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


A free boundary problem for semi-linear elliptic equation and its applications

Jianfeng Cheng1,2,  Lili Du2
Abstract.

In this paper, we consider a free boundary problem of a semilinear nonhomogeneous elliptic equation with Bernoulli’s type free boundary. The existence and regularity of the solution to the free boundary problem are established by use of the variational approach. In particular, we establish the Lipschitz continuity and non-degeneracy of a minimum, and regularity of the free boundary. As a direct and important application, the well-posedness results on the steady, incompressible inviscid jet and cavitational flow with general vorticity are also obtained in this paper.

This work is supported in part by NSFC grant 11971331.
E-Mail: jianfengcheng@126.com (J. Cheng), E-mail: dulili@scu.edu.cn (L. Du).

1 The Institute of Mathematical Sciences,

The Chinese University of Hong Kong, Hong Kong.

2 Department of Mathematics, Sichuan University,

Chengdu 610064, P. R. China.

2010 Mathematics Subject Classification: 76B10; 76B03; 35Q31; 35J25.

Key words: Semilinear elliptic equation; Free boundary; Regularity; Incompressible flow; General vorticity.

1. Introduction

In this paper, we investigate the free boundary problem of a semilinear nonhomogeneous elliptic equation

{Δψ=f(ψ)inΩ{ψ>0},|ψ|=λ(X)onΩ{ψ>0},ψ=Ψ0onΩ,\left\{\begin{array}[]{ll}-\Delta\psi=f(\psi)&\text{in}\ \ \Omega\cap\{\psi>0\},\\ |\nabla\psi|=\lambda(X)&\text{on}\ \ \Omega\cap\partial\{\psi>0\},\\ \psi=\Psi_{0}&\text{on}\ \ \partial\Omega,\end{array}\right. (1.1)

where Ω\displaystyle\Omega is a connected open and bounded domain in 2\displaystyle\mathbb{R}^{2}, Ω\displaystyle\partial\Omega is a local Lipschitz graph. The given functions Ψ0C0,1(Ω¯)\displaystyle\Psi_{0}\in C^{0,1}(\bar{\Omega}) with Ψ00\displaystyle\Psi_{0}\geq 0 and λ(X)C0,β(Ω¯)\displaystyle\lambda(X)\in C^{0,\beta}(\bar{\Omega}) with 0<λ1λ(X)λ2<+\displaystyle 0<\lambda_{1}\leq\lambda(X)\leq\lambda_{2}<+\infty, f(ψ)\displaystyle f(\psi) is so-called vorticity strength function and fC1,β()\displaystyle f\in C^{1,\beta}(\mathbb{R}), 0<β<1\displaystyle 0<\beta<1.

The semilinear nonhomogeneous equation Δψ=f(ψ)\displaystyle-\Delta\psi=f(\psi) characterizes the steady, incompressible flow of an ideal fluid in 2\displaystyle\mathbb{R}^{2} with non-zero variable vorticity. A wide class of vorticity distributions is considered. We are interesting in the existence, regularity and geometric properties of the solution ψ\displaystyle\psi and the free boundary Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\}.

On the other hand, physical motivation for our study lies in the proof of minimizing the functional

J(ψ)=Ω|ψ|2+F(ψ)+λ2(X)I{ψ>0}dX,J(\psi)=\int_{\Omega}|\nabla\psi|^{2}+F(\psi)+\lambda^{2}(X)I_{\{\psi>0\}}dX,

which is related closely to the jet flow problem with variable vorticity. Here, F(t)=20tf(s)𝑑s\displaystyle F(t)=-2\int_{0}^{t}f(s)ds, IE\displaystyle I_{E} is the characteristic function of the set E\displaystyle E. Here and after, denote dX=dxdy\displaystyle dX=dxdy for simplicity.

In the remarkable paper [2] by H. Alt and L. Caffarelli, some results on Lipschitz continuity, non-degeneracy lemma of a minimizer ψ\displaystyle\psi, and the analyticity of the free boundary were obtained for the special case f(ψ)=0\displaystyle f(\psi)=0. The mathematical results of [2] were used immediately in the study of jet flows [3, 4, 5, 15] and impinging jet flows [12, 14] of inviscid, irrotational and incompressible fluid. For the nonhomogeneous problem (f(ψ)0\displaystyle f(\psi)\neq 0), A. Friedman [24] established the first result on the regularity of the solution and the free boundary for the linear nonhomogeneous case Δψ=P(x,y)\displaystyle\Delta\psi=P(x,y). Based on this result in [24], the well-posedness result of cavitational flow [16, 24] and impinging jet flows [13] of inviscid and incompressible fluid with constant vorticity (P(x,y)=P0)\displaystyle(P(x,y)=P_{0}) were obtained. The similar results were extended to the quasilinear homogeneous case

J(ψ)=ΩF(|ψ|2)+λ2(X)I{ψ>0}dXJ(\psi)=\int_{\Omega}F(|\nabla\psi|^{2})+\lambda^{2}(X)I_{\{\psi>0\}}dX

in [6] and nonlinear homogeneous case

J(ψ)=Ωaij(ψ)DiψDjψ+λ2(X)I{ψ>0}dXJ(\psi)=\int_{\Omega}a_{ij}(\psi)D_{i}\psi D_{j}\psi+\lambda^{2}(X)I_{\{\psi>0\}}dX

in [29].

The first purpose of this paper is to establish the existence, regularity of the solution to the free boundary problem (1.1), and extend the classical results in [2] to the semilinear nonhomogeneous elliptic equation. In particular, the Lipschitz continuity, non-degeneracy lemma of the solution and regularity of the free boundary are obtained (please see Theorem 2.4 for Lipschitz continuity and Theorem 3.15 for the regularity of the free boundary).

On the other hand, there is a large number of literatures on the regularity criteria of the free boundary for linear elliptic problem,

{i,j=1naij(x)xixj2ψ=f(x)inΩ{ψ>0},|ψ|=λ(x)onΩ{ψ>0}.\left\{\begin{array}[]{ll}\sum_{i,j=1}^{n}a_{ij}(x)\partial^{2}_{x_{i}x_{j}}\psi=f(x)&\text{in}\ \ \Omega\cap\{\psi>0\},\\ |\nabla\psi|=\lambda(x)&\text{on}\ \ \Omega\cap\partial\{\psi>0\}.\end{array}\right. (1.2)

For the Laplace operator and f0\displaystyle f\equiv 0, L. Caffarelli showed in his pioneer work [9] that Lipschitz free boundary is C1,α\displaystyle C^{1,\alpha}-smooth, furthermore, he also showed in [10] that ”flat” free boundary is Lipschitz. And higher regularity, such as C\displaystyle C^{\infty}-smoothness and analyticity of the free boundary follow from the elegant work of D. Kinderlehrer and L. Nirenberg [27]. In the case of the homogeneous case f0\displaystyle f\equiv 0, the regularity criteria results in spirit of works [9, 10] have been subsequently obtained for more general operators, such as concave fully nonlinear uniform elliptic operators of the form F(D2ψ)\displaystyle F(D^{2}\psi) in [32, 33] and nonconcave fully nonlinear uniform elliptic operators of the form F(D2ψ,Dψ)\displaystyle F(D^{2}\psi,D\psi), and references [11, 21, 22]. Moreover, Silva showed that the Lipschitz free boundary of the problem (1.2) in non-homogenous case (f0)\displaystyle(f\neq 0) is C1,α\displaystyle C^{1,\alpha} in [30]. Recently, Weiss and Zhang [34] investigated the regularity of the free boundary problem of semilinear elliptic equation Δψ=f(ψ)\displaystyle-\Delta\psi=f(\psi), and showed the regularity criteria that the Wloc1,1\displaystyle W_{loc}^{1,1} free boundary is a graph implies the C2,α\displaystyle C^{2,\alpha}-regularity of the free boundary. However, we would like to emphasize that the results in this paper are not the conditional regularity of the free boundary and we do not give any apriori assumptions and topological property on the free boundary and the solution.

An other related interesting problem is the semilinear Dirichlet problem with degenerate gradient on the free boundaries,

{Δψ=f(x,ψ)inΩ{ψ>0},ψ=|ψ|=0onΩ{ψ>0}.\left\{\begin{array}[]{ll}\Delta\psi=f(x,\psi)&\text{in}\ \ \Omega\cap\{\psi>0\},\\ \psi=|\nabla\psi|=0&\text{on}\ \ \Omega\cap\partial\{\psi>0\}.\end{array}\right. (1.3)

The problem is used for modeling the distribution of a gas with density, ψ(x)\displaystyle\psi(x), in reaction with a porous catalyst pellet Ω\displaystyle\Omega. Alt and Phillips [8] investigated the regularity and the geometry properties of the solution and the free boundaries with some special structural conditions on the force term f(x,ψ)\displaystyle f(x,\psi). The degenerate gradient condition on the free boundaries implies that desired optimal regularity of the solution is C1,α\displaystyle C^{1,\alpha} in Ω\displaystyle\Omega. However, in this paper, the gradient of the solution is non-zero on the free boundaries, thus the optimal regularity desired here is only Lipschitz in Ω\displaystyle\Omega. This is the main difference between the free boundary problem (1.1) and the one (1.3).

The second purpose of this paper is to establish the well-posedness theory of the jet flows of inviscid, incompressible fluid with general vorticity issuing from a semi-infinitely long nozzle for symmetric case (see Figure 1). It’s a direct and important application of the mathematical theory of the free boundary problem (1.1).

Refer to caption
Figure 1. Incompressible jet flow

Before we proceed with the bulk of this paper, we would like to mention the previous results on the well-posedness of the free boundary problem for incompressible inviscid jet flows.

For the irrotational inviscid incompressible flow without surface tension, the existence of an axisymmetric jet flow [5], an asymmetric jet flow [3] and impinging jet flow [12, 14] has been established based on the fundamental work [2]. The main advantage for two-dimensional irrotational flow is that the stream function ψ\displaystyle\psi solves the linear elliptic equation with Bernoulli’s type free boundary. And then the conformal mapping and Green function approach for the linear elliptic equation work for the irrotational flow. Furthermore, for the simplified setting of constant vorticity, which corresponds to a constant vorticity strength f(ψ)P0\displaystyle f(\psi)\equiv P_{0} in (1.1), the well-posedness of axially symmetric cavitational flow in [24], symmetric impinging jet flow in [13] and symmetric cavitational flow in an infinitely long nozzle in [16] were established, based on the regularity of the free boundary problem with Poisson equation Δψ=P0\displaystyle\Delta\psi=P_{0}. The assumption of an inviscid incompressible jet with non-zero constant vorticity provides us with the simplest case of a flow that is not irrotational and is attractive for its analytical tractability. However, the simplicity setting is not a mere mathematical convenience, as it is also physical relevant. Indeed, the vorticity remains invariant along the each streamline for inviscid fluid, therefore the fluid possesses constant vorticity as long as we impose the constant vorticity in the upstream.

Generations of scientists working in fluid dynamics has recognized the importance of the vorticity. It has provided a powerful qualitative description for many of the important phenomena of fluid mechanics. From the mathematical point of view, the strong nonlinearity of the equations of vortex motion has made the analysis difficult.

In the present paper, to establish the well-posedness of the incompressible jet issuing from a semi-infinitely long nozzle with general vorticity, we will impose the vorticity in the inlet of the nozzle, and then the vorticity strength function f(ψ)\displaystyle f(\psi) is determined uniquely along the streamlines. To realize this idea, one of the key points is to show the well-posedness of the streamlines. A free boundary problem with semilinear nonhomogeneous elliptic equation as (1.1) is formulated. An associated variational problem is shown to possess a minimizer, which yields a solution of semilinear free boundary problem.

2. Regularity and non-degeneracy of the solutions

To investigate the free boundary of the semilinear elliptic equation (1.1), based on the works [2, 6, 24] by Alt, Caffarelli and Friedman, we study the variational problem in this section, and establish the non-degeneracy lemma and the regularity of the solution.

2.1. The variational problem

Let Ω2\displaystyle\Omega\subset\mathbb{R}^{2} be a bounded and connected open domain and Ω\displaystyle\partial\Omega is a locally Lipschitz graph, IE\displaystyle I_{E} is the characteristic function of a set E2\displaystyle E\subset\mathbb{R}^{2} and f(t)C1,β((,+))\displaystyle f(t)\in C^{1,\beta}((-\infty,+\infty)), which satisfies that

0f(t)Λ for any t0 and Λf(t)0 for any t,\text{$\displaystyle 0\leq f(t)\leq\Lambda$ for any $\displaystyle t\leq 0$ and $\displaystyle-\Lambda\leq f^{\prime}(t)\leq 0$ for any $\displaystyle t\in\mathbb{R}$}, (2.1)

where Λ0\displaystyle\Lambda\geq 0 is a constant. Denote

F(t)=20tf(s)𝑑s.F(t)=-2\int_{0}^{t}f(s)ds.

It is easy to check that

F(0)=0,2ΛF(t)0for any t0 and 0F′′(t)2Λfor any t.F(0)=0,\ -2\Lambda\leq F^{\prime}(t)\leq 0\ \text{for any $\displaystyle t\leq 0$ and}\ 0\leq F^{\prime\prime}(t)\leq 2\Lambda\ \text{for any $\displaystyle t\in\mathbb{R}$}. (2.2)

Define a function ΨC0(Ω¯)C2,α(Ω)\displaystyle\Psi\in C^{0}(\bar{\Omega})\cap C^{2,\alpha}(\Omega) and Ψ\displaystyle\Psi satisfies

ΔΨ+f(Ψ)0in Ω and Ψ>0 in Ω.\Delta\Psi+f(\Psi)\leq 0\ \ \text{in $\displaystyle\Omega$ and $\displaystyle\Psi>0$ in $\displaystyle\Omega$}. (2.3)

Consider an energy functional

J(ψ)=Ω(|ψ|2+F(ψ)+λ2(X)I{ψ>0})𝑑X,J(\psi)=\int_{\Omega}\left(|\nabla\psi|^{2}+F(\psi)+\lambda^{2}(X)I_{\{\psi>0\}}\right)dX,

and an admissible set

K={ϕH1(Ω)ϕΨa.e. in Ω,ϕ=Ψ0onS},K=\{\phi\in H^{1}(\Omega)\mid~\phi\leq\Psi\ \text{a.e. in $\displaystyle\Omega$},~~\phi=\Psi_{0}\ \text{on}~S\},

where S\displaystyle S is a given subset of Ω\displaystyle\partial\Omega with 1(S)>0\displaystyle\mathcal{H}^{1}(S)>0, Ψ0H1(Ω)\displaystyle\Psi_{0}\in H^{1}(\Omega) and 0Ψ0Ψ\displaystyle 0\leq\Psi_{0}\leq\Psi a.e. on S\displaystyle S, λ(X)C0,β(Ω¯)\displaystyle\lambda(X)\in C^{0,\beta}(\bar{\Omega}) and satisfies

0<λ1λ(X)λ2<+for any XΩ¯.0<\lambda_{1}\leq\lambda(X)\leq\lambda_{2}<+\infty\ \ \text{for any $\displaystyle X\in\bar{\Omega}$}. (2.4)

The variational problem (P)\displaystyle(P): Find a ψK\displaystyle\psi\in K, such that

J(ψ)=minϕKJ(ϕ).J(\psi)=\min_{\phi\in K}J(\phi).
Remark 2.1.

For the special case F(t)0\displaystyle F(t)\equiv 0, the variational problem have been studied by Alt and Caffarelli in [2]. They established the Lipschitz continuity and non-degeneracy of the minimizer ψ\displaystyle\psi, and obtained the regularity of the free boundary. Moreover, Friedman in [24] investigated the variational problem for F(t)=2P(x,y)t\displaystyle F(t)=-2P(x,y)t with P(x,y)0\displaystyle P(x,y)\geq 0 and P(x,y)C0,1(Ω¯)\displaystyle P(x,y)\in C^{0,1}(\bar{\Omega}), and established the regularity of the free boundary.

As the first step, the existence of the minimizer ψ\displaystyle\psi to the variational problem (P)\displaystyle(P) can be obtained by using the similar arguments in Lemma 1.3 in [2], we omit it here.

