\int_\Omega \Big(\mathcal{B}(z,\nabla\psi_1) -\mathcal{B}(z,\nabla\psi_0)+(\gamma_1-\gamma_0)e_x\Big).\nabla\zeta=0 \quad\forall\zeta\in \mathcal{D}(\mathbb{R}^2), \] where $\psi_0=\min(\psi_1,\psi_2)$ and $\gamma_0=\min (\gamma_1,\gamma_2)$. \end{lemma} \begin{proof} Let $\zeta\in \mathcal{D}(\mathbb{R}^2)$ and let $K=supp\zeta$, $M=\sup_{K}\vert\zeta\vert$. Then there exists $R_0>a$ such that: $\forall R\geq R_0$, $K\subset (- R,R)\times\mathbb{R}$. Consider $\zeta_R=M.\min\left(1,(-\vert x\vert+R+1)^+\right)$ and set $\zeta_1=\zeta+\zeta_R$, $\zeta_2=\zeta_R-\zeta$. Then for $\epsilon>0$ and $i=1,2$, $\min(\zeta_i,\frac{\psi_1-\psi_0}{\epsilon})$ is a test function and one has by the monotonicity of $\mathcal{B}(z,.)$ \begin{equation}\label{2.1} \begin{aligned} &\int_{[\psi_1-\psi_0\geq \epsilon\zeta_i]}\Big(\mathcal{B} (z,\nabla\psi_1)-\mathcal{B}(z,\nabla\psi_0)\Big) \nabla\zeta_i+\int_\Omega (\gamma_1-\gamma_0).\zeta_{ix}\\ &\leq \int_\Omega (\gamma_1-\gamma_0)\Bigl(\zeta_i-\frac {\psi_1-\psi_0}{\epsilon}\Bigl)^+_x. \end{aligned} \end{equation} Since $Q_{s_1}\leq Q_{s_2}$ and $Q_{f_2}\leq Q_{f_1}$ one has \[ -h_1^*=\phi_1^{-1}\big(\frac{Q_{f_1}}{Q_{s_1}+Q_{f_1}}\big) \leq \phi_1^{-1}\big(\frac{Q_{f_2}}{Q_{s_2}+Q_{f_2}}\big) =-h_2^* \] and then if we denote by $I=\{z\in(-h_2^*,0) : g_2 (z)a$ and ${00] $, we deduce that $\psi_1=-Q_{f_1}$ in $B_r(x_1,0)$. By the monotonicity of $\psi_1$ this leads to $\psi_1=-Q_{f_1}$ in the strip $[x_1,+\infty)\times (-r,0)$ which contradicts the asymptotic behavior of $\psi_1$ at $+\infty$. So there exists $x_1'\in (x_1-r,x_1+r)$ such that $\nabla\psi_1 (x_1',0)\ne 0$. Since $\psi_1\in C^1(D\cup(a,+\infty) \times\{0\})$ there exists $r'>0$ such that $\nabla\psi_1(x,0)\ne 0$ $\forall x\in (x_1'- r',x'_1+r')$. Therefore we get $\psi_0=\psi_1$ in $D$. In particular $\psi_0=\psi_1$ in $[\psi_0<0]$. Similarly one can prove that $\psi_0=\psi_1$ in $[\psi_0 >0]$ and then by continuity $\psi_0=\psi_1$ in $\Omega$. Using Corollary \ref{coro1.1} we deduce that $\gamma_0=\gamma_1$ in $\Omega$. \end{proof} As a direct consequence of Theorem \ref{thm2.1}, we have \begin{corollary}\label{coro2.1} The solution of problem $(P)$ is unique. \end{corollary} According to Theorem \ref{thm2.1} the solution of $(P)$ decreases with respect to $Q_f$ and increases with respect to $Q_s$. Intuitively one would expect that as $Q_f\to 0$ (resp. $Q_s\to 0)$ the aquifer would be saturated by salt (resp. fresh) water only. More precisely we have \begin{theorem}\label{thm2.2} Let $(\psi_{Q_f},\gamma_{Q_f})$ be the solution of $(P(Q_ {s},Q_{f}))$. We have \[ (\psi_{Q_f},\gamma_{Q_f})\to (v_{- \infty},\gamma_s)\quad\hbox{in } W_{\rm loc}^{1,q}(\Omega) \times L_{\rm loc}^{q'}(\Omega)\text{ as } Q_f\to 0, \] where $v_{- \infty}$ is given by (1.15). \end{theorem} \begin{proof} Using the monotonicity of $\psi_{Q_f}$ and $\gamma_{Q_f}$ with respect to $Q_f$ and the fact that the two functions are uniformly bounded, we deduce by Beppo-Levi's theorem that there exists two functions $\psi$, $\gamma$ such that \begin{gather*} \psi_{Q_f}\to \psi \quad\hbox{in }L_{\rm loc}^r (\Omega)\hbox{ a.e. in }\Omega\\ \gamma_{Q_f}\to \gamma \quad\hbox{in }L_{\rm loc} ^r(\Omega)\quad\forall r\geq 1. \end{gather*} Now let $m>a$ and $\eta\in W^{1,\infty}(\mathbb{R})$ such that $0\leq \eta\leq 1$, $\eta=1$ in $(-m,m)$, $\eta=0$ for $\vert x\vert\geq m+1$ and $\vert\eta'\vert\leq 1$. Let $\Phi(x,z)=\phi_0(x)+\phi_1(z)(Q_s-\phi_0(x))$. Then $\eta^q(\psi_{Q_f}-\Phi)$ is a test function for $(P(Q_ {s},Q_{f}))$ and we have \begin{align*} &\int_{\Omega_{m+1}}\eta^q \vert B(z)\nabla\psi_{Q_f} \vert^q\\ &=\int_{\Omega_{m+1}}\eta^q\vert B(z)\nabla\psi_{Q_f} \vert^{q-2}B(z)\nabla\psi_{Q_f}\cdot B(z)\nabla\Phi\\ &\quad-\int_{\Omega_{m+1}}q\eta^{q-1}(\psi_{Q_f}-\Phi)\vert B(z) \nabla\psi_{Q_f}\vert^{q-2}B(z)\nabla\psi_{Q_f}\cdot B(z) \nabla\eta\\ &\quad-\int_{\Omega_{m+1}}\gamma_{Q_f}\eta^q\partial_x\psi_{Q_f} +\int_{\Omega_{m+1}}\gamma_{Q_f}\eta^q\partial_x\Phi-\int_ {\Omega_{m+1}}\gamma_{Q_f}(\psi_{Q_f}-\Phi)q\eta^{q-1} \eta'. \end{align*} Since $\psi_{Q_f},\Phi, \nabla\Phi,\eta,\eta',\gamma_ {Q_f}$ are uniformly bounded, we deduce by using H\"{o}lder's inequality \[ \vert\psi_{Q_f}\vert_{1,q,\Omega_m}\leq c_m \] where $c_m$ is a constant depending on $m$. Then by using a diagonal process we get, up to a subsequence \[ \psi_{Q_f}\rightharpoonup \psi \quad\hbox{in }W_{\rm loc}^ {1,q}(\Omega). \] Let us show that this convergence holds strongly. Let $m>a$ and let $\rho\in\mathcal{D}(\Omega_m)$, $\rho\geq 0$ in $\Omega_m$. Then $\rho^q(\psi_{Q_f}- \psi)$ is a test function for $(P(Q_ {s},Q_{f}))$ and we have \begin{align*} &\int_{\Omega_{m}}\rho^q \vert B(z)\nabla\psi_{Q_f} \vert^q\\ &=\int_{\Omega_{m}}\rho^q\vert B(z)\nabla\psi_{Q_f} \vert^{q-2}B(z)\nabla\psi_{Q_f}\cdot B(z)\nabla\psi\\ &\quad-\int_{\Omega_{m}}q\rho^{q-1}(\psi_{Q_f}-\psi)\vert B(z) \nabla\psi_{Q_f}\vert^{q-2}B(z)\nabla\psi_{Q_f}\cdot B(z) \nabla\rho\\ &\quad-\int_{\Omega_{m}}\rho^q\gamma_{Q_f}\partial_x(\psi_ {Q_f}-\psi) -\int_{\Omega_{m}}\gamma_{Q_f}(\psi_{Q_f}-\psi)q\rho^{q-1} \partial_x\rho. \end{align*} Applying H\"{o}lder's inequality and letting $Q_f\to 0 $, we get \[ \limsup_{Q_f\to 0}\Big(\int_{\Omega_{m}}\rho^q \vert B (z)\nabla\psi_{Q_f}\vert^q\Big)^{1/q}\leq \Bigl(\int_{\Omega_{m}}\rho^q \vert B(z) \nabla\psi\vert^q\Bigl)^{1/q}. \] Hence $\rho^q B(z)\nabla\psi_{Q_f}\to \rho^q B(z) \nabla\psi$ in $L^q(\Omega_{m})$ and in particular \[ \nabla\psi_{Q_f}\to \nabla\psi\quad\hbox{in } W_{\rm loc}^{1,q}(\Omega). \] Now using the monotonicity with respect to $Q_f$ and the continuity of $\psi_{Q_f}$, it follows by Dini's theorem \[\lim_{Q_f\to 0}\big(\lim_{R\to \pm\infty}\psi_{Q_f}(x+R,z) \big)= \lim_{R\to \pm\infty}\big(\lim_{Q_f\to 0}\psi_{Q_f}(x+R,z) \big) \] and \[\lim_{R\to \pm\infty}\psi(x+R,z)=\lim_{Q_f\to 0}v_ {\pm\infty}^{Q_f}(z)=v_{-\infty}(z)\quad\hbox{ for }(x,z) \in\Omega_{0,1}, \] where $v_{-\infty}^{Q_f}$ and $v_{+\infty}^{Q_f}$ were defined by (1.15). Since $\partial_x\psi\leq 0$, we deduce that $\psi(x,z)=v_{- \infty}(z)=Q_s\phi_1(z)$. Moreover $\gamma\in H(\psi)$ and $\psi>0$ in $\Omega$, so $\gamma=\gamma_s$ a.e. in $\Omega$. We conclude that when $Q_f\to 0$, the aquifer is completely saturated by salt water only. \end{proof} \begin{theorem}\label{thm2.3} Let $(\psi_{Q_s},\gamma_{Q_s})$ be the solution of $(P(Q_ {s},Q_{f}))$. We have \[(\psi_{Q_s},\gamma_{Q_s})\to (\psi,\gamma_f) \quad\hbox{in } W_{\rm loc}^{1,q}(\Omega)\times L_{\rm loc}^ {q'}(\Omega)\text{ as } Q_s\to 0, \] where $\psi$ is the solution of the following problem: \end{theorem} \noindent\textbf{Problem($P(Q_f)$):} \begin{itemize} \item[(i)] $\mathop{\rm div}\big(\mathcal{B}(z,\nabla\psi)\big)=0$ in $\mathcal{D}'(\Omega)$ \item[(ii)] $-Q_f<\psi<0$ in $\Omega$ \item[(iii)] $\psi(x,-h)=0$, $\psi(x,0)=\phi_0(x)$ for all $x\in \mathbb{R}$ \item[(iv)] $\lim_{x\to -\infty}\psi(x,z)=0$, $\lim_{x\to +\infty}\psi(x,z)=Q_f(\phi_1(z)-1)$. \end{itemize} \begin{proof}[Proof of Theorem \ref{thm2.3}] Taking into account the monotonicity of $\psi_ {Q_s}$ and $\gamma_{Q_s}$ with respect to $Q_s$ and arguing as in the proof of Theorem \ref{thm2.2}, we deduce the existence of two functions $\bar{\psi}$ and $\bar{\gamma}$ such that $$ \psi_{Q_s}\to \bar{\psi}\quad\mbox{in }W_{\rm loc}^{1,q}(\Omega) \quad \mbox{and}\quad \gamma_{Q_s}\to \bar{\gamma}\quad\mbox{in }L_{\rm loc}^{q'}(\Omega). $$ It follows that $(\bar{\psi},\bar{\gamma})$ satisfies: \begin{itemize} \item[(i)] $\mathop{\rm div}\big(\mathcal{B}(z,\nabla\bar{\psi})\big)=- \partial_x\bar{\gamma}$ in $\mathcal{D}'(\Omega)$ \item[(ii)] $\bar{\gamma}\in H(\bar{\psi})$ \item[(iii)] $-Q_f\leq\bar{\psi}\leq 0$ in $\Omega$ \item[(iv)] $\bar{\psi}(x,-h)=0$, $\bar{\psi}(x,0)=\phi_0(x)$ for all $x\in \mathbb{R}$ \item[(v)] $\lim_{x\to -\infty}\bar{\psi}(x,z)=0$, $\lim_{x\to +\infty}\bar{\psi}(x,z)=Q_f(\phi_1(z)-1)$ \item[(vi)] $\partial_x\bar{\psi}\leq 0$ and $\partial_x\bar{\gamma}\leq 0$ in $\mathcal{D}'(\Omega)$. \end{itemize} By the weak maximum principle, we can compare $\bar{\psi}$ with the solution $\psi$ of $P(Q_f)$. This gives $\bar{\psi} \leq \psi$ in $\Omega$. Since $\psi$ satisfies by the strong maximum principle $-Q_f<\psi<0$, we deduce that $\bar {\psi}<0$ in $\Omega$ and then $\bar{\gamma}=\gamma_f$ a.e in $\Omega$. Consequently $\bar{\psi}= \psi$ in $\Omega$. We conclude that when $Q_s\to 0$, the aquifer is completely saturated by fresh water only.\end{proof} The end of this section is devoted to study the set $\Omega_f=[\psi<0]$. We would like to point out that it was proved in \cite{AD} in the linear case that this set is star shaped with the origin. The proof was based on blow-up arguments. Here we propose a different proof based on a comparison argument that works for the linear case as well as for the nonlinear one with constant permeability. So we assume that $\mathcal{B}$ does not depend on $z$. For any $r>0$, we consider \[ \psi_r(x,z)=\frac{1}{r}\psi(rx,rz)\quad\text{and}\quad \gamma_r(x,z)=\frac{1}{r}\gamma(rx,rz) \] defined on $\Omega_r=\mathbb{R}\times(\frac{-h}{r},0)$. It is easy to check that $(\psi_r,\gamma_r)$ is the solution of the following problem: \noindent\textbf{Problem $(P_r)$} Find $(\psi_r,\gamma_r)\in W_{\rm loc}^{1,q}(\Omega_r)\times L^{\infty}(\Omega_r)$ such that: \begin{itemize} \item[(i)] ${\int}_{\Omega_r}\vert B\nabla\psi_r\vert^{q-2}B \nabla\psi_r\cdot B\nabla\zeta+\gamma_r e_x.\nabla\zeta=0$ for all $\zeta\in W_0^{1,q}(\Omega_r)$ with compact support in $\overline\Omega_r$ \item[(ii)] $\gamma_r\in H(\psi_r)$ a.e. in $\Omega_r$ \item[(iii)] $\psi_r(x,\frac{-h}{r})=\frac{Q_s}{r}$, $\psi_r(x,0)=\frac{1}{r}\phi_0(rx)$ for all $x\in \mathbb{R}$ \item[(iv)] $\partial_x\psi_r\leq 0$ and $\partial_x\gamma_r\leq 0$ in $\mathcal{D}'(\Omega_r)$. \end{itemize} >From the study of problem $(P)$, we know that problem $(P_r) $ has a unique solution with a continuous free boundary $g_r$ and that $\lim_{x\to-\infty}\psi_r(x,z)={\frac{1}{r}v_{-\infty}(rz)}$, $\lim_{x\to+\infty}\psi_r(x,z)={\frac{1}{r} v_{+\infty}(rz)}$. Moreover since we assume that $\mathcal{B}$ does not depend on $z$, we have $v_{-\infty}(z)=-{\frac{Q_s} {h}z}$ and $v_{+\infty}(z)=-{\frac{Q_f}{r}}-{\frac{Q_f+Q_s}{h}z}$. \begin{theorem}\label{thm2.4} For $00]$ it is enough to verify that $\nabla\psi_{r_1}$ does not vanish on some part of the left hand side of the line ${[z=-\frac{h} {r_2}]}$. So assume that for some $x_0$, we have ${\nabla\psi_ {r_1}(x,-h/r_2)=0}$ for all $x\leq x_0$. Then ${\psi_{r_1}(x_0,-h/r_2)}=C\in\mathbb{R}$ for all $x\leq x_0$. By the asymptotic behavior $C=Q_s/r_2>0$ and by continuity and monotonicity, $\psi_{r_1}$ is positive in a strip ${D_\epsilon=(-\infty,x_0)\times(-\frac {h}{r_2}-\epsilon,-\frac{h}{r_2}+\epsilon)}$ for some small $\epsilon>0$. Therefore $\psi_{r_1}$ is $\mathcal{B}$-Harmonic in $D_\epsilon$ and its gradient has non-isolated zeros. This means that $\psi_{r_1}=C$ in $D_\epsilon$ which contradicts the asymptotic behavior. \end{proof} \begin{corollary}\label{coro2.2} $\Omega_f$ is star shaped with the origin. \end{corollary} \begin{proof} Let $X_0\in \Omega_f$ and $t\in (0,1]$. We have by Theorem \ref{thm2.4}, $\psi_t(X_0)\leq\psi_1(X_0)=\psi(X_0)<0 $. So $\psi(tX_0)<0$, which means that $tX_0\in \Omega_f$. \end{proof} \begin{corollary}\label{coro2.3} $(i)$ There exists $z_0\in(-h^*,0)$ such that $g(z)\geq 0$ for all $z\in (-h^*,z_0)$.