2.2. Lipschitz continuity of the minimizer

In this subsection, we will obtain the Lipschitz continuity of the minimizer ψ\displaystyle\psi.

We first focus on the Hölder continuity of the minimizer ψ\displaystyle\psi in Ω\displaystyle\Omega, and show that the minimizer ψ\displaystyle\psi satisfies the semilinear elliptic equation in Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\}.

Lemma 2.1.

(1) ψCα(Ω)\displaystyle\psi\in C^{\alpha}(\Omega) for some α(0,1)\displaystyle\alpha\in(0,1) and ψ0\displaystyle\psi\geq 0 in Ω\displaystyle\Omega.
(2) ψ\displaystyle\psi satisfies that

Δψ+f(ψ)0\displaystyle\Delta\psi+f(\psi)\geq 0 in Ω\displaystyle\Omega (2.5)

in the weak sense and

Δψ+f(ψ)=0\displaystyle\Delta\psi+f(\psi)=0 in Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\}. (2.6)

Furthermore, ψC2,α(D)\displaystyle\psi\in C^{2,\alpha}(D) for any compact subset D\displaystyle D of Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\}.

Proof.

(1) Let BrΩ\displaystyle B_{r}\subset\Omega and define a function ϕ\displaystyle\phi as follows

Δϕ+f(ϕ)=0inBr,andϕ=ψoutsideBr.\Delta\phi+f(\phi)=0\ \ \ \text{in}~~B_{r},\ \ \text{and}\ \ \ \phi=\psi\ \ \text{outside}\ \ B_{r}.

The maximum principle gives that ϕΨ\displaystyle\phi\leq\Psi in Ω\displaystyle\Omega, and thus ϕK\displaystyle\phi\in K. Then we have

0J(ϕ)J(ψ)=Br|(ϕψ)|2+2(ϕψ)ϕ+F(ϕ)F(ψ)dX+Brλ2(X)(I{ϕ>0}I{ψ>0})𝑑XBr|(ϕψ)|2dX+λ22BrI{ϕ>0}𝑑X,\begin{array}[]{rl}0\leq&J(\phi)-J(\psi)\\ =&\int_{B_{r}}-|\nabla(\phi-\psi)|^{2}+2\nabla(\phi-\psi)\cdot\nabla\phi+F(\phi)-F(\psi)dX\\ &+\int_{B_{r}}\lambda^{2}(X)(I_{\{\phi>0\}}-I_{\{\psi>0\}})dX\\ \leq&\int_{B_{r}}-|\nabla(\phi-\psi)|^{2}dX+\lambda_{2}^{2}\int_{B_{r}}I_{\{\phi>0\}}dX,\end{array} (2.7)

which implies that

Br|(ϕψ)|2𝑑Xλ22BrI{ϕ>0}𝑑Xλ22πr2.\begin{array}[]{rl}\int_{B_{r}}|\nabla(\phi-\psi)|^{2}dX\leq\lambda_{2}^{2}\int_{B_{r}}I_{\{\phi>0\}}dX\leq\lambda_{2}^{2}\pi r^{2}.\end{array} (2.8)

With the aid of gradient L2\displaystyle L^{2}-estimate in (2.8), we can now use the method of Morrey in Theorem 5.3.6 in [28] to deduce the Hölder continuity of the minimizer.

Set ψε=ψεmin{ψ,0}\displaystyle\psi^{\varepsilon}=\psi-\varepsilon\min\{\psi,0\} for any ε(0,1)\displaystyle\varepsilon\in(0,1). It is clear that ψεK\displaystyle\psi^{\varepsilon}\in K and ψεψ\displaystyle\psi^{\varepsilon}\geq\psi in Ω\displaystyle\Omega. Furthermore, ψ>0\displaystyle\psi>0 if and only if ψε>0\displaystyle\psi^{\varepsilon}>0 in Ω\displaystyle\Omega, and one has

0J(ψε)J(ψ)Ω((1ε)21)|min{ψ,0}|2εF((1ε)min{ψ,0})min{ψ,0}dX.\begin{array}[]{rl}0\leq&J(\psi^{\varepsilon})-J(\psi)\\ \leq&\int_{\Omega}((1-\varepsilon)^{2}-1)|\nabla\min\{\psi,0\}|^{2}-\varepsilon F^{\prime}((1-\varepsilon)\min\{\psi,0\})\min\{\psi,0\}dX.\end{array} (2.9)

Since F(t)0\displaystyle F^{\prime}(t)\leq 0 for any t0\displaystyle t\leq 0 in (2.2), it follows from (2.9) that

Ω|min{ψ,0}|2𝑑X0,\int_{\Omega}|\nabla\min\{\psi,0\}|^{2}dX\leq 0,

which implies that

ψ(x,y)0 in Ω.\text{$\displaystyle\psi(x,y)\geq 0$ in $\displaystyle\Omega$}.

(2) For any ξC0(Ω)\displaystyle\xi\in C_{0}^{\infty}(\Omega) with ξ0\displaystyle\xi\geq 0 in Ω\displaystyle\Omega, it follows from the statement (1) that ψεξK\displaystyle\psi-\varepsilon\xi\in K and {ψεξ>0}{ψ>0}\displaystyle\{\psi-\varepsilon\xi>0\}\subset\{\psi>0\} for any ε>0\displaystyle\varepsilon>0, then we have

0limε0(1εΩ|ψεξ|2|ψ|2+F(ψεξ)F(ψ)dX)=Ω2ψξ+F(ψ)ξdX,\begin{array}[]{rl}0\leq&\lim_{\varepsilon\rightarrow 0}\left(\frac{1}{\varepsilon}\int_{\Omega}|\nabla\psi-\varepsilon\nabla\xi|^{2}-|\nabla\psi|^{2}+F(\psi-\varepsilon\xi)-F(\psi)dX\right)\\ =&-\int_{\Omega}2\nabla\psi\cdot\nabla\xi+F^{\prime}(\psi)\xi dX,\end{array}

which gives (2.5).

By virtue of the continuity of ψ\displaystyle\psi, we have that Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\} is open. For any ξC0(Ω{ψ>0})\displaystyle\xi\in C_{0}^{\infty}(\Omega\cap\{\psi>0\}), it is easy to check that ψ(x,y)εξ(x,y)Ψ(x,y)\displaystyle\psi(x,y)-\varepsilon\xi(x,y)\leq\Psi(x,y) in Ω\displaystyle\Omega and {ψεξ>0}{ψ>0}\displaystyle\{\psi-\varepsilon\xi>0\}\subset\{\psi>0\} for small |ε|>0\displaystyle|\varepsilon|>0. Obviously, ψεξK\displaystyle\psi-\varepsilon\xi\in K and we have

0εΩ2ψξ+F(ψεξ)ξdX+o(ε),\begin{array}[]{rl}0\leq-\varepsilon\int_{\Omega}2\nabla\psi\cdot\nabla\xi+F^{\prime}(\psi-\varepsilon\xi)\xi dX+o(\varepsilon),\end{array}

which implies that

0=Ω2ψξ+F(ψεξ)ξdXfor any ε.0=\int_{\Omega}2\nabla\psi\cdot\nabla\xi+F^{\prime}(\psi-\varepsilon\xi)\xi dX\ \ \ \text{for any $\displaystyle\varepsilon$}. (2.10)

Taking ε0\displaystyle\varepsilon\rightarrow 0 in (2.10) yields (2.6).

It follows from the Schauder interior estimate in [26] that ψC2,α(D)\displaystyle\psi\in C^{2,\alpha}(D) for any compact subset D\displaystyle D of Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\}.

Denote

Γ=Ω{ψ>0} the free boundary of ψ.\Gamma=\Omega\cap\partial\{\psi>0\}\ \ \text{ the free boundary of $\displaystyle\psi$}.

The dynamic boundary condition on the free boundary Γ\displaystyle\Gamma can be verified in the following.

Proposition 2.2.

If λ(X)C1,β(Ω¯)\displaystyle\lambda(X)\in C^{1,\beta}(\bar{\Omega}), for any minimizer ψ\displaystyle\psi, we have

limε0Ω{ψ>ε}(|ψ|2λ2(X)F(ψ))ηνε𝑑S=0,\lim_{\varepsilon\downarrow 0}\int_{\Omega\cap\partial\{\psi>\varepsilon\}}\left(|\nabla\psi|^{2}-\lambda^{2}(X)-F(\psi)\right)\eta\cdot\nu_{\varepsilon}dS=0, (2.11)

for any 2-vector η(C01(Ω))2\displaystyle\eta\in\left(C_{0}^{1}(\Omega)\right)^{2}, where νε\displaystyle\nu_{\varepsilon} is the outer normal vector to Ω{ψ>ε}\displaystyle\Omega\cap\partial\{\psi>\varepsilon\} with ε>0\displaystyle\varepsilon>0. Similarly, if a segment lΩ{ψ>0}\displaystyle l\subset\partial\Omega\cap\partial\{\psi>0\} is C1,α\displaystyle C^{1,\alpha} and ψ=0\displaystyle\psi=0 on l\displaystyle l, we have

|ψ|λ(X)on l.|\nabla\psi|\geq\lambda(X)\ \ \text{on $\displaystyle l$}. (2.12)
Proof.

We define a diffeomorphism

Y=ξδ(X):ΩΩY=\xi_{\delta}(X):\Omega\rightarrow\Omega

by ξδ(X)=X+δη(X)\displaystyle\xi_{\delta}(X)=X+\delta\eta(X) for any η(X)(C01(Ω))2\displaystyle\eta(X)\in\left(C_{0}^{1}(\Omega)\right)^{2}, where δ\displaystyle\delta is a real number and |δ|>0\displaystyle|\delta|>0 is suitable small.

Denote

ψδ(ξδ(X))=ψ(X).\text{$\displaystyle\psi_{\delta}(\xi_{\delta}(X))=\psi(X)$}.

It’s easy to verify that ψδK\displaystyle\psi_{\delta}\in K and

(Dξδ(X))1=(E+δηEδDη)(detDξδ)1anddetDξδ=1+δη+o(δ),(D\xi_{\delta}(X))^{-1}=(E+\delta\nabla\cdot\eta E-\delta D\eta)(\text{det}D\xi_{\delta})^{-1}\quad\text{and}\quad\text{det}D\xi_{\delta}=1+\delta\nabla\cdot\eta+o(\delta),

where E\displaystyle E is the identity matrix. Then we have

0Ω(|ψδ|2+F(ψδ)+λ2I{ψδ>0})𝑑YΩ(|ψ|2+F(ψ)+λ2I{ψ>0})𝑑X=Ω(|ψ(Dξδ)1|2+F(ψ)+λ2(ξδ(X))I{ψ>0})detDξδ𝑑XΩ(|ψ|2+F(ψ)+λ2I{ψ>0})𝑑X=δΩ{ψ>0}((|ψ|2+F(ψ)+λ2)η2ψDηψ+λ2η)𝑑X+o(δ).\begin{array}[]{rl}0\leq&\int_{\Omega}\left(|\nabla\psi_{\delta}|^{2}+F(\psi_{\delta})+\lambda^{2}I_{\{\psi_{\delta}>0\}}\right)dY-\int_{\Omega}\left(|\nabla\psi|^{2}+F(\psi)+\lambda^{2}I_{\{\psi>0\}}\right)dX\\ =&\int_{\Omega}\left(|\nabla\psi(D\xi_{\delta})^{-1}|^{2}+F(\psi)+\lambda^{2}(\xi_{\delta}(X))I_{\{\psi>0\}}\right)\text{det}D\xi_{\delta}dX\\ &-\int_{\Omega}\left(|\nabla\psi|^{2}+F(\psi)+\lambda^{2}I_{\{\psi>0\}}\right)dX\\ =&\delta\int_{\Omega\cap\{\psi>0\}}\left((|\nabla\psi|^{2}+F(\psi)+\lambda^{2})\nabla\cdot\eta-2\nabla\psi\cdot D\eta\cdot\nabla\psi+\nabla\lambda^{2}\cdot\eta\right)dX+o(\delta).\end{array} (2.13)

Due to the arbitrariness of δ\displaystyle\delta, the linear term of (2.13) in δ\displaystyle\delta has to vanish, and this gives that

0=Ω{ψ>0}((|ψ|2+F(ψ)+λ2)η2ψDηψ+λ2η)𝑑X=Ω{ψ>0}((|ψ|2+λ2+F(ψ))η2(ηψ)ψ)𝑑X=limε0Ω{ψ>ε}((|ψ|2+λ2+F(ψ))η2(ηψ)ψ)νε𝑑S=limε0Ω{ψ>ε}(λ2(X)|ψ|2+F(ψ))ηνε𝑑S.\begin{array}[]{rl}0=&\int_{\Omega\cap\{\psi>0\}}\left((|\nabla\psi|^{2}+F(\psi)+\lambda^{2})\nabla\cdot\eta-2\nabla\psi\cdot D\eta\cdot\nabla\psi+\nabla\lambda^{2}\cdot\eta\right)dX\\ =&\int_{\Omega\cap\{\psi>0\}}\nabla\cdot\left((|\nabla\psi|^{2}+\lambda^{2}+F(\psi))\eta-2(\eta\cdot\nabla\psi)\nabla\psi\right)dX\\ =&\lim_{\varepsilon\downarrow 0}\int_{\Omega\cap\partial\{\psi>\varepsilon\}}\left((|\nabla\psi|^{2}+\lambda^{2}+F(\psi))\eta-2(\eta\cdot\nabla\psi)\nabla\psi\right)\cdot\nu_{\varepsilon}dS\\ =&\lim_{\varepsilon\downarrow 0}\int_{\Omega\cap\partial\{\psi>\varepsilon\}}\left(\lambda^{2}(X)-|\nabla\psi|^{2}+F(\psi)\right)\eta\cdot\nu_{\varepsilon}dS.\end{array} (2.14)

Hence, we obtain the conclusion (2.11). To prove (2.12), define ξδ(X)=X+δη(X)\displaystyle\xi_{\delta}(X)=X+\delta\eta(X) for any η(X)(C1(Ω{ψ>0}))2\displaystyle\eta(X)\in\left(C^{1}(\Omega\cap\partial\{\psi>0\})\right)^{2} with η=0\displaystyle\eta=0 on (Ω{ψ>0})l\displaystyle(\Omega\cap\partial\{\psi>0\})\setminus l and ην0\displaystyle\eta\cdot\nu\leq 0 on l\displaystyle l, where δ>0\displaystyle\delta>0 is suitable small and ν\displaystyle\nu is the outer normal vector of l\displaystyle l.

Along the similar arguments in (2.13) and (2.14), we have

0Ω{ψ>0}((|ψ|2+F(ψ)+λ2)η2ψDηψ+λ2(X)η)𝑑X=Ω{ψ>0}((|ψ|2+λ2+F(ψ))η2(ηψ)ψ)𝑑X=l(λ2(X)|ψ|2+F(ψ))ην𝑑S.\begin{array}[]{rl}0\leq&\int_{\Omega\cap\{\psi>0\}}\left((|\nabla\psi|^{2}+F(\psi)+\lambda^{2})\nabla\cdot\eta-2\nabla\psi\cdot D\eta\cdot\nabla\psi+\nabla\lambda^{2}(X)\cdot\eta\right)dX\\ =&\int_{\Omega\cap\{\psi>0\}}\nabla\cdot\left((|\nabla\psi|^{2}+\lambda^{2}+F(\psi))\eta-2(\eta\cdot\nabla\psi)\nabla\psi\right)dX\\ =&\int_{l}\left(\lambda^{2}(X)-|\nabla\psi|^{2}+F(\psi)\right)\eta\cdot\nu dS.\end{array} (2.15)

Since lC1,α\displaystyle l\in C^{1,\alpha}, ψ\displaystyle\psi is also C1\displaystyle C^{1} up to l\displaystyle l. Therefore, we complete the proof of (2.12) by using (2.15).

Remark 2.2.

Proposition 2.2 implies that if the free boundary Γ\displaystyle\Gamma is C1\displaystyle C^{1} and ψ\displaystyle\psi is C1\displaystyle C^{1} in {ψ>0}\displaystyle\{\psi>0\} uniformly up to the free boundary, then

|ψ(X)|=λ(X)on the free boundary Γ.|\nabla\psi(X)|=\lambda(X)\ \ \text{on the free boundary $\displaystyle\Gamma$}.

Next, we will give the estimate for the Lipschitz norm for ψ\displaystyle\psi near the free boundary, this is the principal part in establishing the Lipschitz continuity of the minimizer ψ\displaystyle\psi. Denote

Ω+=Ω{ψ>0}andΩ0=Ω{ψ=0}.\Omega_{+}=\Omega\cap\{\psi>0\}\ \ \text{and}\ \ \Omega_{0}=\Omega\cap\{\psi=0\}.
Lemma 2.3.

Suppose X0=(x0,y0)Ω\displaystyle X_{0}=(x_{0},y_{0})\in\Omega with d(X0)<12min{dist(X0,Ω),1}\displaystyle d(X_{0})<\frac{1}{2}\min\{\text{dist}(X_{0},\partial\Omega),1\}, where d(X0)=dist(X0,Ω0)\displaystyle d(X_{0})=\text{dist}(X_{0},\Omega_{0}). Then there exists a positive constant C=C(Λ)\displaystyle C=C(\Lambda) depending only on Λ\displaystyle\Lambda, such that

ψ(X0)C(λ2+Λ)d(X0),\psi(X_{0})\leq C(\lambda_{2}+\Lambda)d(X_{0}),

where Λ\displaystyle\Lambda is defined in (2.1).

Proof.

We assume that

ψ(X0)>Md(X0)>0,\psi(X_{0})>Md(X_{0})>0, (2.16)

and derive an upper bound on M\displaystyle M. Denote d=d(X0)\displaystyle d=d(X_{0}) for simplicity. By scaling ψd(X)=ψ(X0+d(XX0))d\displaystyle\psi_{d}(X)=\frac{\psi(X_{0}+d(X-X_{0}))}{d} and λd(X)=λ(X0+d(XX0))\displaystyle\lambda_{d}(X)=\lambda(X_{0}+d(X-X_{0})) with XB1(X0)\displaystyle X\in B_{1}(X_{0}), we have

Δψd+df(dψd)=0\displaystyle\Delta\psi_{d}+df(d\psi_{d})=0\ in B1(X0)\displaystyle B_{1}(X_{0}).

On the other hand, it is easy to check that

|df(dψd)|d|f(0)|+d2maxξ0|f(ξ)|ψddΛ+d2Λψd.|df(d\psi_{d})|\leq d|f(0)|+d^{2}\max_{\xi\geq 0}|f^{\prime}(\xi)|\psi_{d}\leq d\Lambda+d^{2}\Lambda\psi_{d}.

Thanks to Harnack’s inequality in [31] and (2.16), we have

infB34(X0)ψdcMCdΛcMCΛ,\inf_{B_{\frac{3}{4}}(X_{0})}\psi_{d}\geq cM-Cd\Lambda\geq cM-C\Lambda, (2.17)

where the constants c\displaystyle c and C\displaystyle C depend on Λ\displaystyle\Lambda.

Let Y\displaystyle Y be a point in B1(X0){ψd=0}\displaystyle\partial B_{1}(X_{0})\cap\{\psi_{d}=0\}. Define a function ϕ\displaystyle\phi as follows

Δϕ+df(dϕ)=0inB1(Y),andϕ=ψdonB1(Y).\Delta\phi+df(d\phi)=0\ \ \ \text{in}~~B_{1}(Y),\ \ \text{and}\ \ \ \phi=\psi_{d}\ \ \text{on}\ \ \partial B_{1}(Y).

Since ψd\displaystyle\psi_{d} is a minimizer, by virtue of the convexity of F(t)\displaystyle F(t), one has

0B1(Y)|ϕ|2|ψd|2+F(dϕ)F(dψd)+λd2(I{ϕ>0}I{ψd>0})dXB1(Y)|(ϕψd)|2+2(ϕψd)ϕdX+B1(Y)𝑑F(dϕ)(ϕψd)+λd2I{ψd=0}dXB1(Y)|(ϕψd)|2𝑑X+λ22B1(Y)I{ψd=0}𝑑X,\begin{array}[]{rl}0\leq&\int_{B_{1}(Y)}|\nabla\phi|^{2}-|\nabla\psi_{d}|^{2}+F(d\phi)-F(d\psi_{d})+\lambda_{d}^{2}(I_{\{\phi>0\}}-I_{\{\psi_{d}>0\}})dX\\ \leq&\int_{B_{1}(Y)}-|\nabla(\phi-\psi_{d})|^{2}+2\nabla(\phi-\psi_{d})\cdot\nabla\phi dX\\ &+\int_{B_{1}(Y)}dF^{\prime}(d\phi)(\phi-\psi_{d})+\lambda_{d}^{2}I_{\{\psi_{d}=0\}}dX\\ \leq&-\int_{B_{1}(Y)}|\nabla(\phi-\psi_{d})|^{2}dX+\lambda_{2}^{2}\int_{B_{1}(Y)}I_{\{\psi_{d}=0\}}dX,\end{array}

and thus

B1(Y)|(ϕψd)|2𝑑Xλ22B1(Y)I{ψd=0}𝑑X.\begin{array}[]{rl}\int_{B_{1}(Y)}|\nabla(\phi-\psi_{d})|^{2}dX\leq\lambda_{2}^{2}\int_{B_{1}(Y)}I_{\{\psi_{d}=0\}}dX.\end{array} (2.18)

It follows from the statement (2) in Lemma 2.1 that Δψd+df(dψd)0\displaystyle\Delta\psi_{d}+df(d\psi_{d})\geq 0 in B1(Y)\displaystyle B_{1}(Y), and the maximum principle gives that

ϕψdinB1(Y).\phi\geq\psi_{d}\ \ \text{in}\ \ B_{1}(Y). (2.19)

Then it follows from (2.17) and (2.19) that

ϕ(X)ψd(X)cMCΛ>0inB34(X0)B1(Y).\phi(X)\geq\psi_{d}(X)\geq cM-C\Lambda>0\ \ \text{in}\ \ B_{\frac{3}{4}}(X_{0})\cap B_{1}(Y).

Using Harnack’s inequality in [31] again, we have

ϕ(X)cMCΛ=C0>0inB12(Y).\phi(X)\geq cM-C\Lambda=C_{0}>0\ \ \text{in}\ \ B_{\frac{1}{2}}(Y). (2.20)

We take Y=0\displaystyle Y=0 for simplicity and introduce a function

φ(X)=C0(eβ|X|2eβ).\varphi(X)=C_{0}\left(e^{-\beta|X|^{2}}-e^{-\beta}\right).

Then a direct computation gives that

Δφ+df(dφ)C0eβ|X|2(4β(β|X|21)d2Λ)+C0d2Λeβ+df(0)>0,\Delta\varphi+df(d\varphi)\geq C_{0}e^{-\beta|X|^{2}}(4\beta(\beta|X|^{2}-1)-d^{2}\Lambda)+C_{0}d^{2}\Lambda e^{-\beta}+df(0)>0,

for 12<|X|<1\displaystyle\frac{1}{2}<|X|<1, and sufficiently large β>0\displaystyle\beta>0. This gives that

Δ(φϕ)+df(dφ)df(dϕ)>0inB1(0)B12(0),\Delta(\varphi-\phi)+df(d\varphi)-df(d\phi)>0\ \ \text{in}\ \ B_{1}(0)\setminus B_{\frac{1}{2}}(0),

provided that β\displaystyle\beta is sufficiently large. It follows from the maximum principle that

ϕ(X)>φ(X)=C0(eβ|X|2eβ)cC0(1|X|)inB1(0)B12(0),\phi(X)>\varphi(X)=C_{0}\left(e^{-\beta|X|^{2}}-e^{-\beta}\right)\geq cC_{0}(1-|X|)\ \ \text{in}\ \ B_{1}(0)\setminus B_{\frac{1}{2}}(0),

which implies that

ϕ(X)>(cMCΛ)(1|X|)inB1(0)B12(0).\phi(X)>(cM-C\Lambda)(1-|X|)\ \ \text{in}\ \ B_{1}(0)\setminus B_{\frac{1}{2}}(0).

This together with (2.20) gives that

ϕ(X)(cMCΛ)(1|X|)inB1(0).\phi(X)\geq(cM-C\Lambda)(1-|X|)\ \ \text{in}\ \ B_{1}(0). (2.21)

With the aid of (2.18) and (2.21), it follows from the similar arguments in Lemma 3.2 in [2] and Lemma 2.2 in [6] that

(cMCΛ)2CmaxXΩλ2(X)=Cλ22,(cM-C\Lambda)^{2}\leq C\max_{X\in\Omega}\lambda^{2}(X)=C\lambda_{2}^{2},

that is

MC(λ2+Λ),M\leq C(\lambda_{2}+\Lambda),

where the constant C=C(Λ)\displaystyle C=C(\Lambda) depends only on Λ\displaystyle\Lambda.

With the aid of Lemma 2.3, we will obtain the Lipschitz continuity of ψ\displaystyle\psi, which plays an essential part for constructing a Radon measure. To obtain the Lipschitz continuity, we should use the method of Harnack’s inequality in [31].

Theorem 2.4.

(1). The minimizer ψ\displaystyle\psi is Lipschitz continuous in Ω\displaystyle\Omega, namely, ψC0,1(Ω)\displaystyle\psi\in C^{0,1}(\Omega).
(2). For any compact subset D\displaystyle D of Ω\displaystyle\Omega which containing a free boundary point, the Lipschitz coefficient of ψ\displaystyle\psi in D\displaystyle D is estimated by C(λ2+Λ)\displaystyle C(\lambda_{2}+\Lambda), and the constant C\displaystyle C depends only on Λ\displaystyle\Lambda, D\displaystyle D and Ω\displaystyle\Omega.

Proof.

(1). Suppose that d(X0)<13min{dist(X0,Ω),1}\displaystyle d(X_{0})<\frac{1}{3}\min\{\text{dist}(X_{0},\partial\Omega),1\}, where d(X0)=dist(X0,Ω0)\displaystyle d(X_{0})=\text{dist}(X_{0},\Omega_{0}). Set

ψ~(X~)=ψ(X0+d(X0)X~)d(X0).\tilde{\psi}(\tilde{X})=\frac{\psi(X_{0}+d(X_{0})\tilde{X})}{d(X_{0})}.

By virtue of Lemma 2.3, one has

ψ~(X~)C(λ2+Λ)inB1(0),\tilde{\psi}(\tilde{X})\leq C(\lambda_{2}+\Lambda)\ \ \text{in}\ \ B_{1}(0),

where the constant C=C(Λ)\displaystyle C=C(\Lambda) depends only on Λ\displaystyle\Lambda.

Since Δψ~+d(X0)f(d(X0)ψ~)=0\displaystyle\Delta\tilde{\psi}+d(X_{0})f(d(X_{0})\tilde{\psi})=0 in B1(0)\displaystyle B_{1}(0), it follows from the elliptic estimate for ψ~\displaystyle\tilde{\psi} that

|ψ~(0)|C(ψ~C0(B1(0))+f(d(X0)ψ~)L(B1(0))d(X0))C(1+d2(X0)|f(0)|+d2(X0)Λ)ψ~C0(B1(0))C(λ2+Λ),\begin{array}[]{rl}|\nabla\tilde{\psi}(0)|\leq&C(\|\tilde{\psi}\|_{C^{0}(B_{1}(0))}+\|f(d(X_{0})\tilde{\psi})\|_{L^{\infty}(B_{1}(0))}d(X_{0}))\\ \leq&C(1+d^{2}(X_{0})|f(0)|+d^{2}(X_{0})\Lambda)\|\tilde{\psi}\|_{C^{0}(B_{1}(0))}\\ \leq&C(\lambda_{2}+\Lambda),\end{array}

which implies that

|ψ(X0)|=|ψ~(0)|C(λ2+Λ).|\nabla\psi(X_{0})|=|\nabla\tilde{\psi}(0)|\leq C(\lambda_{2}+\Lambda).

Thus, for any compact subset D\displaystyle D of Ω\displaystyle\Omega, |ψ(X)|\displaystyle|\nabla\psi(X)| is bounded in D{ψ>0}U\displaystyle D\cap\{\psi>0\}\cap U, where U\displaystyle U is a small neighborhood of the free boundary. It follows from Lemma 2.1 that ψC2,α\displaystyle\psi\in C^{2,\alpha} in Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\}, this together with ψ=0\displaystyle\nabla\psi=0 a.e. in Ω{ψ=0}\displaystyle\Omega\cap\{\psi=0\} gives that ψC0,1(Ω)\displaystyle\psi\in C^{0,1}(\Omega).

(2). Consider any connected domains DGΩ\displaystyle D\subset\subset G\subset\subset\Omega and D\displaystyle D contains at least one free boundary point. Let r0=15min{dist(G,Ω),1}\displaystyle r_{0}=\frac{1}{5}\min\{\text{dist}(G,\partial\Omega),1\} and XGΩ+\displaystyle X\in G\cap\Omega_{+}. Since D\displaystyle D contains some free boundary points, and we have that G\displaystyle G is not contained in Ω+\displaystyle\Omega_{+}, then we can find some finite points X1,X2,,Xm\displaystyle X_{1},X_{2},...,X_{m} in G\displaystyle G (m\displaystyle m depends only on G\displaystyle G and Ω\displaystyle\Omega), such that

X0=X,XkBr0(Xk1)fork=1,,m,X_{0}=X,\ \ X_{k}\in B_{r_{0}}(X_{k-1})\ \text{for}\ \ k=1,...,m,

and

B2r0(Xk)Ω+fork=1,,m1,and B2r0(Xm) is not contained in Ω+.B_{2r_{0}}(X_{k})\subset\Omega_{+}\ \text{for}\ \ k=1,...,m-1,\text{and $\displaystyle B_{2r_{0}}(X_{m})$ is not contained in $\displaystyle\Omega_{+}$}.

It follows from Lemma 2.1 and Lemma 2.3 that

ψ(Xm)C(λ2+Λ)r0andΔψ+f(ψ)=0inB2r0(Xk)fork=1,,m1.\psi(X_{m})\leq C(\lambda_{2}+\Lambda)r_{0}\ \ \text{and}\ \ \Delta\psi+f(\psi)=0\ \ \text{in}\ \ B_{2r_{0}}(X_{k})\ \text{for}\ \ k=1,...,m-1.

By virtue of Harnack’s inequality in [31], we have

ψ(Xk1)Cψ(Xk)+CΛr0fork=1,,m1,\psi(X_{k-1})\leq C\psi(X_{k})+C\Lambda r_{0}\ \text{for}\ \ k=1,...,m-1,

and

ψ(Xm1)CinfXB2r0(Xm1)ψ(X)+CΛr0Cψ(Xm)+CΛr0.\psi(X_{m-1})\leq C\inf_{X\in B_{2r_{0}}(X_{m-1})}\psi(X)+C\Lambda r_{0}\leq C\psi(X_{m})+C\Lambda r_{0}.

Inductively, we have

ψ(X)=ψ(X0)C(λ2+Λ)for all XG.\psi(X)=\psi(X_{0})\leq C(\lambda_{2}+\Lambda)\ \ \text{for all $\displaystyle X\in G$}. (2.22)

For any XDΩ+\displaystyle X\in D\cap\Omega_{+} and d(X)=dist(X,Ω0)\displaystyle d(X)=\text{dist}(X,\Omega_{0}). We consider the following two cases.

Case 1. d(X)12min{dist(D,G),1}\displaystyle d(X)\geq\frac{1}{2}\min\{dist(D,\partial G),1\}, by using elliptic estimate and (2.22), one has

|ψ(X)|C(supGψ+supG|f(ψ)|)C(λ2+Λ),|\nabla\psi(X)|\leq C\left(\sup_{G}\psi+\sup_{G}|f(\psi)|\right)\leq C(\lambda_{2}+\Lambda),

where C\displaystyle C is a constant depending only on Λ\displaystyle\Lambda, D,G\displaystyle D,G and Ω\displaystyle\Omega.