\\ $(ii)$ $g$ is non-increasing where it is nonnegative. \end{corollary} \begin{proof} $(i)$ First note that the set $[g>0]$ is nonempty. Indeed if $g(z)\leq 0$ $\forall z\in (-h^*,0)$, then for all $(x,z)\in(0,+\infty)\times(-h^*,0)$ we have $\psi (x,z)<0$. Since we also have $\psi(x,z)>0$ in $(0,+\infty)\times(-h,-h^*)$, we get a contradiction with Lemma \ref{lmA.3}. Let $z_0=\inf\{\,z\in (-h^*,0)\,/\,g(z)>0\}$ and take $z\in (-h^*,z_0)$. Since ${r=\frac{z_0}{z}<1}$, we have \begin{align*} 0&=\psi(g(z_0),z_0)=\frac{1}{r}\psi\Big(r\Big(\frac{g(z_0)} {r}\Big),rz\Big)=\psi_r\Big(\frac{g(z_0)}{r},z\Big)\\ &\leq \psi_1\Big(\frac{g(z_0)}{r},z\Big)=\psi\Big(\frac{g (z_0)}{r},z\Big)\leq\psi(g(z_0),z) \end{align*} since $g(z_0)\geq 0$ and ${\frac{1}{r}>1}$. Thus $\psi(g (z_0),z)\geq 0$ and then $g(z)\geq g(z_0)\geq 0$. \noindent $(ii)$ Let $z_1, z_2\in (-h^*,z_0)$ such that $z_1L,\quad\forall z\in (-h^*,z_0), \quad\psi(x,z)<0. \] Since $\psi>0$ for $z<-h^*$ we deduce by continuity that we have necessarily $\psi(x,-h^*)=0$ for all $x>L$. This leads to a contradiction with Lemma \ref{lmA.3}. Thus $L=+\infty$. \noindent $(i)$ Assume that $S\ne\emptyset$. Then there exists $x_0\in \mathbb{R}$ such that $\psi(x_0,-h^*)=0$. For $A_0>x_0$ there exists $\delta>0$ by $(ii)$ such that: for all $z \in(-h^*,-h^*+\delta)$, $g(z)>A_0$. So for $(x,z)\in(-\infty,A_0)\times(-h^*,-h^*+\delta)$ we have $\psi(x,z)>0$. By monotonicity of $\psi$, we deduce that $\psi(x,-h^*)=0$ $\forall x\geq x_0$ since one has $v_{+\infty}(-h^*)=0\leq \psi(x,-h^*)\leq \psi (x_0,-h^*)=0$. Hence we have $\psi>0$ in $(x_0,A_0)\times\Big ((-h^*-\delta,-h^*+\delta)\setminus\{-h^*\}\Big)$ and $\psi(x,-h^*) =0$ $\forall x\in( x_0,A_0)$ which contradicts Lemma \ref{lmA.3}. \noindent\textbf{The general case:} Since we follow the proof given in \cite{CL2} for the linear case, we will only give an outline of it. \noindent $(i)$ Assume that $S\ne\emptyset$. Then $S=[\alpha,+\infty)$ with $\alpha=\inf S>-\infty$. We need some lemmas. \begin{lemma}\label{lm3.1} There exists $u\in L^\infty(-h,0)$ such that $u>0$ for a.e $z\in(-h,0) $ and \[ E\big(u(z)\big)=\left(\overline{A}_{22}u^2(z)-2\overline{A}_{12}u(z)+ \overline{A}_{11}\right)^{\frac{q-2}{2}} \left(\overline{A}_{22}u(z)-\overline{A}_{12}\right)-C_0=0 \] for some constant $C_0>-\big(\overline{A}_{11}^{\frac{q-2}{2}}\overline{A}_{12} \big)(z)$ for a.e $z\in(-h,0)$. $\overline{A}_{ij}$ being the entries of the matrix $\overline{A}$ introduced in Section 1. \end{lemma} \begin{proof} The function $E: \mathbb{R}\to \mathbb{R}$ is continuous and satisfies ${\lim_{t\to +\infty}E(t)=+\infty}$, $E(0)=-\overline{A}_{11}^{\frac{q-2}{2}}\overline{A}_{12}-C_0<0$. So we deduce that for a.e $z\in(-h,0)$, there exists $u(z)>0$ such that $E\big(u(z)\big)=0$. This clearly defines a positive and uniformly bounded function $u$ on $(-h,0)$. \end{proof} Set $\alpha'=\alpha+\frac{1}{2}$ and define $f(z)=\int_{-h}^z u(s)ds+\alpha'$. For $k>0$ define $v$ and $\theta$ by \begin{gather*} v(x,z)=\big(k(\gamma_s-\gamma_f)\big)^{1/(q-1)} (f(z)-x)^+\\ \theta(x,z)=\gamma_s\chi\big([xf(z)]\big) \end{gather*} for $(x,z)\in D(z_1)=(\alpha',+\infty)\times(- h^*,z_1)$ with $z_1\in (-h^*,0).$ Then we have the following lemma. \begin{lemma}\label{lm3.2} There exists $k>0$ such that \[ \int_{D(z_1)}\Big(\mathcal{B}(z,\nabla v)+\theta e_x\Big) \nabla\xi\geq 0\quad\forall \xi\in\mathcal{D}\big(D(z_1) \big). \] \end{lemma} To prove this lemma, one can adapt the proof of \cite[Lemma 6.3]{CL2}. \begin{lemma}\label{lm3.3} Let $(\psi,\gamma)$ be the solution of $(P)$. Then there exists $z_0\in (-h^*,0)$ such that \[ \int_{D(z_0)}\big(\mathcal{B}(z,\nabla \psi^+)-\mathcal{B} (z,\nabla v_0)+(\gamma-\theta_0) e_x\big)\nabla\zeta=0 \quad\forall \zeta\in\mathcal{D}\big(\mathbb{R}^2\big) \] where $v_0=\min(\psi^+,v)$ and $\theta_0=\min(\gamma,\theta) $. \end{lemma} To prove this lemma, one can adapt the proof in \cite[Lemma 6.4]{CL2}. \begin{proof}[completion of the proof for Theorem \ref{thm3.1}] Let $z_*\in (-h^*,z_0)$. \\ $\bullet$ If $\psi^+(\alpha',z_*)=0$, then by monotonicity we have $\psi^+(x,z_*)=0\,\,\forall x\geq \alpha'$. \\ $\bullet$ If $\psi^+(\alpha',z_*)>0$, then since we can choose $k$ such that $\psi^+(\alpha',z_*)0$ and $\psi^+< v$. Then $\psi^+= v_0$ in $B_r(\alpha',z_*) $. Let us denote by $C_*$ the connected component of $D(z_0) \cap[\psi^+>0]$ which contains $B_r(\alpha',z_*)\cap D(z_0)$. Let $D^+(z_0)=D(z_0)\cap [x0$, $v>0$, it follows that $\gamma=\theta=\theta_0=\gamma_s$ and then \[ \mathop{\rm div}\big(\mathcal{B}(z,\nabla \psi^+)\big)= \mathop{\rm div}\big(\mathcal{B}(z,\nabla v_0)\big)=0\quad\hbox{in } \mathcal{D}'\big(C_*\cap D^+(z_0)\big). \] So we have \begin{gather*} \mathop{\rm div}\big(\mathcal{B}(z,\nabla \psi^+)\big)=0,\quad \mathop{\rm div}\big (\mathcal{B}(z,\nabla v_0)=0\quad\hbox{in } \mathcal{D}'\big(C_*\cap D^+(z_0)\big)\\ \psi^+\geq v_0 \quad \hbox{in }C_*\cap D^+(z_0)\\ \quad\psi^+= v_0 \quad \hbox{in }B_r(\alpha',z_*) \cap D^+(z_0) \end{gather*} which leads by Lemma \ref{lmA.1} (see Appendix) to $\psi^+= v_0$ in $C_*\cap D^+(z_0)$. We then conclude as in the end of the proof of Theorem 6.1 in \cite{CL2}. \noindent $(ii)$ See Corollary 6.6 of \cite{CL2}. \end{proof} \section{Behavior of the free boundary near $z=0$ }\label{4} Now we study the free boundary near $z=0$. We first prove the existence of a limit $g(0-)\leq 0$ as $z$ approaches zero. Then we prove that this limit is finite in two cases. \subsection{Existence of the limit of $g$ at $z=0$}\label{4.1} \begin{theorem} \label{thm4.1} The function $g$ has a limit when $z$ approaches zero. \end{theorem} \begin{proof} It suffices to show that ${\liminf_{z\to 0}g(z) =\limsup_{z\to 0}g(z)}$.\\ First, if ${\limsup_{z\to 0}g(z)=-\infty}$, then $$\liminf_ {z\to 0}g(z)=\limsup_{z\to 0}g(z)=\lim_{z\to 0}g(z)=-\infty. $$ Now assume that $\limsup_{z\to 0}g(z)=a^+>-\infty$. Note that $a^+\leq 0$. Indeed if $a^+>0$, then $\psi(a^+,0)<0$ and by continuity of $\psi$ there exists $\epsilon>0$ such that $\psi<0$ in $(a^+-\epsilon,a^++\epsilon)\times(-\epsilon,0) $. So $g(z)\leq a^+-\epsilon\quad\forall z\in(-\epsilon,0)$ and then $a^+\leq a^+-\epsilon$ which is impossible. Set $a^-=\liminf_{z\to 0}g(z)$ and assume that $a^-0$ we define $v$ and $\theta$ by \begin{equation} \begin{gathered} v(x,z)=\big(k(\gamma_s-\gamma_f)\big)^{\frac{1}{q-1}}(f(z)-x)^+\\ \theta(x,z)=\gamma_s\chi\big([xf(z)]\big) \end{gathered}\end{equation} for $(x,z)\in D(z_n)=(x_1,+\infty)\times(z_n,0)$ with $n\geq n_2$. Then as in Lemma \ref{lm3.2}, one can deduce from (4.2) and(4.4) the existence of $k>0$ such that \[ \int_{D(z_n)}\Big(\mathcal{B}(z,\nabla v)+\theta e_x\Big) \nabla\xi\geq 0\quad\forall \xi\in\mathcal{D}\big(D(z_n) \big). \] Using the fact that $\psi^+(x_1,0)=0$, the continuity of $\psi$, the monotonicity of $f$, (4.1) and (4.3), there exists $n\geq\sup(n_1,n_2)$ such that $\psi^+(x_1,z)\leq v (x_1,z)\quad\forall z\in(z_n,0)$. Moreover one can check that \[ \psi^+\leq v\quad\hbox{on } \partial D(z_n)\quad\hbox {and }\quad \lim_{x\to+\infty}\psi^+(x,z)=\lim_ {x\to+\infty}v(x,z)=0. \] Arguing as in the proof of Theorem \ref{thm3.1}, we obtain $\psi^+\leq v$ in $D(z_n)$ from which we deduce that $\psi^+ (x,z)=0\quad\forall (x,z)\in(x_2,a^+)\times(z_n,0)$ and then $g(z) \leq x_2\quad\forall z\in(z_n,0)$. So ${a^+=\limsup_{z\to 0}g (z)\leq x_2-\infty.