Case 2. d(X)<12min{dist(D,G),1}\displaystyle d(X)<\frac{1}{2}\min\{dist(D,\partial G),1\}, it follows from Lemma 2.3 that

|ψ(X)|C(λ2+Λ),|\nabla\psi(X)|\leq C(\lambda_{2}+\Lambda),

where C\displaystyle C is a constant depending only on Λ\displaystyle\Lambda.

Hence, we complete the proof of Theorem 2.4.

2.3. Non-degeneracy of the minimizer

As a consequence of Theorem 2.4, we will obtain the non-degeneracy lemma of the minimizer ψ\displaystyle\psi in this subsection.

Lemma 2.5.

Let ψ\displaystyle\psi be a minimizer, for any compact subset D\displaystyle D of Ω\displaystyle\Omega, there exists a positive constant C\displaystyle C^{*} (depending only on Λ,D\displaystyle\Lambda,D and Ω\displaystyle\Omega), such that for any disc Br(X0)D\displaystyle B_{r}(X_{0})\subset D,

1rBr(X0)ψ𝑑SC(λ2+Λ)\frac{1}{r}\fint_{\partial B_{r}(X_{0})}\psi dS\geq C^{*}(\lambda_{2}+\Lambda) (2.23)

implies that

ψ(x,y)>0inBr(X0).\psi(x,y)>0\ \ ~~~~\text{in}\ ~~B_{r}(X_{0}).
Proof.

Suppose that there exists a free boundary point XBr(X0)Ω0\displaystyle X^{\prime}\in B_{r}(X_{0})\cap\Omega_{0}. It follows from Theorem 2.4 that |ψ|C(λ2+Λ)\displaystyle|\nabla\psi|\leq C(\lambda_{2}+\Lambda), where C\displaystyle C is a constant depending only on Λ,D\displaystyle\Lambda,D and Ω\displaystyle\Omega. Then we have

ψ(X)=ψ(X)ψ(X)C(λ2+Λ)rinB¯r,\psi(X)=\psi(X)-\psi(X^{\prime})\leq C(\lambda_{2}+\Lambda)r\ \ \text{in}\ \ \overline{B}_{r},

which contradicts to (2.23), provided that C\displaystyle C^{*} is sufficiently large. ∎

The non-degeneracy lemma will be stated in the following.

Lemma 2.6.

For any κ(0,1)\displaystyle\kappa\in(0,1), there exists a positive constant cκ=c(κ)\displaystyle c^{*}_{\kappa}=c^{*}(\kappa), such that for any disc Br(X0)Ω\displaystyle B_{r}(X_{0})\subset\Omega with r<1Λmin{cκλ1,1}\displaystyle r<\frac{1}{\Lambda}\min\{c_{\kappa}^{*}\lambda_{1},1\},

1rBr(X0)ψ𝑑Scκλ1\frac{1}{r}\fint_{\partial B_{r}(X_{0})}\psi dS\leq c_{\kappa}^{*}\lambda_{1} (2.24)

implies that

ψ(x,y)0inBκr(X0).\psi(x,y)\equiv 0~~~~\text{in}~~B_{\kappa r}(X_{0}).
Proof.

Without loss of generality, we assume that X0=0\displaystyle X_{0}=0. Set ψr(X)=ψ(rX)r\displaystyle\psi_{r}(X)=\frac{\psi(rX)}{r} and λr(X)=λ(rX)\displaystyle\lambda_{r}(X)=\lambda(rX), we have

Δψr+rf(rψr)=0inΩr{ψr>0},\Delta\psi_{r}+rf(r\psi_{r})=0\ \ \ \text{in}~~\Omega_{r}\cap\{\psi_{r}>0\},

where Ωr={XrXΩ}\displaystyle\Omega_{r}=\{X\mid rX\in\Omega\}. Denote

ε=supXBκ(0)ψr(X).\varepsilon=\sup_{X\in B_{\sqrt{\kappa}}(0)}\psi_{r}(X).

Since Δψrrf(rψr)rf(0)Λr\displaystyle\Delta\psi_{r}\geq-rf(r\psi_{r})\geq-rf(0)\geq-\Lambda r in Ω\displaystyle\Omega, it follows from the maximum principle that

ψr(X)Λr1|X|24+1|X|22πB1(0)ψr(Y)|XY|2𝑑SY\psi_{r}(X)\leq\Lambda r\frac{1-|X|^{2}}{4}+\frac{1-|X|^{2}}{2\pi}\int_{\partial B_{1}(0)}\frac{\psi_{r}(Y)}{|X-Y|^{2}}dS_{Y}

for any XBκ(0)\displaystyle X\in B_{\sqrt{\kappa}}(0), which implies that

ε=supXBκ(0)ψr(X)Λr+C(κ)B1(0)ψr(Y)𝑑SY.\varepsilon=\sup_{X\in B_{\sqrt{\kappa}}(0)}\psi_{r}(X)\leq\Lambda r+C(\kappa)\fint_{\partial B_{1}(0)}\psi_{r}(Y)dS_{Y}. (2.25)

Define a function ϕ\displaystyle\phi solving the following problem

{Δϕ+rf(rϕ)=0inBκ(0)Bκ(0),ϕ=0inBκ(0),ϕ=εoutsideBκ(0).\left\{\begin{array}[]{ll}&\Delta\phi+rf(r\phi)=0~~~~\text{in}\ \ \ B_{\sqrt{\kappa}}(0)\setminus B_{\kappa}(0),\\ &\phi=0~~~~\text{in}\ \ \ B_{\kappa}(0),\ \ \ \ \ \phi=\varepsilon~~~~\text{outside}\ \ \ B_{\sqrt{\kappa}}(0).\end{array}\right.

It is easy to check that

0Bκ(0)|min{ψr,ϕ}|2+F(rmin{ψr,ϕ})+λr2I{min{ψr,ϕ}>0}dXBκ(0)|ψr|2+F(rψr)+λr2I{ψr>0}dX,\begin{array}[]{rl}0\leq&\int_{B_{\sqrt{\kappa}}(0)}|\nabla\min\{\psi_{r},\phi\}|^{2}+F(r\min\{\psi_{r},\phi\})+\lambda_{r}^{2}I_{\{\min\{\psi_{r},\phi\}>0\}}dX\\ &-\int_{B_{\sqrt{\kappa}}(0)}|\nabla\psi_{r}|^{2}+F(r\psi_{r})+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX,\end{array}

which yields that

Bκ(0)|ψr|2+F(rψr)+λr2I{ψr>0}dXBκ(0)Bκ(0)|min{ψr,ϕ}|2+F(rmin{ψr,ϕ})|ψr|2F(rψr)dXBκ(0)Bκ(0)|min{ϕψr,0}|2+2ϕmin{ϕψr,0}+rF(rϕ)min{ϕψr,0}dX2Bκ(0)min{ϕψr,0}ϕν𝑑S,\begin{array}[]{rl}&\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}+F(r\psi_{r})+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX\\ \leq&\int_{B_{\sqrt{\kappa}}(0)\setminus B_{\kappa}(0)}|\nabla\min\{\psi_{r},\phi\}|^{2}+F(r\min\{\psi_{r},\phi\})-|\nabla\psi_{r}|^{2}-F(r\psi_{r})dX\\ \leq&\int_{B_{\sqrt{\kappa}}(0)\setminus B_{\kappa}(0)}-|\nabla\min\{\phi-\psi_{r},0\}|^{2}+2\nabla\phi\cdot\nabla\min\{\phi-\psi_{r},0\}\\ &+rF^{\prime}(r\phi)\min\{\phi-\psi_{r},0\}dX\\ \leq&2\int_{\partial B_{\kappa}(0)}\min\{\phi-\psi_{r},0\}\frac{\partial\phi}{\partial\nu}dS,\end{array} (2.26)

where ν\displaystyle\nu is the outer normal vector, we have used the fact

Bκ(0)Bκ(0)I{min{ψr,ϕ}>0}I{ψr>0}dX0,\int_{B_{\sqrt{\kappa}}(0)\setminus B_{\kappa}(0)}I_{\{\min\{\psi_{r},\phi\}>0\}}-I_{\{\psi_{r}>0\}}dX\leq 0,

due to the fact {min{ψr,ϕ}>0}{ψr>0}\displaystyle\{\min\{\psi_{r},\phi\}>0\}\subset\{\psi_{r}>0\}.

On the other hand, by virtue of the convexity of F(t)\displaystyle F(t), we have that

F(rψr)F(0)rψr2Λrψr,F(r\psi_{r})\geq F^{\prime}(0)r\psi_{r}\geq-2\Lambda r\psi_{r},

which gives that

Bκ(0)|ψr|22Λrψr+λr2I{ψr>0}dXBκ(0)|ψr|2+F(rψr)+λr2I{ψr>0}dX.\begin{array}[]{rl}&\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}-2\Lambda r\psi_{r}+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX\leq\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}+F(r\psi_{r})+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX.\end{array} (2.27)

The standard elliptic estimates give that

supXBκ(0)|ϕ(X)|C(supXBκ(0)ϕ(X)+rf(rϕ)L(Bκ(0)))C(Λ,κ)(ε+Λr).\sup_{X\in\partial B_{\kappa}(0)}|\nabla\phi(X)|\leq C\left(\sup_{X\in B_{\sqrt{\kappa}}(0)}\phi(X)+r\|f(r\phi)\|_{L^{\infty}(B_{\sqrt{\kappa}}(0))}\right)\leq C(\Lambda,\kappa)(\varepsilon+\Lambda r). (2.28)

Furthermore, with the aid of the trace theorem, it follows from (2.26) - (2.28) that

Bκ(0)|ψr|2+λr2I{ψr>0}dX2Bκ(0)min{ϕψr,0}ϕν𝑑S+2Bκ(0)Λrψr𝑑XC(Λ,κ)ε0(Bκ(0)ψr𝑑S+Bκ(0)ψr𝑑X)C(Λ,κ)ε0(Bκ(0)ψrI{ψr>0}𝑑X+CBκ(0)|ψr|I{ψr>0}𝑑X)C(Λ,κ)ε0(ε0λ12Bκ(0)I{ψr>0}𝑑X+Cλ1Bκ(0)|ψr|2+λr2I{ψr>0}dX)C(Λ,κ)(ε02λ12+ε0λ1)Bκ(0)|ψr|2+λr2I{ψr>0}dX,\begin{array}[]{rl}&\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX\\ \leq&2\int_{\partial B_{\kappa}(0)}\min\{\phi-\psi_{r},0\}\frac{\partial\phi}{\partial\nu}dS+2\int_{B_{\kappa}(0)}\Lambda r\psi_{r}dX\\ \leq&C(\Lambda,\kappa)\varepsilon_{0}\left(\int_{\partial B_{\kappa}(0)}\psi_{r}dS+\int_{B_{\kappa}(0)}\psi_{r}dX\right)\\ \leq&C(\Lambda,\kappa)\varepsilon_{0}\left(\int_{B_{\kappa}(0)}\psi_{r}I_{\{\psi_{r}>0\}}dX+C\int_{B_{\kappa}(0)}|\nabla\psi_{r}|I_{\{\psi_{r}>0\}}dX\right)\\ \leq&C(\Lambda,\kappa)\varepsilon_{0}\left(\frac{\varepsilon_{0}}{\lambda_{1}^{2}}\int_{B_{\kappa}(0)}I_{\{\psi_{r}>0\}}dX+\frac{C}{\lambda_{1}}\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX\right)\\ \leq&C(\Lambda,\kappa)\left(\frac{\varepsilon^{2}_{0}}{\lambda_{1}^{2}}+\frac{\varepsilon_{0}}{\lambda_{1}}\right)\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX,\end{array} (2.29)

where ε0=ε+Λr\displaystyle\varepsilon_{0}=\varepsilon+\Lambda r. By virtue of (2.24) and (2.25), one has

ε0λ1=ε+Λrλ1C(Λ,κ)(1λ1B1(0)ψr𝑑S+Λrλ1)<C(Λ,κ)cκ.\frac{\varepsilon_{0}}{\lambda_{1}}=\frac{\varepsilon+\Lambda r}{\lambda_{1}}\leq C(\Lambda,\kappa)\left(\frac{1}{\lambda_{1}}\fint_{\partial B_{1}(0)}\psi_{r}dS+\frac{\Lambda r}{\lambda_{1}}\right)<C(\Lambda,\kappa)c_{\kappa}^{*}.

This implies that ε0λ1\displaystyle\frac{\varepsilon_{0}}{\lambda_{1}} is small enough, provided that cκ\displaystyle c_{\kappa}^{*} is small enough. This together with (2.29) implies that

Bκ(0)|ψr|2+λr2I{ψr>0}dX=0,\int_{B_{\kappa}(0)}|\nabla\psi_{r}|^{2}+\lambda_{r}^{2}I_{\{\psi_{r}>0\}}dX=0,\ \

for sufficiently small cκ\displaystyle c_{\kappa}^{*}. This implies that ψ=0\displaystyle\psi=0 in Bκ(0)\displaystyle B_{\kappa}(0), if cκ\displaystyle c_{\kappa}^{*} is small enough.

Remark 2.3.

It should be noted that Lemma 2.5 remains valid, provided that Br(X0)\displaystyle B_{r}(X_{0}) is not contained in Ω\displaystyle\Omega and ψ=0\displaystyle\psi=0 on Br(X0)Ω\displaystyle B_{r}(X_{0})\cap\partial\Omega.

Finally, we give the density estimate of the free boundary point, which gives that 2(D{ψ>0})=0\displaystyle\mathcal{L}^{2}(D\cap\partial\{\psi>0\})=0 for any compact subset D\displaystyle D of Ω\displaystyle\Omega.

Lemma 2.7.

Let G\displaystyle G be a compact subset of Ω\displaystyle\Omega, there exists a positive constant c(0,1)\displaystyle c\in(0,1), such that for any disc Br=Br(X0)G\displaystyle B_{r}=B_{r}(X_{0})\subset G with X0Γ\displaystyle X_{0}\in\Gamma and small r\displaystyle r,

c<2(Br{ψ>0})2(Br)<1c,c<\frac{\mathcal{L}^{2}(B_{r}\cap\{\psi>0\})}{\mathcal{L}^{2}(B_{r})}<1-c, (2.30)

where 2\displaystyle\mathcal{L}^{2} is the two-dimensional Lebesgue measure.

Proof.

Step 1. Without loss of generality, we assume that X0=0\displaystyle X_{0}=0. It follows from Lemma 2.6 that there exists a point YBr\displaystyle Y\in\partial B_{r}, such that ψ(Y)>cr\displaystyle\psi(Y)>cr. Set φ(X)=ψ(X)+Λ|X|24\displaystyle\varphi(X)=\psi(X)+\frac{\Lambda|X|^{2}}{4}, one has

Δφ=Δψ+Λf(0)f(ψ)0in Br,\Delta\varphi=\Delta\psi+\Lambda\geq f(0)-f(\psi)\geq 0\ \ \ \text{in $\displaystyle B_{r}$},

which implies that

Λr2κ+1κrBκr(Y)ψ(Z)𝑑SZ1κrBκr(Y)φ(Z)𝑑SZφ(Y)κrcκ,\frac{\Lambda r}{2\kappa}+\frac{1}{\kappa r}\fint_{\partial B_{\kappa r}(Y)}\psi(Z)dS_{Z}\geq\frac{1}{\kappa r}\fint_{\partial B_{\kappa r}(Y)}\varphi(Z)dS_{Z}\geq\frac{\varphi(Y)}{\kappa r}\geq\frac{c}{\kappa},

for small κ>0\displaystyle\kappa>0, that is

1κrBκr(Y)ψ(Z)𝑑SZcΛr2κ>c4κif r is small enough.\frac{1}{\kappa r}\fint_{\partial B_{\kappa r}(Y)}\psi(Z)dS_{Z}\geq\frac{c-\Lambda r}{2\kappa}>\frac{c}{4\kappa}\ \ \text{if $\displaystyle r$ is small enough}.

By virtue of non-degeneracy Lemma 2.6, one has

ψ>0\displaystyle\psi>0 in Bκr(Y)\displaystyle B_{\kappa r}(Y),

this gives the left-hand side of (2.30).