$$ \end{theorem} The idea of the proof is to compare $\psi$ with a suitable function. First let $R, b>0$ and consider \begin{equation*} f(x)= \begin{cases} -R+\sqrt{R^2-(x+b)^{+2}} &\hbox{if }x\leq {x_R=\frac{R(a+b)} {\sqrt{R^2+(a+b)^2}}-b},\\ z_R-{\frac{a+b}{R}}(x-x_R) &\hbox{if }x\geq x_R \end{cases} \end{equation*} with ${z_R=\frac{R(a+b)}{\sqrt{R^2+(a+b)^2}}-R}$ (see Figure 2). Set \[\omega(z)=h\frac{{\int}_0^z\frac{ds}{B_{22}(s)}}{ {\int}_{-h}^0 \frac{ds}{B_{22}(s)}}={\kappa\int_0^z\frac{ds}{B_{22}(s)}} \quad\hbox{ for }z\in(-h,0). \] For $t>0$, consider $$G(t)=\lambda[(t+1)^2-1]\quad \text{with}\quad \lambda=\frac{Q_s}{h^2+2h}$$ and define $v_1$ by \[ v_1(x,z)=G(f(x)-\omega(z))\quad\hbox{for }(x,z)\in D_1=\{(x,z)\in \Omega/\,-h0$, set $f_d(x)=f(x)+d$ and for $t>0$ consider $$ K(t) =\mu \log(1+t)\quad \text{with}\quad \mu=\frac{Q_f}{\log(1+d)}. $$ Then we define $v_2$ by \[ v_2(x,z)=-K(\omega(z)-f(x))\quad\hbox{for } (x,z)\in D_2 \] with $D_2=\{(x,z)\in \Omega :\omega^{-1}(x) 0$ such that \begin{equation} \forall\alpha\in(0,\alpha_1),\;\forall R\geq \max\Big({\frac{2M^2}{\kappa^2},\frac{a}{\alpha}}\Big),\quad \mathop{\rm div}\big(B(z)\nabla v_1\big)\geq 0\quad\hbox{ in }\mathcal{D}' (D_1). \end{equation} \end{lemma} \begin{proof} Indeed let $C_1$ be a constant such that $B_{21}(z)+C_1>0$ for a.e $z\in(-h,0)$. Then we have \begin{equation} \begin{aligned} &\mathop{\rm div}\big(B(z)\nabla v_1\big)\\ &=\frac{\partial}{\partial x}\big(B_{11}\frac{\partial v_1}{\partial x} + B_{12}\frac {\partial v_1}{\partial z}\big) +\frac{\partial}{\partial z}\big(-C_1\frac{\partial v_1}{\partial x} + B_{22}\frac{\partial v_1}{\partial z}\big) +\frac{\partial}{\partial z}\big((B_{21}+C_1)\frac {\partial v_1}{\partial x}\big)\\ &=B_{11}\big(G''(f(x)-\omega(z))(f')^2+G'(f(x)-\omega(z))f''\big)\\ &\quad+(B_{12}-C_1)\big(\frac{-\kappa}{B_{22}}f' G''(f(x)-\omega(z))\big) +\frac{\kappa^2}{B_{22}}G''(f(x)-\omega(z)) \\ &\quad -\frac{\kappa}{B_{22}}(B_{21}+C_1)G''(f(x)- \omega(z))f'+B_{21}' G'(f(x)-\omega(z))f'\\ &\geq 2\lambda B_{11}\big((f')^2+\frac{\kappa^2}{B_ {22}B_{11}}-\frac{\kappa}{B_{22}B_{11}}(B_{12}-C_1)f' +(f(x)-\omega(z)+1)f''\big)\\ &=2\lambda B_{11}I_1 \end{aligned} \end{equation} since $B_{21}+C_1>0,\,G''=2\lambda,\,f' (x)\leq 0,\,B_{21}'\leq 0\hbox{ in } \mathcal{D}'(- h,0)$ and $G'(t)=2\lambda (t+1)>0$ for $t>0$. We shall distinguish three cases: \begin{itemize} \item For $x\leq -b$, we have $f(x)=0$ and then ${I_1=\frac{\kappa^2}{B_{22}B_{11}}>0}$. \item For $x\in(x_R,f^{-1}(-h))$, we have $f''(x)=0$, ${f'(x)=-\frac{a+b}{R}=- \alpha}$, $m\leq B_{11},B_{22}\leq M$, $\vert B_{12}\vert\leq M$, where $m$ and $M$ are two positive constants such that \begin{equation} m\vert\xi\vert^2\leq \langle B(z) \xi,\xi\rangle \leq M\vert\xi\vert^2\quad\forall\xi \in\mathbb{R}^2,\text{ a.e. } z\in(-h,0). \end{equation} So \begin{align*} I_1 &=\alpha^2+\frac{\kappa^2}{B_{22}B_{11}} +\frac{\alpha\kappa}{B_{22}B_{11}}(B_{12}-C_1)\\ &\geq \alpha^2+\frac{\kappa^2}{M^2}-\frac{\alpha\kappa}{m^2} (M+C_1)\to \frac{\kappa^2}{M^2}>0\quad\hbox{as } \alpha\to 0. \end{align*} Therefore, there exists $\alpha_0>0$ such that $\forall\alpha\in(0,\alpha_0)\quad I_1\geq 0$. \item For $x\in (-b,x_R)$, we have \begin{gather*} f'(x)=\frac{-(x+b)}{\sqrt{R^2-(x+b)^2}}, \quad f''(x)=-\frac{R^2}{(R^2-(x+b)^2)^{3/2}},\\ 1+{f'}^2(x)=\frac{R^2}{R^2-(x+b)^2}. \end{gather*} Then \begin{align*} I_1 &=\frac{R^2}{(R^2-(x+b)^2)^{3/2}}\Big(\sqrt{R^2-(x+b)^2}\\ &-f(x)+\frac{(R^2-(x+b)^2)^{3/2}}{R^2}\big(\frac{\kappa^2} {B_{22}B_{11}}-1\big)\\ &+\frac{\kappa(B_{12}-C_1)}{B_{22}B_{11}} \frac{(R^2-(x+b)^2)}{R^2}(x+b)+\omega(z)-1\Big). \end{align*} \end{itemize} Note that \begin{itemize} \item $\sqrt{R^2-(x+b)^2}-f(x)=R$. \item ${\frac{(R^2-(x+b)^2)^{3/2}}{R^2}\left(\frac{\kappa^2}{B_{22} B_{11}}-1\right)\geq R\left(\frac{\kappa^2}{M^2}-1\right)}$ provided that $M>\kappa$. \item ${\frac{\kappa(B_{12}-C_1)}{B_{22}B_{11}}\frac{(R^2-(x+b) ^2)}{R^2}(x+b)\geq-C_2\frac{\alpha R}{\sqrt{1+\alpha^2}} \geq-C_2(x_R+b)}$ since \\ ${\frac{\kappa(B_{12}-C_1)}{B_{22}B_{11}}>-C_2}$ for some positive constant $C_2$. \item $\omega(z)\geq -h$. \end{itemize} Therefore \begin{align*} I_1 &\geq\frac{R^2}{(R^2-(x+b)^2)^{3/2}}\big(R\frac {\kappa^2}{M^2}-\frac{C_2\alpha }{\sqrt{1+\alpha^2}}R-h-1\big)\\ &\geq\frac{R^3}{(R^2-(x+b)^2)^{3/2}}\Big(\big(\frac {\kappa^2}{2M^2}-C_2\alpha \big)+\big(\frac{\kappa^2}{2M^2}- \frac{h+1}{R}\big)\Big) &\geq 0 \end{align*} provided that $\alpha\leq \frac{\kappa^2}{2M^2 C_2}$ and $R>\frac{2M^2}{\kappa^2}(h+1)$. Finally for ${\alpha\in\big(0,\alpha_1=\min\big(\alpha_0,\frac{\kappa^2}{2M^2C_2}\big)\big)}$ and ${R>\frac {2M^2}{\kappa^2}(h+1)}$, one has $I_1\geq 0$ and then by (4.6) we obtain (4.5). \end{proof} \begin{lemma}\label{lm4.2} There exists $\alpha_2>0$ such that \begin{equation} \forall\alpha\in(0,\alpha_2),\; \forall R\geq \max\big({\frac{2M^2}{\kappa^2},\frac{a}{\alpha}}\big),\quad \mathop{\rm div}\left(B(z)\nabla v_2\right)\geq 0\quad\hbox{in } \mathcal{D}'(D_2). \end{equation} \end{lemma} \begin{proof} Indeed we have \begin{equation} \begin{aligned} &\mathop{\rm div}\left(B(z)\nabla v_2\right)\\ &=\frac{\partial}{\partial x} \big(B_{11}\frac{\partial v_2}{\partial x}+ B_{12}\frac {\partial v_2}{\partial z}\big)+\frac{\partial}{\partial z} \big(-C_1\frac{\partial v_2}{\partial x}+ B_{22}\frac {\partial v_2}{\partial z}\big) +\frac{\partial}{\partial z}\big((B_{21}+C_1)\frac {\partial v_2}{\partial x}\big)\\ &=B_{11}\big(-K''(\omega(z)-f(x))(f') ^2+K'(\omega(z)-f(x))f''\big)\\ &\quad +(B_{12}-C_1)\frac{\kappa}{B_{22}}f'K''(\omega(z)-f(x)) -\frac{\kappa^2}{B_{22}}K''(\omega(z)-f(x))\\ &\quad +\frac{\kappa}{B_{22}}(B_{21}+C_1)K''(\omega(z)-f(x))f' +B_{21}' K'(\omega(z)-f(x))f'\\ &\geq \frac{\mu B_{11}}{(1+\omega(z)-f(x))^2}\Big((f')^2 +\frac{\kappa^2}{B_{22}B_{11}}-\frac{\kappa(B_{12}-C_1)}{B_{22}B_{11}}f' +(1+\omega(z)-f(x))f''\Big)\\ &=\frac{\mu B_{11}}{(1+\omega(z)-f(x))^2}I_2, \end{aligned} \end{equation} since $B_{21}+C_1>0,\,K''(t)=-{\frac{\mu}{(1+t) ^2}<0,\,K'(t)=\frac{\mu}{1+t}>0,}$ $f'(x)\leq 0$, $B_{21}'\leq 0$ in $\mathcal{D}'(-h,0)$. We distinguish two cases: \noindent$\bullet$ For $x\in (x_R,f_d^{-1}(-h))$, we have $f''(x)=0$, ${f'(x)=-\frac{a+b}{R}=- \alpha}$ and then \begin{align*} I_2 &=\alpha^2+\frac{\kappa^2}{B_{22}B_{11}} +\frac{\alpha\kappa}{B_{22}B_{11}}(B_{12}-C_1)\\ &\geq \alpha^2+\frac{\kappa^2}{M^2}-\frac{\alpha\kappa}{m^2} (M+C_1)\to \frac{\kappa^2}{M^2}>0\quad\hbox{as } \alpha\to 0. \end{align*} So there exists $\alpha_0>0$ such that $I_2\geq 0 \quad\forall\alpha\in(0,\alpha_0)$. \noindent $\bullet$ For $x\in(-b,x_R)$, we have \begin{align*} I_2 &=\frac{R^2}{(R^2-(x+b)^2)}+\frac{\kappa^2}{B_{22}B_{11}}-1 +\frac{\kappa(B_{12}-C_1)}{B_{22}B_{11}}\frac{(x+b)}{\sqrt{R^2-(x+b)^2}}\\ &\quad -\frac{R^2}{(R^2-(x+b)^2)^{3/2}}(1+\omega(z)-f(x))\\ &=\frac{R^2}{(R^2-(x+b)^2)^{3/2}}\Big( \sqrt{R^2-(x+b)^2}+\frac{(R^2-(x+b)^2)^{3/2}}{R^2} \big(\frac{\kappa^2}{B_{22}B_{11}}-1\big)\\ &\quad+\frac{\kappa(B_{12}-C_1)}{B_{22}B_{11}}(x+b) \frac{(R^2-(x+b)^2)}{R^2}-(1+\omega(z)-f(x))\Big). \end{align*} Taking into account the fact that $x\in (-b,x_R]$ and (4.7), it follows that \begin{align*} &I_2 \\ &\geq \frac{R^2}{(R^2-(x+b)^2)^{3/2}}\Big(\!-R+2\sqrt{R^2-(x_R+b)^2} + R\big(\frac{\kappa^2}{M^2}-1\big)-C_2(x_R+b)-1\Big)\\ &=\frac{R^2}{(R^2-(x+b)^2)^{3/2}}\Big(-R+\frac{2R}{\sqrt{1+\alpha^2}} +R\big(\frac{\kappa^2}{M^2}-1\big)-C_2\frac{\alpha}{\sqrt{1+\alpha^2}}R-1\Big)\\ &=\frac{R^3}{(R^2-(x+b)^2)^{3/2}}\Big(\frac{2}{\sqrt{1+\alpha^2}}-2 +\big(\frac{\kappa^2}{2M^2}-\frac{C_2\alpha}{\sqrt{1+\alpha^2}}\big) +\frac{\kappa^2}{2M^2}-\frac{1}{R}\Big)\geq 0 \end{align*} provided that ${R>\frac{2M^2}{\kappa^2}}$ and $\alpha\in (0,\alpha'_0)$ for a small $\alpha'_0>0$. Finally for $\alpha\in(0,\alpha_2=\min(\alpha_0,{\alpha_0} '))$, we have $I_2\geq 0$ and from (4.9) we obtain (4.8). \end{proof} \begin{lemma}\label{lm4.3} There exists $\alpha_*>0$ and $R_*>0$ such that \begin{equation} \forall\alpha\in(0,\alpha_*),\;\forall R\geq R_*,\quad \mathop{\rm div}\big(B(z)\nabla v+\theta e_x\big)\geq 0\quad \hbox{in } \mathcal{D}'(D). \end{equation} \end{lemma} \begin{proof} Let $\xi\in \mathcal{D}(D)$, $\xi\geq 0$. Then by (4.5) and (4.8), \begin{align*} I(\xi) &=\int_D \big(B(z)\nabla v+\theta e_x\big)\nabla\xi\\ &=\int_{D_1} \big(B(z)\nabla v_1+\gamma_s e_x\big)\nabla\xi +\int_{D_2} \big(B(z)\nabla v_2+\gamma_f e_x\big)\nabla\xi\\ &=\langle-\mathop{\rm div}\big(B(z)\nabla v_1\big),\xi\rangle +\langle -\mathop{\rm div}\big(B(z)\nabla v_2\big),\xi\rangle\\ &\quad +\int_{[z=\omega^{-1}(x)]\cap\Omega}\big(B(z)(\nabla v_1- \nabla v_2).\nu+(\gamma_s-\gamma_f)\nu_x\big)\xi\\ &\leq \int_{[z=\omega^{-1}(x)]\cap\Omega}\big(B(z)(\nabla v_1-\nabla v_2).\nu+(\gamma_s-\gamma_f)\nu_x\big)\xi \end{align*} provided that $\alpha\in(0,\min(\alpha_1,\alpha_2))$ and ${R>\frac{2M^2}{\kappa^2}(h+1)}$. Moreover one has \begin{align*} \nabla v_1(x,\omega^{-1}(x)) &=G'(0) ^t\big(f'(x),-\frac{\kappa}{B_{22}(\omega^{-1}(x))}\big)\\ &=\frac{2Q_s}{h^2+2h} ^t\big(f'(x),-\frac{\kappa}{B_ {22}(\omega^{-1}(x))}\big) \end{align*} \begin{align*} \nabla v_2(x,\omega^{-1}(x))&=K'(0) ^t\big(f' (x),-\frac{\kappa}{B_{22}(\omega^{-1}(x))}\big)\\ &=\frac{Q_f}{\log(1+d)} ^t\big(f'(x),-\frac{\kappa}{B_ {22}(\omega^{-1}(x))}\big) \end{align*} \begin{align*} \nu(x,\omega^{-1}(x)) &=\frac{1}{\sqrt{1+{(\omega^{-1}of)'}^2(x)}} ^t \big(-(\omega^{-1}of)'(x),1\big)\\ &=\frac{1}{\sqrt{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)^2+{f'}^2(x)}} ^t\big(-f'(x),\frac {\kappa}{B_{22}(\omega^{-1}(x))}\big). \end{align*} Then \begin{align*} J_1&=B(z)(\nabla v_1-\nabla v_2).\nu+(\gamma_s-\gamma_f) \nu_x\\ &=\frac{-1}{\sqrt{{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)}^2+{f'}^2(x)}}{\Big(\frac{2Q_s}{h^2+2h}-\frac {Q_f}{\log(1+d)}\Big)}\\ & \times B(z) ^t\Big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1} (x))}\Big)^t{\big(f'(x),\frac{-\kappa}{B_{22}(\omega^ {-1}(x))}\big)}\\ &-\frac{f'(x)}{\sqrt{{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)}^2+{f'}^2(x)}}(\gamma_s-\gamma_f)\\ &=\frac{-1}{\sqrt{{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)}^2+{f'}^2(x)}}{\Big(\frac{Q_s}{h^2+2h}-\frac {Q_f}{\log(1+d)}\Big)}\\ & \times B(z) ^t{\Big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1} (x))}\Big)}^t{\big(f'(x),\frac{-\kappa}{B_{22}(\omega^ {-1}(x))}\big)}\\ &-\frac{1}{\sqrt{{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)}^2+{f'}^2(x)}}\Big(\frac{Q_s}{h^2+2h} B(z) ^t{\big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1}of (x))}\big)}\\ &\times {}^t{\big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1}of (x))}\big)}+(\gamma_s-\gamma_f)f'(x)\Big). \end{align*} Using (4.7), we obtain \begin{align*} &J_2\\ &=-\frac{1}{\sqrt{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)^2+{f'}^2(x)}}\Big(\frac{Q_s}{h^2+2h} B(z) ^t\big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1}of (x))}\big)\\ & \times{} ^t\big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1}of (x))}\big)+(\gamma_s-\gamma_f)f'(x)\Big)\\ &\leq \sqrt{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)^2+{f'}^2(x)}\Big(\frac{-mQ_s}{h^2+2h}+ (\gamma_s-\gamma_f)\frac{-f'(x)}{{f'}^2(x)+ \big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\Big)^2}\big). \end{align*} Note that $${F(x)=\frac{-f'(x)}{{f'}^2(x) +\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)^2}\leq -\frac { M^2}{\kappa^2}f'(x)} $$ and then \begin{itemize} \item If $x\in (x_R,f_d^{-1}(-h))$ we have $f'(x)=-\alpha$ and ${F(x)\leq\frac{ \alpha M^2} {\kappa^2}}$. \item If $x\in(-b,x_R)$ \begin{align*} F(x)&\leq-\frac{ M^2}{\kappa^2}f'(x)=\frac{ M^2} {\kappa^2}\frac{x+b}{\sqrt{(R^2-(x+b)^2)}} \\ &\leq \frac{ M^2}{\kappa^2}\frac{x_R+b}{R}=\frac{ M^2} {\kappa^2}\frac{\alpha }{\sqrt{1+\alpha^2}}\leq \alpha\frac { M^2}{\kappa^2}. \end{align*} So