Step 2. Define a function ϕ\displaystyle\phi solving the following problem

Δϕ+f(ϕ)=0inBr,ϕ=ψoutsideBr.\Delta\phi+f(\phi)=0~~~~\text{in}~~B_{r},\ \ \phi=\psi~~~~~~\text{outside}~~B_{r}.

The maximum principle gives that ψϕΨ\displaystyle\psi\leq\phi\leq\Psi in Br\displaystyle B_{r}, and thus ϕK\displaystyle\phi\in K, where Ψ\displaystyle\Psi is defined in (2.3).

Then we have

0J(ϕ)J(ψ)Br|(ϕψ)|2dX+λ22BrI{ψ=0}𝑑X,\begin{array}[]{rl}0\leq J(\phi)-J(\psi)\leq\int_{B_{r}}-|\nabla(\phi-\psi)|^{2}dX+\lambda_{2}^{2}\int_{B_{r}}I_{\{\psi=0\}}dX,\end{array}

which together with Poincaré’s inequality and Hölder inequality implies that

λ22BrI{ψ=0}𝑑XBr|(ϕψ)|2𝑑Xcr2Br|ϕψ|2𝑑X.\begin{array}[]{rl}\lambda_{2}^{2}\int_{B_{r}}I_{\{\psi=0\}}dX\geq&\int_{B_{r}}|\nabla(\phi-\psi)|^{2}dX\geq\frac{c}{r^{2}}\int_{B_{r}}|\phi-\psi|^{2}dX.\end{array} (2.31)

For any point YBκr\displaystyle Y\in B_{\kappa r}, since ΔϕM0Λf(0)\displaystyle\Delta\phi\leq M_{0}\Lambda-f(0) with M0=supXΩΨ(X)\displaystyle M_{0}=\sup_{X\in\Omega}\Psi(X), one has

ϕ(Y)r2|Y|22πrBrψ|XY|2𝑑SX+(M0Λf(0))|Y|2r24(1Cκ)Brψ𝑑SXCr2.\begin{array}[]{rl}\phi(Y)\geq&\frac{r^{2}-|Y|^{2}}{2\pi r}\int_{\partial B_{r}}\frac{\psi}{|X-Y|^{2}}dS_{X}+(M_{0}\Lambda-f(0))\frac{|Y|^{2}-r^{2}}{4}\\ \geq&(1-C\kappa)\fint_{\partial B_{r}}\psi dS_{X}-Cr^{2}.\end{array} (2.32)

It follows from (2.32), Lemma 2.3 and Lemma 2.6 that

ϕ(Y)ψ(Y)ϕ(Y)Cκr(1Cκ)Brψ𝑑SXCκrCr2cλ1r,\begin{array}[]{rl}\phi(Y)-\psi(Y)\geq&\phi(Y)-C\kappa r\\ \geq&(1-C\kappa)\fint_{\partial B_{r}}\psi dS_{X}-C\kappa r-Cr^{2}\\ \geq&c\lambda_{1}r,\end{array} (2.33)

provided that κ\displaystyle\kappa and r\displaystyle r are small.

Combining (2.31) and (2.33), we have

λ22BrI{ψ=0}𝑑Xcr2Br|ϕψ|2𝑑Xcr2Bκr|ϕψ|2𝑑Xcκλ12r2,\begin{array}[]{rl}\lambda_{2}^{2}\int_{B_{r}}I_{\{\psi=0\}}dX\geq\frac{c}{r^{2}}\int_{B_{r}}|\phi-\psi|^{2}dX\geq\frac{c}{r^{2}}\int_{B_{\kappa r}}|\phi-\psi|^{2}dX\geq&c_{\kappa}\lambda_{1}^{2}r^{2},\end{array} (2.34)

for small r>0\displaystyle r>0 and small κ>0\displaystyle\kappa>0. This gives the right-hand side of (2.30).

3. Regularity of the free boundary

Based on the Lipschitz continuity and non-degeneracy lemma of the minimizer ψ\displaystyle\psi in Section 2, we will establish the regularity of the free boundary in this section.

3.1. Measure estimate of the free boundary

In this subsection, the main objective is to show that the set Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\} is finite perimeter locally in Ω\displaystyle\Omega.

Set μ=Δψ+f(ψ)\displaystyle\mu=\Delta\psi+f(\psi) and μ0=Δψ+f(ψ)I{ψ>0}\displaystyle\mu_{0}=\Delta\psi+f(\psi)I_{\{\psi>0\}}. First, we will show that μ0\displaystyle\mu_{0} is a Radon measure supported on the free boundary Γ\displaystyle\Gamma.

Lemma 3.1.

μ0=Δψ+f(ψ)I{ψ>0}\displaystyle\mu_{0}=\Delta\psi+f(\psi)I_{\{\psi>0\}} is a positive Radon measure with support in Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\}. Moreover, μ=Δψ+f(ψ)\displaystyle\mu=\Delta\psi+f(\psi) is a Radon measure supported on Ω{ψ=0}\displaystyle\Omega\cap\{\psi=0\} and its singular point is contained in the free boundary.

Proof.

For any ξC0(Ω)\displaystyle\xi\in C_{0}^{\infty}(\Omega) with ξ0\displaystyle\xi\geq 0, denote k(t)=max{min{2t,1},0}\displaystyle k(t)=\max\{\min\{2-t,1\},0\} for any t>0\displaystyle t>0. Then we have

Ωψ(ξk(tψ))dX=ΩψξdX+Ω{ψ1t}ψ(ξk(tψ)ξ)dX=ΩψξdXΩ{ψ>0}f(ψ)ξ𝑑X+Ω{1tψ2t}(2ψt)f(ψ)ξ𝑑X+Ω{ψ2t}f(ψ)ξ𝑑XΩψξdXΩ{ψ>0}f(ψ)ξ𝑑X,\begin{array}[]{rl}\int_{\Omega}\nabla\psi\cdot\nabla(\xi k(t\psi))dX=&\int_{\Omega}\nabla\psi\cdot\nabla\xi dX+\int_{\Omega\cap\{\psi\geq\frac{1}{t}\}}\nabla\psi\cdot\nabla(\xi k(t\psi)-\xi)dX\\ =&\int_{\Omega}\nabla\psi\cdot\nabla\xi dX-\int_{\Omega\cap\{\psi>0\}}f(\psi)\xi dX\\ &+\int_{\Omega\cap\{\frac{1}{t}\leq\psi\leq\frac{2}{t}\}}(2-\psi t)f(\psi)\xi dX+\int_{\Omega\cap\{\psi\leq\frac{2}{t}\}}f(\psi)\xi dX\\ \geq&\int_{\Omega}\nabla\psi\cdot\nabla\xi dX-\int_{\Omega\cap\{\psi>0\}}f(\psi)\xi dX,\end{array} (3.1)

for sufficiently large t>0\displaystyle t>0, where we have used the fact that f(ψ)f(0)20\displaystyle f(\psi)\geq\frac{f(0)}{2}\geq 0 in Ω{ψ2t}\displaystyle\Omega\cap\left\{\psi\leq\frac{2}{t}\right\} for large t>0\displaystyle t>0.

On the other hand, one has

Ωψ(ξk(tψ))dX=Ω{1tψ2t}(2ψt)ψξξt|ψ|2dX+Ω{0<ψ1t}ψξdXΩ{0<ψ2t}|ψ||ξ|𝑑X.\begin{array}[]{rl}\int_{\Omega}\nabla\psi\cdot\nabla(\xi k(t\psi))dX=&\int_{\Omega\cap\{\frac{1}{t}\leq\psi\leq\frac{2}{t}\}}(2-\psi t)\nabla\psi\cdot\nabla\xi-\xi t|\nabla\psi|^{2}dX\\ &+\int_{\Omega\cap\{0<\psi\leq\frac{1}{t}\}}\nabla\psi\cdot\nabla\xi dX\\ \leq&\int_{\Omega\cap\{0<\psi\leq\frac{2}{t}\}}|\nabla\psi||\nabla\xi|dX.\end{array} (3.2)

It follows from (3.1) and (3.2) that

ΩψξdXΩ{ψ>0}f(ψ)ξ𝑑XΩ{0<ψ2t}|ψ||ξ|𝑑X.\begin{array}[]{rl}\int_{\Omega}\nabla\psi\cdot\nabla\xi dX-\int_{\Omega\cap\{\psi>0\}}f(\psi)\xi dX\leq\int_{\Omega\cap\{0<\psi\leq\frac{2}{t}\}}|\nabla\psi||\nabla\xi|dX.\end{array} (3.3)

Taking t+\displaystyle t\rightarrow+\infty in (3.3), we conclude that Δψ+f(ψ)I{ψ>0}0\displaystyle\Delta\psi+f(\psi)I_{\{\psi>0\}}\geq 0 in the sense of distributions. Consequently, there exists a positive Radon measure μ0\displaystyle\mu_{0} supported on Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\}, such that μ0=Δψ+f(ψ)I{ψ>0}\displaystyle\mu_{0}=\Delta\psi+f(\psi)I_{\{\psi>0\}}.

Recalling f(0)0\displaystyle f(0)\geq 0, there exists a positive Radon measure μ\displaystyle\mu supported on Ω{ψ=0}\displaystyle\Omega\cap\{\psi=0\}, such that μ=μ0+f(0)I{ψ=0}=Δψ+f(ψ)\displaystyle\mu=\mu_{0}+f(0)I_{\{\psi=0\}}=\Delta\psi+f(\psi).

Next, we will give the estimate of the Radon measure μ\displaystyle\mu.

Lemma 3.2.

Let G\displaystyle G be a compact subset of Ω\displaystyle\Omega. There exist some positive constants r0,c\displaystyle r_{0},c and C\displaystyle C, such that for any disc Br=Br(X0)G\displaystyle B_{r}=B_{r}(X_{0})\subset G with X0Γ\displaystyle X_{0}\in\Gamma and r<r0\displaystyle r<r_{0},

crBr𝑑μCr.cr\leq\int_{B_{r}}d\mu\leq Cr. (3.4)

Furthermore,

crBr{ψ>0}𝑑μ=Br𝑑μ0Cr.cr\leq\int_{B_{r}\cap\partial\{\psi>0\}}d\mu=\int_{B_{r}}d\mu_{0}\leq Cr. (3.5)
Proof.

For any ε>0\displaystyle\varepsilon>0, taking a test function dε,Br(X)=min{dist(X,2Br)ε,1}\displaystyle d_{\varepsilon,B_{r}}(X)=\min\left\{\frac{\text{dist}(X,\mathbb{R}^{2}\setminus B_{r})}{\varepsilon},1\right\} for a set BrG\displaystyle B_{r}\subset G, it is easy to check that dε,Br(X)\displaystyle d_{\varepsilon,B_{r}}(X) converges to IBr\displaystyle I_{B_{r}} as ε0\displaystyle\varepsilon\rightarrow 0 and

Ωdε,Br𝑑μ=Ωψdε,Br+f(ψ)dε,BrdX.\begin{array}[]{rl}\int_{\Omega}d_{\varepsilon,B_{r}}d\mu=&\int_{\Omega}\nabla\psi\cdot\nabla d_{\varepsilon,B_{r}}+f(\psi)d_{\varepsilon,B_{r}}dX.\end{array} (3.6)

Taking ε0\displaystyle\varepsilon\rightarrow 0 in (3.6), it follows from Lemma 2.4 that

Br𝑑μ=Brψνd1+Brf(ψ)𝑑XCr+Λπr2Cr,\int_{B_{r}}d\mu=\int_{\partial B_{r}}\nabla\psi\cdot\nu d\mathcal{H}^{1}+\int_{B_{r}}f(\psi)dX\leq Cr+\Lambda\pi r^{2}\leq Cr, (3.7)

where 1\displaystyle\mathcal{H}^{1} is the one-dimensional Hausdorff measure on 2\displaystyle\mathbb{R}^{2} and ν\displaystyle\nu is the outer normal vector.

Let YBr\displaystyle Y\in B_{r} and GY(X)>0\displaystyle G_{Y}(X)>0 be the Green function for Laplacian in Br\displaystyle B_{r} with pole Y\displaystyle Y. If ψ(Y)>0\displaystyle\psi(Y)>0, then the pole Y\displaystyle Y is outside the support of the Radon measure μ\displaystyle\mu, and thus

BrGY(X)𝑑μ=ψ(Y)Brf(ψ)GY(X)𝑑X+BrψνGY(X)d1.\int_{B_{r}}G_{Y}(X)d\mu=-\psi(Y)-\int_{B_{r}}f(\psi)G_{Y}(X)dX+\int_{\partial B_{r}}\psi\partial_{-\nu}G_{Y}(X)d\mathcal{H}^{1}. (3.8)

Thanks to the non-degeneracy Lemma 2.6, there exists a point YBκr\displaystyle Y\in\partial B_{\kappa r}, such that ψ(Y)cκλr>0\displaystyle\psi(Y)\geq c^{*}_{\kappa}\lambda r>0 for all small κ>0\displaystyle\kappa>0. Recalling that ψ\displaystyle\psi is Lipschitz continuous, we have

ψ(Y)=ψ(Y)ψ(X0)C|YX0|=Cκrandψ>0inB𝔠(κ)r(Y),\psi(Y)=\psi(Y)-\psi(X_{0})\leq C|Y-X_{0}|=C\kappa r\ \ \text{and}\ \ \psi>0\ \ \text{in}\ \ B_{\mathfrak{c}(\kappa)r}(Y), (3.9)

for some small constant 𝔠(κ)\displaystyle\mathfrak{c}(\kappa). Then it follows from (3.8) and (3.9) that

BrGY(X)𝑑μCκrCΛr2+c(1κ)Brψ𝑑1c0r,(c0>0),\int_{B_{r}}G_{Y}(X)d\mu\geq-C\kappa r-C\Lambda r^{2}+c(1-\kappa)\fint_{\partial B_{r}}\psi d\mathcal{H}^{1}\geq c_{0}r,\ \ (c_{0}>0), (3.10)

provided that κ\displaystyle\kappa and r\displaystyle r are small enough.

On the other hand, we have

BrGY(X)𝑑μ=Br{ψ=0}GY(X)𝑑μ(supBr{ψ=0}GY(X))Br𝑑μCkBr𝑑μ,\int_{B_{r}}G_{Y}(X)d\mu=\int_{B_{r}\cap\{\psi=0\}}G_{Y}(X)d\mu\leq\left(\sup_{B_{r}\cap\{\psi=0\}}G_{Y}(X)\right)\int_{B_{r}}d\mu\leq C_{k}\int_{B_{r}}d\mu, (3.11)

where we have used the fact that dist(Y,Br{ψ=0})𝔠(κ)r\displaystyle\text{dist}(Y,B_{r}\cap\{\psi=0\})\geq\mathfrak{c}(\kappa)r. Combining (3.10) and (3.11), one has

Br𝑑μcr,\int_{B_{r}}d\mu\geq cr,

which together with (3.7) gives (3.4).

It is clear that

Br{ψ>0}𝑑μBr𝑑μCr,\int_{B_{r}\cap\partial\{\psi>0\}}d\mu\leq\int_{B_{r}}d\mu\leq Cr,

this gives the right-hand side of (3.5). On the other hand, we have

Br𝑑μ=Br{ψ>0}𝑑μ+Br{ψ=0}𝑑μBr{ψ>0}𝑑μ+Br{ψ=0}f(0)𝑑XBr{ψ>0}𝑑μ+Λπr2,\begin{array}[]{rl}\int_{B_{r}}d\mu=&\int_{B_{r}\cap\partial\{\psi>0\}}d\mu+\int_{B_{r}\cap\{\psi=0\}}d\mu\\ \leq&\int_{B_{r}\cap\partial\{\psi>0\}}d\mu+\int_{B_{r}\cap\{\psi=0\}}f(0)dX\\ \leq&\int_{B_{r}\cap\partial\{\psi>0\}}d\mu+\Lambda\pi r^{2},\end{array}

which implies that

Br{ψ>0}𝑑μBr𝑑μΛπr2crΛπr2crfor small r,\int_{B_{r}\cap\partial\{\psi>0\}}d\mu\geq\int_{B_{r}}d\mu-\Lambda\pi r^{2}\geq cr-\Lambda\pi r^{2}\geq cr\ \ \text{for small $\displaystyle r$},

this gives the left-hand side of (3.5).

We next introduce the representation theorem as follows, which implies that Ω+=Ω{ψ>0}\displaystyle\Omega_{+}=\Omega\cap\{\psi>0\} has finite perimeter.

Proposition 3.3.

(Representation Theorem) Let ψ\displaystyle\psi be a minimizer, there holds that:

(1). 1(D{ψ>0})<+\displaystyle\mathcal{H}^{1}(D\cap\partial\{\psi>0\})<+\infty for any compact subset D\displaystyle D of Ω\displaystyle\Omega.

(2). There exists a Borel function ψ\displaystyle\mathcal{B}_{\psi}, such that

Δψ+f(ψ)I{ψ>0}=ψ1{ψ>0},\Delta\psi+f(\psi)I_{\{\psi>0\}}=\mathcal{B}_{\psi}\mathcal{H}^{1}\lfloor_{\partial\{\psi>0\}},

that is, for any ξC0(Ω)\displaystyle\xi\in C_{0}^{\infty}(\Omega),

ΩψξdX+Ω{ψ>0}f(ψ)ξ𝑑X=Ω{ψ>0}ψξ𝑑1.-\int_{\Omega}\nabla\psi\cdot\nabla\xi dX+\int_{\Omega\cap\{\psi>0\}}f(\psi)\xi dX=\int_{\Omega\cap\partial\{\psi>0\}}\mathcal{B}_{\psi}\xi d\mathcal{H}^{1}.

(3). For any compact subset D\displaystyle D of Ω\displaystyle\Omega, Br(X0)D\displaystyle B_{r}(X_{0})\subset D with X0Γ\displaystyle X_{0}\in\Gamma and small r>0\displaystyle r>0,

0<cψ(X)C<+andcr1(Br(X0){ψ>0})Cr,0<c\leq\mathcal{B}_{\psi}(X)\leq C<+\infty\ \ \text{and}\ \ cr\leq\mathcal{H}^{1}(B_{r}(X_{0})\cap\partial\{\psi>0\})\leq Cr,

for some positive constants c,C\displaystyle c,C independent of r\displaystyle r and X0\displaystyle X_{0}.

Proof.

Thanks to the fact (3.5), and along the similar arguments in the proof of Theorem 4.5 in [2], we can obtain that

1Cμ0(G)1(G)Cμ0(G)for any compact subset G of D{ψ>0},\frac{1}{C}\mu_{0}(G)\leq\mathcal{H}^{1}(G)\leq C\mu_{0}(G)\ \ \text{for any compact subset $\displaystyle G$ of $\displaystyle D\cap\partial\{\psi>0\}$},

Denote μ0=Δψ+f(ψ)I{ψ>0}\displaystyle\mu_{0}=\Delta\psi+f(\psi)I_{\{\psi>0\}}, this gives the assertion (1).

Therefore, μ0\displaystyle\mu_{0} is absolutely continuous with respect to 1{ψ>0}\displaystyle\mathcal{H}^{1}\lfloor_{\partial\{\psi>0\}}. It follows from Theorem 2.5.8 in [20] that there exists a Radon-Nikodym derivative ψ=dμ0d(1{ψ>0})\displaystyle\mathcal{B}_{\psi}=\frac{d\mu_{0}}{d(\mathcal{H}^{1}\lfloor_{\partial\{\psi>0\}})}, namely,

μ0(G)=Gψd(1{ψ>0})for any GΩ,\mu_{0}(G)=\int_{G}\mathcal{B}_{\psi}d(\mathcal{H}^{1}\lfloor_{\partial\{\psi>0\}})\ \ \text{for any $\displaystyle G\subset\Omega$},

which together with (3.5) gives the assertion (3).

Then we have

Ω{ψ=0}ξ𝑑μ+Ωξ𝑑μ0=Ωξ𝑑μ=ΩψξdX+Ωf(ψ)ξ𝑑X,\int_{\Omega\cap\{\psi=0\}}\xi d\mu+\int_{\Omega}\xi d\mu_{0}=\int_{\Omega}\xi d\mu=-\int_{\Omega}\nabla\psi\cdot\nabla\xi dX+\int_{\Omega}f(\psi)\xi dX,

for any ξC0(Ω)\displaystyle\xi\in C_{0}^{\infty}(\Omega), which gives that

ΩψξdX+Ω{ψ>0}f(ψ)ξ𝑑X=Ω{ψ>0}Bψξ𝑑1,-\int_{\Omega}\nabla\psi\cdot\nabla\xi dX+\int_{\Omega\cap\{\psi>0\}}f(\psi)\xi dX=\int_{\Omega\cap\partial\{\psi>0\}}B_{\psi}\xi d\mathcal{H}^{1},

due to Ω{ψ=0}ξ𝑑μ=Ω{ψ=0}f(0)ξ𝑑X\displaystyle\int_{\Omega\cap\{\psi=0\}}\xi d\mu=\int_{\Omega\cap\{\psi=0\}}f(0)\xi dX. This yields the assertion (2).

By virtue of (1) in Proposition 3.3, it follows from Theorem 1 in §5.11\displaystyle\S 5.11 in [19] that the set Ω+=Ω{ψ>0}\displaystyle\Omega_{+}=\Omega\cap\{\psi>0\} has finite perimeter, that is, 𝒱ψ=IΩ+\displaystyle\mathcal{V}_{\psi}=-\nabla I_{\Omega_{+}} is Borel measure and the total variation |𝒱ψ|\displaystyle|\mathcal{V}_{\psi}| is a Radon measure. Denote the reduced boundary of the set Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\} as

red{ψ>0}={XΩ{ψ>0}|νψ|=1},\partial_{red}\{\psi>0\}=\{X\in\Omega\cap\partial\{\psi>0\}\mid|\nu_{\psi}|=1\},

where νψ\displaystyle\nu_{\psi} is the unique unit vector with

Br(X)|I{ψ>0}I{Y(YX)νψ<0}|𝑑Y=o(r2)as r0\int_{B_{r}(X)}|I_{\{\psi>0\}}-I_{\{Y\mid(Y-X)\cdot\nu_{\psi}<0\}}|dY=o(r^{2})\ \ \text{as $\displaystyle r\rightarrow 0$, }

if such a vector exists, and νψ=0\displaystyle\nu_{\psi}=0 otherwise. Furthermore, it follows from Theorem 4.5.6 in [20] that

𝒱ψ=νψ1red{ψ>0}.\mathcal{V}_{\psi}=\nu_{\psi}\mathcal{H}^{1}\lfloor\ \partial_{red}\{\psi>0\}.

3.2. Blow-up limits

In this subsection, we study some properties of the so-called blow-up limits.

For any X0Ω\displaystyle X_{0}\in\Omega, take two sequences {Xn}\displaystyle\{X_{n}\} and {ρn}\displaystyle\{\rho_{n}\} with XnΩ\displaystyle X_{n}\in\Omega and ρn>0\displaystyle\rho_{n}>0, such that ψ(Xn)=0\displaystyle\psi(X_{n})=0, XnX0\displaystyle X_{n}\rightarrow X_{0} and ρn0\displaystyle\rho_{n}\rightarrow 0 as n0\displaystyle n\rightarrow 0. We call the sequence of functions defined by

ψn(X)=ψ(Xn+ρnX)ρnandλn(X)=λ(Xn+ρnX),\psi_{n}(X)=\frac{\psi(X_{n}+\rho_{n}X)}{\rho_{n}}\ \ \text{and}\ \ \lambda_{n}(X)=\lambda(X_{n}+\rho_{n}X),

as the blow-up sequence with respect to Bρn(Xn)\displaystyle B_{\rho_{n}}(X_{n}), X2\displaystyle X\in\mathbb{R}^{2}. It follows from Lemma 2.4 that |ψn(X)|C\displaystyle|\nabla\psi_{n}(X)|\leq C in any compact set of 2\displaystyle\mathbb{R}^{2}, provided that n\displaystyle n is large enough. Since ψn(0)=0\displaystyle\psi_{n}(0)=0, there exists a blow-up limit ψ0:2\displaystyle\psi_{0}:\mathbb{R}^{2}\rightarrow\mathbb{R}, such that for a subsequence {ψn}\displaystyle\{\psi_{n}\},

ψnψ0in Cloc0,α(2) for any α(0,1),\psi_{n}\rightarrow\psi_{0}\ \ \text{in $\displaystyle C_{loc}^{0,\alpha}(\mathbb{R}^{2})$ for any $\displaystyle\alpha\in(0,1)$}, (3.12)

and

ψnψ0weakly star in Lloc(2).\nabla\psi_{n}\rightarrow\nabla\psi_{0}\ \ \text{weakly star in $\displaystyle L_{loc}^{\infty}(\mathbb{R}^{2})$.}

Recalling the definition of the Hausdorff distance d(E,F)\displaystyle d(E,F) between two sets E\displaystyle E and F\displaystyle F as the infimum of the numbers ε\displaystyle\varepsilon, such that

FXEBε(X)andEXFBε(X).F\subset\bigcup_{X\in E}B_{\varepsilon}(X)\ \ \text{and}\ \ E\subset\bigcup_{X\in F}B_{\varepsilon}(X).

Next, we will give the several convergences of the blow-up sequence, the idea borrows from the similar arguments for Laplace equation in Lemma 3.6 in Chapter 3 in [23].

Lemma 3.4.

The following properties hold:

(1) {ψn>0}{ψ0>0}locally in Hausdorff distance.\displaystyle\partial{\{\psi_{n}>0\}}\rightarrow\partial{\{\psi_{0}>0\}}\ \ \text{locally in Hausdorff distance.}

(2) I{ψn>0}I{ψ0>0} in Lloc1(2).\displaystyle I_{\{\psi_{n}>0\}}\rightarrow I_{\{\psi_{0}>0\}}\ \ \text{ in $\displaystyle L_{loc}^{1}(\mathbb{R}^{2})$.}

(3) If Xn{ψ>0}\displaystyle X_{n}\in\partial\{\psi>0\}, then 0{ψ0>0}\displaystyle 0\in\partial\{\psi_{0}>0\}.

(4) ψnψ0a.e. in 2.\displaystyle\nabla\psi_{n}\rightarrow\nabla\psi_{0}\ \ \text{a.e. in $\displaystyle\mathbb{R}^{2}$}.

(5) If ψ(Xn)=0\displaystyle\psi(X_{n})=0 and XnX0Ω\displaystyle X_{n}\rightarrow X_{0}\in\Omega, then every blow-up limit ψ0\displaystyle\psi_{0} with respect to Bρn(Xn)\displaystyle B_{\rho_{n}}(X_{n}) is an absolute minimum for the functional J0(ϕ)\displaystyle J_{0}(\phi) in Br=Br(0)\displaystyle B_{r}=B_{r}(0) for any r>0\displaystyle r>0, namely,

J0(ψ0)=minJ0(ϕ)for any ϕH1(Br) and ϕ=ψ0 on Br,J_{0}(\psi_{0})=\min J_{0}(\phi)\ \ \text{for any $\displaystyle\phi\in H^{1}(B_{r})$ and $\displaystyle\phi=\psi_{0}$ on $\displaystyle\partial B_{r}$}, (3.13)

where J0(ϕ)=Br|ϕ|2+λ2(X0)I{ϕ>0}dX\displaystyle J_{0}(\phi)=\int_{B_{r}}|\nabla\phi|^{2}+\lambda^{2}(X_{0})I_{\{\phi>0\}}dX.

Proof.

(1). For any X0Ω\displaystyle X_{0}\in\Omega, if X0{ψ0>0}\displaystyle X_{0}\notin\partial\{\psi_{0}>0\}, then there exists a small r>0\displaystyle r>0 such that Br(X0){ψ0>0}=\displaystyle B_{r}(X_{0})\cap\partial\{\psi_{0}>0\}=\varnothing with Br(X0)Ω\displaystyle B_{r}(X_{0})\subset\Omega. We next claim that

Br4(X0){ψn>0}=\displaystyle B_{\frac{r}{4}}(X_{0})\cap\partial\{\psi_{n}>0\}=\varnothing for sufficiently large n\displaystyle n. (3.14)

In fact, it follows from (3.12) that the claim (3.14) is true, if ψ0>0\displaystyle\psi_{0}>0 in Br(X0)\displaystyle B_{r}(X_{0}).

If ψ00\displaystyle\psi_{0}\equiv 0 in Br(X0)\displaystyle B_{r}(X_{0}), for any fixed small ε>0\displaystyle\varepsilon>0, there exists a N=N(ε)\displaystyle N=N(\varepsilon), such that ψn<ε\displaystyle\psi_{n}<\varepsilon in Br2(X0)\displaystyle B_{\frac{r}{2}}(X_{0}) for any n>N\displaystyle n>N, and one has

2rBr2(X0)ψn𝑑S<2εrc12λ1for sufficiently large n,\frac{2}{r}\fint_{\partial B_{\frac{r}{2}}(X_{0})}\psi_{n}dS<\frac{2\varepsilon}{r}\leq c_{\frac{1}{2}}^{*}\lambda_{1}\ \text{for sufficiently large $\displaystyle n$},

which together with the non-degeneracy Lemma 2.6 implies that ψn0inBr4(X0)\displaystyle\psi_{n}\equiv 0\ \text{in}\ B_{\frac{r}{4}}(X_{0}) for n>N\displaystyle n>N, this gives the claim (3.14).

Reversely, for any X0Ω\displaystyle X_{0}\in\Omega, if X0{ψn>0}\displaystyle X_{0}\notin\partial\{\psi_{n}>0\} for a subsequence {ψn}\displaystyle\{\psi_{n}\}, then Br(X0){ψn>0}=\displaystyle B_{r}(X_{0})\cap\partial\{\psi_{n}>0\}=\varnothing for small r>0\displaystyle r>0. Next, we claim that

Br4(X0){ψ0>0}=.\text{$\displaystyle B_{\frac{r}{4}}(X_{0})\cap\partial\{\psi_{0}>0\}=\varnothing$}. (3.15)

If ψn>0\displaystyle\psi_{n}>0 in Br(X0)\displaystyle B_{r}(X_{0}), we have

Δψn+ρnf(ρnψn)=0inBr(X0),\Delta\psi_{n}+\rho_{n}f(\rho_{n}\psi_{n})=0\ \ \text{in}\ \ B_{r}(X_{0}),

which implies that

Δψ0=0inBr2(X0),ψ00inBr2(X0).\Delta\psi_{0}=0\ \text{in}\ B_{\frac{r}{2}}(X_{0}),\ \psi_{0}\geq 0\ \text{in}\ B_{\frac{r}{2}}(X_{0}).

The strong maximum principle yields that

eitherψ00orψ0>0 inBr2(X0),\text{either}\ \ \psi_{0}\equiv 0\ \ \text{or}\ \ \psi_{0}>0\ \text{ in}\ B_{\frac{r}{2}}(X_{0}),

which gives the claim (3.15).

It is easy to check that the claim (3.15) holds, in the case of ψn0\displaystyle\psi_{n}\equiv 0 in Br(X0)\displaystyle B_{r}(X_{0}).

Hence, we obtain the convergence of the free boundary in the Hausdorff distance.