if ${\alpha<\frac{mQ_s}{h^2+2h}\frac{1}{\gamma_s-\gamma_f} \frac{\kappa^2}{ M^2}=\alpha_3}$, we get $J_2\leq 0$ and then \begin{align*} J_1\leq & \frac{1} {\sqrt{\big({\frac{\kappa}{B_{22}(\omega^{-1}of (x))}}\big)^2+{f'}^2(x)} } \Big(\frac{Q_f}{\log(1+d)}- \frac{Q_s}{h^2+2h}\Big)\\ & \times B(z){} ^t\Big(f'(x),\frac{-\kappa}{B_{22}(\omega^{-1} (x))}\Big)^t\Big(f'(x),\frac{-\kappa}{B_{22}(\omega^ {-1}(x))}\Big). \end{align*} Note that \[ \frac{Q_f}{\log(1+d)}<\frac{Q_s}{h^2+2h}\Leftrightarrow R>\frac{e^{\frac{Q_f}{Q_s}(h^2+2h)}-1+a\alpha}{\sqrt {1+\alpha^2}\big(\sqrt{1+\alpha^2}-1\big)}. \] Thus if $\alpha\in(0,\alpha_*=\min (\alpha_1,\alpha_2,\alpha_3))$ and \[ R\geq R_*(\alpha)=\max\Big(\frac{a}{\alpha},\frac{ 2M^2} {\kappa^2}(h+1),\frac{e^{\frac{Q_f}{Q_s}(h^2+2h)}-1+a\alpha} {\sqrt{1+\alpha^2}\big(\sqrt{1+\alpha^2}-1\big)}\Big), \] we get $I(\xi)\leq 0$ for all $\xi\in \mathcal{D}(D)$, $\xi\geq 0$. Therefore (4.10) holds. \end{itemize} \end{proof} \begin{lemma}\label{lm4.4} For $R>\max\big(\frac{a}{\alpha},\frac{a\alpha} {\sqrt{1+\alpha^2}(\sqrt{1+\alpha^2}-1)}\big)$, \begin{equation} v\leq\psi\quad\hbox{on }\partial D\quad\hbox{and} \quad\lim_{x\to-\infty}v(x,z)\leq \lim_{x\to-\infty}\psi (x,z). \end{equation} \end{lemma} \begin{proof} Indeed we have \begin{itemize} \item $v(x,-h)=G(f(x)-\omega(-h))=G(f(x)+h)\leq G(h)=Q_s=\psi(x,-h)$ for $x0$ which holds if \[R>\frac{a\alpha}{\sqrt{1+\alpha^2}(\sqrt{1+\alpha^2}- 1)}. \] \item $v(x,\omega^{-1}(x))=-K(f_d(x)-f(x)) =-K(d)=-Q_f\leq\psi(x,\omega^{-1}(x))$ \item $\lim_{x\to-\infty}v(x,z)=\lim_{x\to-\infty}v_1(x,z)=\lim_ {x\to-\infty}G(f(x)-\omega(z))\\ =G(-\omega(z))\leq {\frac{Q_s}{h}\omega(z)}=v_{-\infty} (z)=\lim_{x\to-\infty}\psi(x,z).$ \end{itemize} This (4.11) holds. \end{proof} \begin{lemma}\label{lm4.5} For all $\alpha\in(0,\alpha_*)$, $R>\max\big(R_* (\alpha),\frac{a\alpha}{\sqrt{1+\alpha^2}(\sqrt{1+\alpha^2}-1)}\big)$, one has \begin{equation} \int_D \left(B(z)\nabla(v-\psi_0)+(\theta-\gamma_0)e_x\right) \nabla\zeta=0\quad\forall\zeta\in\mathcal{D}(\mathbb{R}^2) \end{equation} where $\psi_0=\min(\psi,v)$ and $\gamma_0=\min(\gamma,\theta)$. \end{lemma} \begin{proof} Let $\zeta\in\mathcal{D}(\mathbb{R}^2)$ and consider for $i=1,2$ the function $\zeta_i$ defined as in the proof of Lemma \ref{lm2.1}. Then for $\epsilon>0$, ${\min\big(\zeta_i,\frac{v- \psi_0}{\epsilon}\big)\in H^1_0(D)}$ by (4.11) and is nonnegative with compact support. By (4.10) and $(P)i)$, we get \begin{align*} &\int_{D\cap[v-\psi_0\geq\epsilon\zeta_i]} B(z)\nabla(v- \psi_0)\nabla\zeta_i+\int_D (\theta-\gamma_0)\zeta_{ix}\\ &\leq \int_D (\theta-\gamma_0)\big(\zeta_i-\frac{v-\psi_0} {\epsilon}\big)_x^+\\ &\leq(\gamma_s-\gamma_f)\int_{I_0}\big(\zeta_i+\frac{\psi(f^ {-1}o\omega(z),z)}{\epsilon}\big)^+ -\big(\zeta_i-\frac{v(g(z),z)}{\epsilon}\big)^+ \end{align*} where $I_0=\{z\in(-h^*,0)/\,g(z)\omega^{- 1}(x)]\cap\Omega)$. Note that $\theta=\gamma_f=\gamma_0$ in $D_2$. We deduce from (4.12) \[ \int_{[z>\omega^{-1}(x)]\cap D} B(z)\nabla u.\nabla\zeta=0. \] Since $u$ vanishes on $[z=\omega^{-1}of_d(x)]$, then $u\in H_{\rm loc}^1(\Omega)$ and \[ \int_{[z>\omega^{-1}(x)]\cap \Omega} B(z)\nabla u.\nabla\zeta=0. \] Moreover $u\geq 0$ and $u=0$ in $\Omega\setminus D$. So by the strong maximum principle we get $u=0$ in $[z>\omega^{-1} (x)]\cap \Omega$ which leads to $v\leq \psi$ in $D_2$. In particular we obtain $\psi(f^{-1}o\omega(z),z)\geq v(f^{-1} o\omega(z),z)=v(x,\omega^{-1}(x))=0$. So $g(z)\geq f^{-1} o\omega(z)\quad\forall z\in(-h^*,0)$ and then \[ g(0-)\geq f^{-1}o\omega(0)=f^{-1}(0)=-b>-\infty. \] \end{proof} \subsection{The Nonlinear Homogeneous Case} Assume that $\mathcal{B} (z,\xi)=\vert\xi\vert^{q-2}\xi$. Then we have the following theorem. \begin{theorem}\label{thm4.3} $g(0-)=\lim_{z\to-\infty}g(z)>-\infty$. \end{theorem} We follow the proof of Theorem \ref{thm4.2} and we use the same notation with $\omega(z)=z$. We will need the following three Lemmas. \begin{lemma}\label{lm4.6} We have \begin{equation} \Delta_q v_1>0\quad\text{in}\quad D_1,\quad \forall R>R_1=\frac{h+1} {q-1}(\vert q-2\vert+q-1). \end{equation} \end{lemma} \begin{proof} Indeed we have \begin{align*} \Delta_q v_1 &=\mathop{\rm div}(\vert\nabla v_1\vert^{q-2}\nabla v_1)\\ &=\frac{\partial}{\partial x}\big(\vert\nabla v_1\vert^{q-2} \frac{\partial v_1}{\partial x}\big)+\frac{\partial} {\partial z}\big(\vert\nabla v_1\vert^{q-2}\frac{\partial v_1}{\partial z}\big)\\ &=(G'(f(x)-z))^{q-2}(1+{f'}^2(x))^{\frac{q-2}{2}}\\ &\times\Big((q-1)G''(f(x)-z)(1+{f'}^2(x)) +G'(f(x)-z)f''(x)\frac{1+(q-1){f'}^2(x)}{1+{f'}^2(x)}\Big). \end{align*} We distinguish two cases: \noindent $\bullet$ If $x<-b$ or $x>x_R$, we have $f''(x)=0$ and then $\Delta_q v_1>0\hbox{ in }D_1.$ \noindent $\bullet$ If $-bR_1$ we get $\Delta_q v_1>0$ in $D_1$ \end{proof} \begin{lemma}\label{lm4.7} For all $$R>R_2(\alpha) =\max\Big(\frac{a}{\alpha},\frac{(q-1)+\vert q-2\vert(h+1)} {(q-1)\Big(\frac{2}{\sqrt{1+\alpha^2}}-1\Big)}\Big), $$ we have \begin{equation} \Delta_q v_2> 0\quad\text{in}\quad D_2,\quad \forall\alpha\in(0,1). \end{equation} \end{lemma} \begin{proof} We have \begin{align*} \Delta_q v_2 &=(K'(z-f(x)))^{q-2}(1+{f'}^2(x))^ {\frac{q-2}{2}}\Big(-(q-1)K''(z-f(x))(1+{f'}^2(x))\\ &+K'(z-f(x))f''(x)\big(q-1-\frac{q-2}{1+{f'}^2(x)}\big)\Big). \end{align*} We distinguish two cases: \noindent $\bullet$ For $x_R 0\hbox{ in }D_2.$ \noindent $\bullet$ If $-b0}$ and ${R\geq\max(\frac{a}{\alpha},R_2(\alpha))}$, we get (4.14). \end{proof} \begin{lemma}\label{lm4.8} There exists $\alpha_*>0$ and $R_*>0$ such that \begin{equation} \forall\alpha\in(0,\alpha_*),\quad\forall R>R_*,\, \hbox{ one has }\,\Delta_q v +\theta_x\geq 0\hbox{ in }\mathcal{D}'(D). \end{equation} \end{lemma} \begin{proof} Let $\xi\in \mathcal{D}(D)$, $\xi\geq 0$. For $\alpha\in(0,1)$ and $R\geq\max\big(R_1, R_2(\alpha),\frac{a}{\alpha}\big)$, we have by (4.14)-(4.15) \begin{align*} I(\xi)&=\int_D(\vert\nabla v\vert^{q-2}\nabla v+\theta e_x) \nabla\xi\\ &=\int_{D_1}(\vert\nabla v_1\vert^{q-2}\nabla v_1+\gamma_s e_x)\nabla\xi+\int_{D_2}(\vert\nabla v_2\vert^{q-2}\nabla v_2+\gamma_f e_x)\nabla\xi\\ &=\langle-\Delta_q v_1,\xi>+<-\Delta_q v_2,\xi\rangle\\ &\quad+\int_{[z=f(x)]\cap D}\big((\vert\nabla v_1\vert^{q-2} \nabla v_1-\vert\nabla v_2\vert^{q-2}\nabla v_2).