(2). For any Y0{ψ0>0}\displaystyle Y_{0}\in\partial\{\psi_{0}>0\}, it follows from the results in Step 1 that there exists a sequence {Yn}\displaystyle\{Y_{n}\} with Yn{ψn>0}\displaystyle Y_{n}\in\partial\{\psi_{n}>0\}, such that YnY0\displaystyle Y_{n}\rightarrow Y_{0}. Since ρnYn+Xn{ψ>0}\displaystyle\rho_{n}Y_{n}+X_{n}\in\partial\{\psi>0\}, by using Lemma 2.5 and Lemma 2.6 for ψn\displaystyle\psi_{n}, we have

c12λ11rBr(Yn)ψn(X)𝑑SX=1rρnBrρn(ρnYn+Xn)ψ(Z)𝑑SZC(λ2+Λ),c_{\frac{1}{2}}^{*}\lambda_{1}\leq\frac{1}{r}\fint_{\partial B_{r}(Y_{n})}\psi_{n}(X)dS_{X}=\frac{1}{r\rho_{n}}\fint_{\partial B_{r\rho_{n}}(\rho_{n}Y_{n}+X_{n})}\psi(Z)dS_{Z}\leq C^{*}(\lambda_{2}+\Lambda), (3.16)

for any r>0\displaystyle r>0, provided that n\displaystyle n is sufficiently large. Then taking n+\displaystyle n\rightarrow+\infty in (3.16) gives that

c12λ11rBr(Y0)ψ0𝑑SC(λ2+Λ)forY0{ψ0>0},c_{\frac{1}{2}}^{*}\lambda_{1}\leq\frac{1}{r}\fint_{\partial B_{r}(Y_{0})}\psi_{0}dS\leq C^{*}(\lambda_{2}+\Lambda)\ \ \text{for}\ \ Y_{0}\in\partial\{\psi_{0}>0\}, (3.17)

which together with Theorem 4.5 in [2] imply that

1({ψ0>0}D)<+for any compact subset D of 2.\mathcal{H}^{1}(\partial\{\psi_{0}>0\}\cap D)<+\infty\ \ \text{for any compact subset $\displaystyle D$ of $\displaystyle\mathbb{R}^{2}$}. (3.18)

Here, 1\displaystyle\mathcal{H}^{1} is the one-dimensional Hausdorff measure on 2\displaystyle\mathbb{R}^{2}. Consequently,

2({ψ0>0}D)=0,\mathcal{L}^{2}(\partial\{\psi_{0}>0\}\cap D)=0,

where 2\displaystyle\mathcal{L}^{2} is the two-dimensional Lebesgue measure on 2\displaystyle\mathbb{R}^{2}.

Let Oεn\displaystyle O_{\varepsilon_{n}} be an εn\displaystyle\varepsilon_{n}-neighborhood of {ψ0>0}\displaystyle\partial\{\psi_{0}>0\}, such that

{ψn>0}Oεnand2(DOεn)0asεn0.\partial\{\psi_{n}>0\}\subset O_{\varepsilon_{n}}\ \ \text{and}\ \ \mathcal{L}^{2}(D\cap O_{\varepsilon_{n}})\rightarrow 0\ \quad\text{as}\ \varepsilon_{n}\rightarrow 0. (3.19)

Hence, it follows from the results in Step 1 that

D|I{ψn>0}I{ψ0>0}|𝑑XDOεn1𝑑X=2(DOεn),\int_{D}\left|I_{\{\psi_{n}>0\}}-I_{\{\psi_{0}>0\}}\right|dX\leq\int_{D\cap O_{\varepsilon_{n}}}1dX=\mathcal{L}^{2}(D\cap O_{\varepsilon_{n}}),

for sufficiently large n\displaystyle n, which together with (3.19) gives that

I{ψn>0}I{ψ0>0}\displaystyle I_{\{\psi_{n}>0\}}\rightarrow I_{\{\psi_{0}>0\}}\ in L1(D)\displaystyle\ L^{1}(D).

(3). If Xn\displaystyle X_{n} is a free boundary point of the minimizer ψ\displaystyle\psi, it follows from Lemma 2.5 and Lemma 2.6 that

c12λ11rBr(0)ψn(X)𝑑SX=1rρnBrρn(Xn)ψ(Z)𝑑SZC(λ2+Λ),c_{\frac{1}{2}}^{*}\lambda_{1}\leq\frac{1}{r}\fint_{\partial B_{r}(0)}\psi_{n}(X)dS_{X}=\frac{1}{r\rho_{n}}\fint_{\partial B_{r\rho_{n}}(X_{n})}\psi(Z)dS_{Z}\leq C^{*}(\lambda_{2}+\Lambda), (3.20)

for any r>0\displaystyle r>0, provided that n\displaystyle n is sufficiently large. Then taking n+\displaystyle n\rightarrow+\infty in (3.20), one has

c12λ11rBr(0)ψ0𝑑SC(λ2+Λ)for any r>0,c_{\frac{1}{2}}^{*}\lambda_{1}\leq\frac{1}{r}\fint_{\partial B_{r}(0)}\psi_{0}dS\leq C^{*}(\lambda_{2}+\Lambda)\ \ \text{for any $\displaystyle r>0$},

which gives that 0{ψ0>0}\displaystyle 0\in\partial\{\psi_{0}>0\}.

(4). Let D\displaystyle D be any compact subset of {ψ0>0}\displaystyle\{\psi_{0}>0\}, thanks to the standard elliptic estimates for ψn\displaystyle\psi_{n}, one has

ψnψ0uniformly in D.\nabla\psi_{n}\rightarrow\nabla\psi_{0}\ \ \text{uniformly in $\displaystyle D$}. (3.21)

Next, we claim that

ψnψ0a.e. in {ψ0=0}.\nabla\psi_{n}\rightarrow\nabla\psi_{0}\ \ \text{a.e. in $\displaystyle\{\psi_{0}=0\}$}. (3.22)

Since {ψ0=0}\displaystyle\{\psi_{0}=0\} is 2\displaystyle\mathcal{L}^{2}-measurable, it follows from Corollary 3 of Section 1.7 in [19] that

limr02(Br(X){ψ0=0})2(Br(X))=1for 2 a.e. X{ψ0=0}.\lim_{r\rightarrow 0}\frac{\mathcal{L}^{2}(B_{r}(X)\cap\{\psi_{0}=0\})}{\mathcal{L}^{2}(B_{r}(X))}=1\ \ \text{for $\displaystyle\mathcal{L}^{2}$ a.e. $\displaystyle X\in\{\psi_{0}=0\}$}.

Denote

𝒮={X{ψ0=0}limr02(Br(X){ψ0=0})2(Br(X))=1}.\mathcal{S}=\left\{X\in\{\psi_{0}=0\}\mid\lim_{r\rightarrow 0}\frac{\mathcal{L}^{2}(B_{r}(X)\cap\{\psi_{0}=0\})}{\mathcal{L}^{2}(B_{r}(X))}=1\right\}.

We next show that

ψ0(X0+X)=o(|X|) for any X0𝒮.\text{$\displaystyle\psi_{0}(X_{0}+X)=o(|X|)\ $ for any $\displaystyle X_{0}\in\mathcal{S}$}. (3.23)

In fact, suppose not, we assume that ψ0(Y)>kr\displaystyle\psi_{0}(Y)>kr for some YBr(X0)\displaystyle Y\in B_{r}(X_{0}) with r0\displaystyle r\rightarrow 0 and k>0\displaystyle k>0. With the aid of (3.17), it follows from Theorem 4.3 and Remark 4.4 in [2] that ψ0\displaystyle\psi_{0} is Lipschitz continuous, which implies that

ψ0(X)>k2rin Bεkr(Y) for some small ε>0.\psi_{0}(X)>\frac{k}{2}r\ \ \text{in $\displaystyle B_{\varepsilon kr}(Y)$ for some small $\displaystyle\varepsilon>0$}.

This gives that {ψ0>0}\displaystyle\{\psi_{0}>0\} has positive density at X0\displaystyle X_{0}, which contradicts to X0𝒮\displaystyle X_{0}\in\mathcal{S}. Thus, we obtain the fact (3.23).

With the aid of (3.12) and (3.23), for any ε>0\displaystyle\varepsilon>0, we have

ψnr<εin Br(X0) for small r,\frac{\psi_{n}}{r}<\varepsilon\ \ \text{in $\displaystyle B_{r}(X_{0})$ for small $\displaystyle r$},

provided that n\displaystyle n is sufficiently large, that is n>N(ε,r)\displaystyle n>N(\varepsilon,r). It follows from the non-degeneracy Lemma 2.6 that ψn0\displaystyle\psi_{n}\equiv 0 in Br2(X0)\displaystyle B_{\frac{r}{2}}(X_{0}), which implies that ψ00\displaystyle\psi_{0}\equiv 0 in Br2(X0)\displaystyle B_{\frac{r}{2}}(X_{0}), and thus 𝒮\displaystyle\mathcal{S} is open. Furthermore, one has

ψnψ0\displaystyle\psi_{n}\equiv\psi_{0} in any compact subset of 𝒮\displaystyle\mathcal{S}, provided that n\displaystyle n is sufficiently large.

This completes the proof of the claim (3.22).

Since 2({ψ0>0})=0\displaystyle\mathcal{L}^{2}(\partial\{\psi_{0}>0\})=0, it follows from (3.21) and (3.22) that ψnψ0\displaystyle\nabla\psi_{n}\rightarrow\nabla\psi_{0} a.e. in 2\displaystyle\mathbb{R}^{2}.

(5). For any ϕH1(Br)\displaystyle\phi\in H^{1}(B_{r}) and ϕ=ψ0\displaystyle\phi=\psi_{0} on Br\displaystyle\partial B_{r} with Br=Br(0)\displaystyle B_{r}=B_{r}(0), it suffices to show that

J0(ψ0)J0(ϕ).J_{0}(\psi_{0})\leq J_{0}(\phi). (3.24)

Taking ηε(X)=min{1εdist(X,2Br),1}\displaystyle\eta_{\varepsilon}(X)=\min\left\{\frac{1}{\varepsilon}dist(X,\mathbb{R}^{2}\setminus B_{r}),1\right\}, it is easy to see that ηεC00,1(Br)\displaystyle\eta_{\varepsilon}\in C_{0}^{0,1}(B_{r}) and 0ηε1\displaystyle 0\leq\eta_{\varepsilon}\leq 1. Set

ϕn=ϕ+(1ηε)(ψnψ0).\phi_{n}=\phi+(1-\eta_{\varepsilon})(\psi_{n}-\psi_{0}).

It is easy to check that

ϕn=ψnoutsideBr,and{ϕn>0}{ϕ>0}{ηε<1}.\phi_{n}=\psi_{n}\ \text{outside}\ \ B_{r},\ \ \text{and}\ \ \{\phi_{n}>0\}\subset\{\phi>0\}\cup\{\eta_{\varepsilon}<1\}.

Then we have

Br|ψn|2+F(ρnψn)+λn2I{ψn>0}dXBr|ϕn|2+F(ρnϕn)+λn2I{ϕn>0}dXBr|ϕn|2+F(ρnϕn)+λn2I{ϕn>0}dX+Brλn2I{ηε<1}𝑑X.\begin{array}[]{rl}&\int_{B_{r}}|\nabla\psi_{n}|^{2}+F(\rho_{n}\psi_{n})+\lambda_{n}^{2}I_{\{\psi_{n}>0\}}dX\\ \leq&\int_{B_{r}}|\nabla\phi_{n}|^{2}+F(\rho_{n}\phi_{n})+\lambda_{n}^{2}I_{\{\phi_{n}>0\}}dX\\ \leq&\int_{B_{r}}|\nabla\phi_{n}|^{2}+F(\rho_{n}\phi_{n})+\lambda_{n}^{2}I_{\{\phi_{n}>0\}}dX+\int_{B_{r}}\lambda_{n}^{2}I_{\{\eta_{\varepsilon}<1\}}dX.\end{array} (3.25)

By virtue of the results in the statements (2) and (4), taking n+\displaystyle n\rightarrow+\infty in (3.25) gives that

Br|ψ0|2+λ2(X0)I{ψ0>0}dXBr|ϕ|2+λ2(X0)I{ϕ>0}dX+λ2(X0)BrI{ηε<1}𝑑X.\begin{array}[]{rl}&\int_{B_{r}}|\nabla\psi_{0}|^{2}+\lambda^{2}(X_{0})I_{\{\psi_{0}>0\}}dX\\ \leq&\int_{B_{r}}|\nabla\phi|^{2}+\lambda^{2}(X_{0})I_{\{\phi>0\}}dX+\lambda^{2}(X_{0})\int_{B_{r}}I_{\{\eta_{\varepsilon}<1\}}dX.\end{array} (3.26)

Taking ε0\displaystyle\varepsilon\rightarrow 0 in (3.26), we complete the proof of the claim (3.24).

3.3. Linear growth near the free boundary

In this subsection, we will obtain the gradient estimate of ψ\displaystyle\psi near the free boundary, and show that ψ(X)\displaystyle\psi(X) should grow linearly away from the free boundary. Namely,

ψ(X+X0)=λ(X0)max{Xν(X0),0}+o(|X|),\psi(X+X_{0})=\lambda(X_{0})\max\{-X\cdot\nu(X_{0}),0\}+o(|X|), (3.27)

for X0\displaystyle X\rightarrow 0, where X0Ω{ψ>0}\displaystyle X_{0}\in\Omega\cap\partial\{\psi>0\} and ν(X0)\displaystyle\nu(X_{0}) is the unit vector.

Lemma 3.5.

For any compact subset D\displaystyle D of Ω\displaystyle\Omega, there exist some positive constants C\displaystyle C and 0<γ<1\displaystyle 0<\gamma<1 depending only on Λ\displaystyle\Lambda, D\displaystyle D and Ω\displaystyle\Omega, such that for any disc Br(X0)BR(X0)D\displaystyle B_{r}(X_{0})\subset B_{R}(X_{0})\subset D with X0Ω{ψ>0}\displaystyle X_{0}\in\Omega\cap\partial\{\psi>0\}, then

supXBr(X0)|ψ(X)|supXBR(X0)λ(X)+C(rR)γ,\sup_{X\in B_{r}(X_{0})}|\nabla\psi(X)|\leq\sup_{X\in B_{R}(X_{0})}\lambda(X)+C\left(\frac{r}{R}\right)^{\gamma}, (3.28)

Furthermore, there exists a α0(0,1)\displaystyle\alpha_{0}\in(0,1), such that

|ψ(X)|λ(X)+Crα0,α0(0,1),|\nabla\psi(X)|\leq\lambda(X)+Cr^{\alpha_{0}},\ \ \alpha_{0}\in(0,1), (3.29)

for any disc Br(X)D\displaystyle B_{r}(X)\subset D touching the free boundary with small r>0\displaystyle r>0.

Proof.

Denote Br=Br(X0)\displaystyle B_{r}=B_{r}(X_{0}) for simplicity, and one has

Δψ+f(ψ)=0inΩ{ψ>0}.\Delta\psi+f(\psi)=0\ \ \text{in}\ \ \Omega\cap\{\psi>0\}.

Step 1. In this step, we will show that

lim supXX0,ψ(X)>0|ψ(X)|=λ(X0),X0Ω{ψ>0}.\limsup_{X\rightarrow X_{0},\psi(X)>0}|\nabla\psi(X)|=\lambda(X_{0}),\ \ \ X_{0}\in\Omega\cap\partial\{\psi>0\}. (3.30)

Denote κ=lim supXX0,ψ(X)>0|ψ(X)|\displaystyle\kappa=\limsup_{X\rightarrow X_{0},\psi(X)>0}|\nabla\psi(X)|, it suffices to prove that κ=λ(X0)\displaystyle\kappa=\lambda(X_{0}). In view of definition of κ\displaystyle\kappa, there exists a sequence {Zn}\displaystyle\{Z_{n}\} with ψ(Zn)>0\displaystyle\psi(Z_{n})>0 and |ψ(Zn)|κ\displaystyle|\nabla\psi(Z_{n})|\rightarrow\kappa. Let YnΓ\displaystyle Y_{n}\in\Gamma be the nearest point to Zn\displaystyle Z_{n} and denote ρn=|YnZn|\displaystyle\rho_{n}=|Y_{n}-Z_{n}|. Let ψ0\displaystyle\psi_{0} be a blow-up limit of a sequence ψ(Yn+ρnX)ρn\displaystyle\frac{\psi(Y_{n}+\rho_{n}X)}{\rho_{n}} with respect to Bρn(Yn)\displaystyle B_{\rho_{n}}(Y_{n}), without loss of generality, we assume that

ZnYnρne2,e2=(0,1).\frac{Z_{n}-Y_{n}}{\rho_{n}}\rightarrow-e_{2},\ \ \ e_{2}=(0,1).