\nu+ (\gamma_s-\gamma_f)\nu_x\big)\xi\\ &\leq \int_{[z=f(x)]\cap D}\big((\vert\nabla v_1\vert^{q-2} \nabla v_1-\vert\nabla v_2\vert^{q-2}\nabla v_2).\nu+ (\gamma_s-\gamma_f)\nu_x\big)\xi\\ &=\int_{[z=f(x)]\cap D}J_1\xi. \end{align*} Recall that \begin{eqnarray*} &&\nabla v_1(x,f(x))=\frac{2Q_s}{h^2+2h}(f'(x),-1)\cr &&\vert\nabla v_1\vert^{q-2}\nabla v_1=\Big(\frac{2Q_s}{h^2+2h}\Big)^{q-1}\left(1+{f'}^2(x)\right)^{\frac{q-2}{2}}(f'(x),-1)\cr &&\nabla v_2(x,f(x))=\frac{Q_f}{\log(1+d)}(f'(x),-1)\cr &&\vert\nabla v_1\vert^{q-2}\nabla v_1=\Big(\frac{Q_f}{\log(1+d)}\Big)^{q-1}\left(1+{f'}^2(x)\right)^{\frac{q-2}{2}}(f'(x),-1)\cr &&\nu(x,f(x))=\frac{1}{\sqrt{1+{f'}^2(x)}}(-f'(x),1). \end{eqnarray*} Then \begin{align*} J_1&=(1+{f'}^2(x))^{\frac{q-1}{2}}\Big(\big(\frac {Q_f}{\log(1+d)}\big)^{q-1}-\big(\frac{2Q_s}{h^2+2h}\big)^{q-1} \Big)\\ &\quad +(\gamma_s-\gamma_f)\frac{-f'(x)}{\sqrt{1+{f'}^2(x)}}\\ &=\left(1+{f'}^2(x)\right)^{\frac{q-1}{2}}\Big(\big(\frac{Q_f} {\log(1+d)}\big)^{q-1}-\frac{1}{2}\big(\frac{2Q_s}{h^2+2h} \big)^{q-1}\Big)\\ &\quad+\left(1+{f'}^2(x)\right)^{\frac{q-1}{2}}\Big((\gamma_s- \gamma_f)\frac{-f'(x)}{(1+{f'}^2(x))^{q/2}}- \frac{1}{2}{\big(\frac{2Q_s}{h^2+h}\big)}^{q-1}\Big). \end{align*} Moreover we have \[ \frac{-f'(x)}{(1+{f'}^2(x))^{q/2}} \leq -f'(x)\leq \alpha\quad\forall x\in(-b,f_d^{-1}(- h)) \] and for \[ R>R_3(\alpha)= \frac{ e^{ \frac{Q_f(h^2+2h)}{2^{\frac{q-2}{q-1}}Q_s} } -1+a\alpha } {\sqrt{1+\alpha^2}(\sqrt{1+\alpha^2}-1)} \] one has \[ \left(\frac{Q_f}{\log(1+d)}\right)^{q-1}<\frac{1}{2}\left(\frac {2Q_s}{h^2+2h}\right)^{q-1}=\Big(\frac{Q_s}{2^{\frac{1} {q-1}}(h^2+2h)}\Big)^{q-1}. \] Thus if \[ \alpha \in\Big(0,\alpha_*=\min\big(1,\frac{1}{2 (\gamma_s-\gamma_f)}\big(\frac{2Q_s}{h^2+2h}\big)^{q-1}\big)\Big) \] and $R\geq R_*=\max\left(R_1,R_2(\alpha),R_3 (\alpha),\frac{a}{\alpha}\right)$, we get $J_1\leq 0$ and then $I(\xi)\leq 0$ for all $\xi\in \mathcal{D}(D)$, $\xi\geq 0$. Hence (4.15) holds.\end{proof} \begin{proof}[Proof of Theorem \ref{thm4.3}] Clearly for ${R>\max\big(\frac{a\alpha}{\sqrt{1+\alpha^2}(\sqrt {1+\alpha^2}-1)},\frac{a}{\alpha}\big)}$. Then we have (see Lemma \ref{lm4.4}) \[ v\leq\psi\quad\hbox{on }\partial D\quad\hbox{and} \quad\lim_{x\to-\infty}v(x,z)\leq \lim_{x\to-\infty}\psi (x,z). \] Using (4.15) and arguing as in the proof of Lemma \ref{lm2.1}, we can establish that for all $\alpha\in (0,\alpha_*)$ and ${R>\max\big(R_*,\frac{a \alpha}{\sqrt {1+\alpha^2}(\sqrt{1+\alpha^2}-1)}\big)}$, \[ (\Delta_q v-\Delta_q\psi_0)+(\gamma-\theta_0)_x=0 \quad\text{in}\quad \mathcal{D}'(D). \] Assume that $D_1\cap [\psi<0]\ne\emptyset$ and note that we have $\gamma=\gamma_f=\gamma_0$ and $\theta=\gamma_s$ in $D_1\cap[\psi<0]$. We also have $\psi=\psi_0$ in $D_1\cap[\psi<0]$ since $\psi<00$ in $D_1$. Therefore $D_1\cap[\psi<0]=\emptyset$ and $\psi\geq 0$ in $D_1$. In particular we have $g(z)\geq f^{- 1}(z)\quad\forall z\in(-h^*,0)$ and then $g(0-)\geq f^{-1} (0)=-b>-\infty$. \end{proof} \subsection{The Linear Homogeneous Case} Under the assumption $\mathcal{B} (z,\xi)=\xi$ we show that $g(0-)<0$. This result was announced in \cite{AD} without an explicit proof. In that paper the authors indicated that this result can be obtained by using hodograph techniques to get a semi-explicit expression for $\bar{\psi}$ which is the unique harmonic function satisfying the same boundary and limit conditions as $\psi$. However an explicit proof was not given. That is why we propose here a proof of this result. \begin{theorem}\label{thm4.4} \begin{equation} \lim_{z\to 0^-}g(z)=g(0-)\leq -\frac{h}{\pi}\log \Big(\frac {e^{\frac{a\pi(Q_s+Q_f)}{h Q_f}}-1}{e^{\frac{a\pi(Q_s+Q_f)} {h Q_f}}-e^{\frac{a\pi}{h }}}\Big)<0. \end{equation} \end{theorem} \begin{proof} Since we have $\Delta\psi\geq 0$ and $\Delta \bar{\psi}=0$ in $\Omega$, we get by taking into account the boundary conditions and the limit behavior at infinity that $\psi\leq \bar{\psi}$ in $\Omega$. Moreover if we denote by $\bar{g}$ the function defined by $\bar{\psi}(\bar{g}(z),z)=0$ $\forall z\in(-h^*,0)$, we get $g(z)\leq \bar{g}(z)$ for all $z\in(-h^*,0)$ which leads to \begin{equation} g(0-)\leq \bar{g}(0-). \end{equation} Next, we will prove that \begin{equation}\bar{g}(0-)=-\frac{h}{\pi}\log\left(\frac{e^{\frac{a\pi (Q_s+Q_f)}{h Q_f}}-1}{e^{\frac{a\pi(Q_s+Q_f)}{h Q_f}}-e^ {\frac{a\pi}{h }}}\right)<0. \end{equation} For simplicity we introduce the function \[ \psi_0(x,z)=\bar{\psi}(hx,h(z-1))\quad\hbox{for }(x,z) \in \mathbb{R}\times[0,1] \] which is harmonic in $\mathbb{R}\times(0,1)$ and satisfies $\psi_0(x,0)=Q_s$, $\psi_0(x,1)=\phi_0(x)$ and $\lim_{x\to \pm\infty}\psi_0(x,z)=v_{\pm\infty}(h(z-1))$. In \cite{LC}, one can find that \begin{align*} \psi_0(x,z)&=\mathop{\rm Re}\left( \frac{i}{2} \int_{-\infty}^{+\infty}Q_s \coth\frac{\pi(t-x-iz)}{2}+\phi_0 (ht)th\frac{\pi(t-x-iz)}{2}dt\right)\\ &=-\frac{Q_s}{2}\int_{-\infty}^{+\infty}\frac{ch\pi(t-x) +\cos\pi z}{sh^2\pi(t-x)+\sin^2\pi z}\sin\pi z dt\\ &\quad+\frac{1}{2}\int_{-\infty}^{+\infty}\phi_0(ht)\frac{\sin\pi z}{ch\pi(t-x)+\cos\pi z }dt\\ &=-\frac{Q_s}{2}\int_{-\infty}^{+\infty}\frac{ch\pi s+\cos\pi z}{sh^2\pi s+\sin^2\pi z}\sin\pi z ds+\frac{1}{2}I (x,z) \end{align*} where \begin{align*} I(x,z)&=\int_0^{\frac{a}{h}}-\frac{Q_f}{a}ht\frac{\sin\pi z} {ch\pi(t-x)+\cos\pi z}dt +\int_{\frac{a}{h}}^{+\infty}-Q_f\frac{\sin\pi z}{ch\pi(t-x) +\cos\pi z}dt\\ &=-\frac{Q_f}{a}h\int_{-x}^{\frac{a}{h}-x}t\frac{\sin\pi z} {ch\pi t+\cos\pi z}dt -\frac{Q_f}{a}hx\int_{-x}^{\frac{a}{h}-x}\frac{\sin\pi z} {ch\pi t+\cos\pi z}dt\\ &\quad-Q_f\int_{\frac{a}{h}-x}^{+\infty}\frac{\sin\pi z}{ch\pi t+\cos\pi z}dt. \end{align*} Then we deduce that \begin{align*} \frac{\partial \psi_0}{\partial x}&=\frac{1}{2}\frac {\partial I}{\partial x}=-\frac{Q_f}{2a}h\int_{-x}^{\frac{a} {h}-x}\frac{\sin\pi z}{ch\pi t+\cos\pi z}dt\\ &=-\frac{hQ_f }{a\pi}\Big[\mathop{\rm Arctan}\Big(\frac{e^{(\frac{a} {h}-x)\pi}+\cos\pi z} {\sin \pi z}\Big)-\mathop{\rm Arctan}\left(\frac{e^{-\pi x}+\cos\pi z}{\sin \pi z}\right)\Big] \end{align*} and for $(x,z)\in\Omega$ \begin{multline*} \bar{\psi}(x,z)=v_{-\infty}(z)\\ -\frac{hQ_f }{a\pi}\int_{- \infty}^{\frac{x}{h}} \Big(\mathop{\rm Arctan}\left(\frac{e^{-\pi t}-\cos\frac{\pi z}{h}}{\sin \frac{\pi z}{h}}\right)-\mathop{\rm Arctan}\Big(\frac{e^{\big(\frac{a}{h}-t\big)\pi}-\cos\frac {\pi z}{h}}{\sin \frac{\pi z}{h}}\Big)\Big)dt. \end{multline*} Using this formula we give an asymptotic behavior of $\bar {\psi}$ near $z=0$. Note that for $x\leq 0$ and ${t<\frac{x} {h}}$, we have $e^{-\pi t}>1$ and ${e^{\big({\frac{a}{h}}-t\big) \pi}>1}$. Therefore \begin{gather*} \mathop{\rm Arctan}\left(\frac{e^{-\pi t}-\cos\frac{\pi z}{h}}{\sin \frac {\pi z}{h}}\right)=-\frac{\pi}{2}+\frac{\pi}{h}\frac{1}{1-e^{- \pi t}}z+\frac{z^2}{2}p(z,t)\\ \mathop{\rm Arctan}\Big(\frac{e^{\big(\frac{a}{h}-t\big)\pi}-\cos\frac {\pi z}{h}}{\sin \frac{\pi z}{h}}\Big)=-\frac{\pi}{2}+\frac {\pi}{h}\frac{1}{1-e^{\big(\frac{a}{h}-t\big)\pi}}z+\frac {z^2}{2}q(z,t) \end{gather*} with $p(z,t), q(z,t)\approx e^{\pi t}$ as $t\to -\infty$. It follows that \begin{align*} 0=\bar{\psi}(\bar{g}(z),z)&=-\frac{Q_s}{h}z-\frac{Q_f}{a} z\int_{-\infty}^{\bar{g}(z)}\Big(\frac{1}{1-e^{-\pi t}}- \frac{1}{1-e^{\big(\frac{a}{h}-t\big)\pi}}\Big)dt\\ &\quad-\frac{hQ_f}{2\pi a}z^2\int_{-\infty}^{\bar{g}(z)}(p(z,t)-q (z,t))dt-\frac{Q_s}{h}\\ &\quad-\frac{Q_f}{a}\int_{-\infty}^{\bar{g}(z)}\left (\frac{1}{1-e^{-\pi t}}-\frac{1}{1-e^{\big(\frac{a}{h}- t\big)\pi}}\right)dt\\ &=z\int_{-\infty}^{\bar{g}(z)}(p(z,t)-q(z,t))dt. \end{align*} Letting $z\to 0$, we obtain \[ \frac{Q_f}{a}\int_{-\infty}^{\bar{g}(0)}\big(\frac{1}{1-e^ {-\pi t}}-\frac{1}{1-e^{\big(\frac{a}{h}-t\big)\pi}}\big) dt=\frac{Q_s}{h} \] and by evaluating the last integral we get (4.17). Taking into account (4.17) and (4.18), we obtain (4.16).\end{proof} \begin{remark}\label{rmk4.1} \rm As a consequence of the above theorem, we have also $g(0-)<0 $ when \[ {\mathcal{B}(z,\xi)=\left(\frac{C^2}{B(z)}\xi_1,B(z)\xi_2\right) \quad\hbox{with } C={\frac{-h}{{\int_{-h}^0}{\frac{ds}{B (s)}}}.} } \] \end{remark} Let $v(x,z)=\bar{\psi}(x,\omega(z))$ with ${\omega(z) =C{\int_z^0\frac{ds}{B(s)}}}$. We remark that \[ \mathop{\rm div}(\mathcal{B}(z,\nabla v))=\frac{C^2}{B(z)}\Delta\bar{\psi} (x,\omega(z))=0. \] Moreover $v=\psi$ on $\partial\Omega$ and $\lim_ {x\to\pm\infty}v=\lim_{x\to\pm\infty}\psi$. Then we deduce that $\psi\leq v$ in $\Omega$ which leads to \[ \psi(\bar{g}(\omega(z)),z)\leq v (\bar{g}(\omega(z)),z) =\bar{\psi}(\bar{g}(\omega(z)),\omega(z))=0\quad\forall z\in (-h^*,0). \] Therefore $g(z)\leq \bar{g}(\omega(z))$ for all $z\in(-h^*,0)$ and then $g(0-)\leq \bar{g}(0-)<0$. \section*{Appendix}\label{A:Append} This section is devoted to some technical Lemmas. \begin{lemma}[Strong maximum principle] \label{lmA.1} Let $\Omega$ be a domain of $\mathbb{R}^2$ and let $\mathcal{B}$ defined on $\Omega\times\mathbb{R}^2$ by $\mathcal{B}(X,\xi)=\vert B(X)\xi\vert^{q-2}B^2(X) \xi$, where $B(X)$ is a locally Lipschitz continuous symmetric and strictly elliptic matrix. Assume that $u_1$, $u_2 \in W_{\rm loc}^{1,q}(\Omega)$ are such that \begin{gather} \mathop{\rm div}(\mathcal{B}(X,\nabla u_1)) =\mathop{\rm div}(\mathcal{B}(X,\nabla u_2))=0 \quad\hbox{in }\mathcal{D}'(\Omega),\label{A.1} \\ u_1\leq u_2 \quad\hbox{in }\Omega.\label{A.2} \end{gather} Then we have either $u_1=u_2$ in $\Omega$ or $u_10$ and assume that $S_{eg}=\{(x,z_0)/\,\vert x-x_0\vert\leq r\}\subset\Gamma$, then we cannot have \[ \forall(x,z)\in B_r(x_0,z_0)\setminus S_{eg}\quad\psi(x,z) \ne 0 \] where $B_r(x_0,z_0)$ is the open ball of center $(x_0,z_0)$ and radius $r$. \end{lemma} The proof follows as in \cite[Lemma 5.1]{CCL} and uses Lemma \ref{lmA.1}. \subsection*{Acknowledgments} We would like to thank KFUPM for the facilities provided and for its support. \begin{thebibliography}{99} \bibitem{AD} H. W. Alt \& C. J. Van Duijn: \emph{A stationary flow of fresh and salt groundwater in coastal aquifer}. Nonlinear Analysis. Theory. Methods \& Applications. Vol.14 No.8 pp.625-656, (1990). \bibitem{AS} G. Alessandrini \& M. Sigalotti: \emph{ Geometric properties of solutions to the inhomogeneous anisotropic p- Laplace equation in dimension two}. Ann. Acad. Sci. Fenn. Math. 26 (2001), no. 1, 249-266. \bibitem{B} J. Bear: \emph{Dynamics of fluids in porous media}. American Elsevier Environmental Science Series (1972). \bibitem{CCL} J. Carrillo, S. Challal \& A. Lyaghfouri: \emph{A free boundary problem for a flow of fresh and salt groundwater with nonlinear Darcy's law}. Advances in Mathematical Sciences and Applications 12 , no. 1, 191-215, (2002). \bibitem{CL1} S. Challal \& A. Lyaghfouri: \emph{ A stationary flow of fresh and salt groundwater in a coastal aquifer with nonlinear Darcy's law}. Applicable Analysis Vol. 67, 295-312, (1997). \bibitem{CL2} S. Challal \& A. Lyaghfouri: \emph{A stationary flow of fresh and salt groundwater in a heterogeneous coastal aquifer}. Bollettino della Unione Matematica Italiana (8) 3-B, 505-533, (2000). \bibitem{Ci} P. Ciarlet: \emph{Introduction to numerical linear algebra and optimisation}. Cambridge University Press, (1989). \bibitem{D} L. Damascelli: \emph{Comparison theorems for some quasilinear degenerate elliptic operators and applications to symmetry and monotonicity results}. Ann. Inst. H. Poincar\'e 15, 493-516, (1998). \bibitem{DL} R. Dautray \& J.L. Lions: \emph{Analyse Math\'ematiques et Calcul Num\'erique Pour les Sciences et les Techniques,} Tome 2 \emph{L'op\'erateur de Laplace.} Masson, Paris (1987). \bibitem{K} J. Heinonen, T. Kilpel\"ainen, and O. Martio: \emph {Nonlinear Potential Theory of Degenerate Elliptic Equations.} Oxford Science Publications, (1993). \bibitem{L} A. Lyaghfouri: \emph{On the Uniqueness of the Solution of a Nonlinear Filtration Problem through a Porous medium.} Calculus of Variations and Partial Differential Equations Vol. 6, No. 1, (1998), 67-94. \bibitem{LC} Lavrentiev \& Chabat: \emph{Problems of hydrodynamics and their mathematical models}. Second edition. Izdat. Nauka, Moscow, (1977). \end{thebibliography} \end{document}