By virtue of the statement (5) in Lemma 3.4, we have that ψ0\displaystyle\psi_{0} is a minimizer. It follows from the similar arguments in Lemma 2.2 and Lemma 2.4 in [2] that ψ0\displaystyle\psi_{0} is subharmonic in 2\displaystyle\mathbb{R}^{2} and Δψ0=0\displaystyle\Delta\psi_{0}=0 in {ψ0>0}\displaystyle\{\psi_{0}>0\}. Furthermore, ψ0(0)=0\displaystyle\psi_{0}(0)=0 and

|ψ0|κin {ψ0>0}|ψ0(e2)|=κ and B1(e2){ψ0>0},|\nabla\psi_{0}|\leq\kappa\ \ \text{in $\displaystyle\{\psi_{0}>0\}$, $\displaystyle|\nabla\psi_{0}(-e_{2})|=\kappa$ and $\displaystyle B_{1}(-e_{2})\subset\{\psi_{0}>0\}$}, (3.31)

this gives that κ>0\displaystyle\kappa>0. Define ϕ0=ψ0ν0\displaystyle\phi_{0}=\frac{\partial\psi_{0}}{\partial\nu_{0}}, where ν0=ψ0(e2)|ψ0(e2)|\displaystyle\nu_{0}=-\frac{\nabla\psi_{0}(-e_{2})}{|\nabla\psi_{0}(-e_{2})|}. It is easy to check that ϕ0\displaystyle\phi_{0} is harmonic in {ψ0>0}\displaystyle\{\psi_{0}>0\}. Moreover, it follows from (3.31) that

ϕ0=ψ0ψ0(e2)|ψ0(e2)|κin B1(e2) and ϕ0(e2)=κ.\phi_{0}=-\frac{\nabla\psi_{0}\cdot\nabla\psi_{0}(-e_{2})}{|\nabla\psi_{0}(-e_{2})|}\geq-\kappa\ \text{in $\displaystyle B_{1}(-e_{2})$ and $\displaystyle\phi_{0}(-e_{2})=-\kappa$}.

The strong maximum principle gives that

ϕ0κinB1(e2),\phi_{0}\equiv-\kappa\ \ \text{in}\ \ B_{1}(-e_{2}),

which together with ψ0(0)=0\displaystyle\psi_{0}(0)=0 implies that

ψ0(X)=κXν0inB1(e2).\psi_{0}(X)=-\kappa X\cdot\nu_{0}\ \ \text{in}\ \ B_{1}(-e_{2}).

Since ψ0>0\displaystyle\psi_{0}>0 in B1(e2)\displaystyle B_{1}(-e_{2}), we have that ν0=e2\displaystyle\nu_{0}=e_{2}. Due to the uniqueness of the Cauchy problem for the Laplace equation, one has

ψ0=κyin{y0}.\psi_{0}=-\kappa y\ \ \text{in}\ \ \{y\leq 0\}.

Next, we claim that

ψ0=0in some strip{0<y<ε0}.\psi_{0}=0\ \ \text{in some strip}\ \ \{0<y<\varepsilon_{0}\}. (3.32)

Suppose that the claim (3.32) is not true. Define

l=lim supy0,ψ0(x,y)>0ψ0(x,y)y.l=\limsup_{y\downarrow 0,\psi_{0}(x,y)>0}\frac{\partial\psi_{0}(x,y)}{\partial y}.

Since ψ0\displaystyle\psi_{0} is a local minimizer, it follows from Corollary 3.3 in [2] that ψ0\displaystyle\psi_{0} is Lipschitz continuous, and thus l<\displaystyle l<\infty. Suppose l>0\displaystyle l>0, and let

ψ0(xn,yn)ylasyn0.\frac{\partial\psi_{0}(x_{n},y_{n})}{\partial y}\rightarrow l\ \ \text{as}\ \ y_{n}\downarrow 0.

Choose a blow-up sequence ψ0(xn+ynx,yny)yn\displaystyle\frac{\psi_{0}(x_{n}+y_{n}x,y_{n}y)}{y_{n}} with respect to Byn(xn,0)\displaystyle B_{y_{n}}(x_{n},0), let ψ00\displaystyle\psi_{00} be the blow-up limit. Using above arguments again, we have

ψ00(x,y)=lyin{y>0},andψ00(x,y)=κyin{y<0}.\psi_{00}(x,y)=ly\ \ \text{in}\ \ \{y>0\},\ \ \text{and}\ \ \psi_{00}(x,y)=-\kappa y\ \ \text{in}\ \ \{y<0\}.

Since ψ00\displaystyle\psi_{00} is a minimizer for (3.13) and (x,0)\displaystyle(x,0) is a free boundary point of ψ00\displaystyle\psi_{00}, we can show that the set {ψ00=0}\displaystyle\{\psi_{00}=0\} has density zero at any point (x,0)\displaystyle(x,0), which contradicts to Lemma 3.7 in [2].

Therefore, we obtain that l=0\displaystyle l=0 and ψ0(x,y)=o(y)\displaystyle\psi_{0}(x,y)=o(y) as y0\displaystyle y\downarrow 0. For any ε>0\displaystyle\varepsilon>0, one has

1rBr(X0)ψ0𝑑S<ε,where X0=(x0,y0) and r=y0,\frac{1}{r}\fint_{\partial B_{r}(X_{0})}\psi_{0}dS<\varepsilon,\ \ \text{where $\displaystyle X_{0}=(x_{0},y_{0})$ and $\displaystyle r=y_{0}$},

provided that y0>0\displaystyle y_{0}>0 is small enough. It follows from the non-degeneracy Lemma 3.4 in [2] that ψ0=0\displaystyle\psi_{0}=0 in some strip {0<y<ε0}\displaystyle\{0<y<\varepsilon_{0}\}. Thus, the proof of the claim (3.32) is done.

Finally, noting that ψ0\displaystyle\psi_{0} is a minimizer to the variational problem (3.13), by means of Theorem 2.5 in [2], we can conclude that |ψ0|=λ(X0)\displaystyle|\nabla\psi_{0}|=\lambda(X_{0}) on the free boundary of ψ0\displaystyle\psi_{0}, and thus λ(X0)=κ\displaystyle\lambda(X_{0})=\kappa.

Step 2. In this step, we will show that there exists a constant C>0\displaystyle C>0, such that

supXBr(X0)|ψ(X)|supXBR(X0)λ(X)+C(rR)γ,\sup_{X\in B_{r}(X_{0})}|\nabla\psi(X)|\leq\sup_{X\in B_{R}(X_{0})}\lambda(X)+C\left(\frac{r}{R}\right)^{\gamma}, (3.33)

for any disc Br(X0)BR(X0)D\displaystyle B_{r}(X_{0})\subset B_{R}(X_{0})\subset D. Denote Br=Br(X0)\displaystyle B_{r}=B_{r}(X_{0}) and BR=BR(X0)\displaystyle B_{R}=B_{R}(X_{0}) for simplicity.

Define 𝒬ε(Y)=max{|ψ(Y)|2λ02ε,0}\displaystyle\mathcal{Q}_{\varepsilon}(Y)=\max\{|\nabla\psi(Y)|^{2}-\lambda_{0}^{2}-\varepsilon,0\} for any ε>0\displaystyle\varepsilon>0, where λ0=supXBRλ(X)\displaystyle\lambda_{0}=\sup_{X\in B_{R}}\lambda(X). It is easy to check that

Δ𝒬ε=2|D2ψ|22f(ψ)|ψ|20in Ω{ψ>0}.\Delta\mathcal{Q}_{\varepsilon}=2|D^{2}\psi|^{2}-2f^{\prime}(\psi)|\nabla\psi|^{2}\geq 0\ \ \text{in $\displaystyle\Omega\cap\{\psi>0\}$}.

And thus 𝒬ε\displaystyle\mathcal{Q}_{\varepsilon} is subharmonic function in Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\}. It follows from (3.30) that 𝒬ε=0\displaystyle\mathcal{Q}_{\varepsilon}=0 in a small neighborhood of the free boundary Γ\displaystyle\Gamma. We extend 𝒬ε\displaystyle\mathcal{Q}_{\varepsilon} by 0\displaystyle 0 and set

Pε(r)=supYBr𝒬ε(Y)for any r>0.P_{\varepsilon}(r)=\sup_{Y\in B_{r}}\mathcal{Q}_{\varepsilon}(Y)\ \ \text{for any $\displaystyle r>0$}.

Then Pε(r)𝒬ε\displaystyle P_{\varepsilon}(r)-\mathcal{Q}_{\varepsilon} is superharmonic in Br\displaystyle B_{r}. Furthermore, Pε(r)𝒬ε0\displaystyle P_{\varepsilon}(r)-\mathcal{Q}_{\varepsilon}\geq 0 in Br\displaystyle B_{r} and Pε(r)𝒬ε=Pε(r)\displaystyle P_{\varepsilon}(r)-\mathcal{Q}_{\varepsilon}=P_{\varepsilon}(r) in Br{ψ=0}\displaystyle B_{r}\cap\{\psi=0\}. It follows from Theorem 8.26 in [26] that

infYBr2(Pε(r)𝒬ε(Y))cr2Pε(r)𝒬εL1(Br)cr2Pε(r)𝒬εL1(Br{ψ=0})cPε(r),\begin{array}[]{rl}\inf_{Y\in B_{\frac{r}{2}}}(P_{\varepsilon}(r)-\mathcal{Q}_{\varepsilon}(Y))\geq&cr^{-2}\|P_{\varepsilon}(r)-\mathcal{Q}_{\varepsilon}\|_{L^{1}(B_{r})}\\ \geq&cr^{-2}\|P_{\varepsilon}(r)-\mathcal{Q}_{\varepsilon}\|_{L^{1}(B_{r}\cap\{\psi=0\})}\\ \geq&cP_{\varepsilon}(r),\end{array} (3.34)

where we have used the fact 2(Br{ψ=0})cr2\displaystyle\mathcal{L}^{2}(B_{r}\cap\{\psi=0\})\geq cr^{2} in Lemma 2.7. Taking ε0\displaystyle\varepsilon\rightarrow 0 in (3.34), we have

infYBr2(P0(r)𝒬0(Y))cP0(r), 0<c<1,\inf_{Y\in B_{\frac{r}{2}}}(P_{0}(r)-\mathcal{Q}_{0}(Y))\geq cP_{0}(r),\ \ \ 0<c<1,

which implies that

supYBr2𝒬0(Y)(1c)P0(r).\sup_{Y\in B_{\frac{r}{2}}}\mathcal{Q}_{0}(Y)\leq(1-c)P_{0}(r).

Thus we have

P0(r2)(1c)P0(r).P_{0}\left(\frac{r}{2}\right)\leq(1-c)P_{0}(r).

It follows from Lemma 8.23 in [26] that

P0(r)C(rR)γP0(R),P_{0}(r)\leq C\left(\frac{r}{R}\right)^{\gamma}P_{0}(R),

which gives that

supXBr|ψ(X)|λ0+C(rR)γ,γ(0,1),\sup_{X\in B_{r}}|\nabla\psi(X)|\leq\lambda_{0}+C\left(\frac{r}{R}\right)^{\gamma},\ \ \gamma\in(0,1),

for small r>0\displaystyle r>0.

Step 3. For any disc Br(X)\displaystyle B_{r}(X) touching the free boundary, there exists a free boundary point X0Br(X)\displaystyle X_{0}\in B_{r}(X), such that Br(X)B2r(X0)\displaystyle B_{r}(X)\subset B_{2r}(X_{0}). Since λ(X)C0,β(Ω)\displaystyle\lambda(X)\in C^{0,\beta}(\Omega), the gradient estimate (3.28) gives that

|ψ(X)|supYB2r(X0)|ψ(Y)|supYB2r(X0)λ(Y)+Crβ2λ(X)+Crα0,|\nabla\psi(X)|\leq\sup_{Y\in B_{2r}(X_{0})}|\nabla\psi(Y)|\leq\sup_{Y\in B_{2\sqrt{r}}(X_{0})}\lambda(Y)+Cr^{\frac{\beta}{2}}\leq\lambda(X)+Cr^{\alpha_{0}}, (3.35)

where α0=12min{γ,β}\displaystyle\alpha_{0}=\frac{1}{2}\min\{\gamma,\beta\}, provided that r\displaystyle r is small.

To obtain the linear growth (3.27) of ψ\displaystyle\psi near the free boundary, we next show that

Lemma 3.6.

For any compact subset D\displaystyle D of Ω\displaystyle\Omega and disc Br(X0)D\displaystyle B_{r}(X_{0})\subset D with X0Ω{ψ>0}\displaystyle X_{0}\in\Omega\cap\partial\{\psi>0\}, then

Br(X0){ψ>0}max{λ2(Y)|ψ(Y)|2,0}𝑑Y0as r0.\fint_{B_{r}(X_{0})\cap\{\psi>0\}}\max\{\lambda^{2}(Y)-|\nabla\psi(Y)|^{2},0\}dY\rightarrow 0\ \ \text{as $\displaystyle r\rightarrow 0$}. (3.36)
Proof.

Without loss of generality, we assume that X0=0\displaystyle X_{0}=0, and denote Br=Br(0)\displaystyle B_{r}=B_{r}(0) for simplicity. For any ξC0(Ω)\displaystyle\xi\in C_{0}^{\infty}(\Omega) with ξ0\displaystyle\xi\geq 0, define a function

ψε=max{ψεξ,0}for any ε>0.\psi_{\varepsilon}=\max\{\psi-\varepsilon\xi,0\}\ \ \text{for any $\displaystyle\varepsilon>0$}.

Since ψεK\displaystyle\psi_{\varepsilon}\in K, we have that J(ψ)J(ψε)\displaystyle J(\psi)\leq J(\psi_{\varepsilon}), namely,

0Ω|ψ|2|ψε|2+F(ψ)F(ψε)dX+Ωλ2(I{ψ>0}I{ψε>0})𝑑XΩ|min{ψ,εξ}|2+2ψmin{ψ,εξ}F(ψ)min{ψ,εξ}Λ|min{ψ,εξ}|2dX+Ω{0<ψεξ}λ2𝑑X=Ω|min{ψ,εξ}|2+Λ|min{ψ,εξ}|2dX+Ω{0<ψεξ}λ2𝑑X.\begin{array}[]{rl}0\geq&\int_{\Omega}|\nabla\psi|^{2}-|\nabla\psi_{\varepsilon}|^{2}+F(\psi)-F(\psi_{\varepsilon})dX+\int_{\Omega}\lambda^{2}(I_{\{\psi>0\}}-I_{\{\psi_{\varepsilon}>0\}})dX\\ \geq&\int_{\Omega}-|\nabla\min\{\psi,\varepsilon\xi\}|^{2}+2\nabla\psi\cdot\nabla\min\{\psi,\varepsilon\xi\}-F^{\prime}(\psi)\min\{\psi,\varepsilon\xi\}\\ &-\Lambda|\min\{\psi,\varepsilon\xi\}|^{2}dX+\int_{\Omega\cap\{0<\psi\leq\varepsilon\xi\}}\lambda^{2}dX\\ =&-\int_{\Omega}|\nabla\min\{\psi,\varepsilon\xi\}|^{2}+\Lambda|\min\{\psi,\varepsilon\xi\}|^{2}dX+\int_{\Omega\cap\{0<\psi\leq\varepsilon\xi\}}\lambda^{2}dX.\end{array}

This gives that

Ω{0<ψεξ}λ2|ψ|2dXε2Ω{ψ>εξ}|ξ|2𝑑X+ΛΩ|min{ψ,εξ}|2𝑑X.\int_{\Omega\cap\{0<\psi\leq\varepsilon\xi\}}\lambda^{2}-|\nabla\psi|^{2}dX\leq\varepsilon^{2}\int_{\Omega\cap\{\psi>\varepsilon\xi\}}|\nabla\xi|^{2}dX+\Lambda\int_{\Omega}|\min\{\psi,\varepsilon\xi\}|^{2}dX. (3.37)

Taking BrBρ=Bρ(0)BR\displaystyle B_{r}\subset B_{\rho}=B_{\rho}(0)\subset B_{R}, the Lipschitz continuity of ψ\displaystyle\psi gives that

ψ(X)=ψ(X)ψ(0)Cr\displaystyle\psi(X)=\psi(X)-\psi(0)\leq Cr in Br\displaystyle B_{r}.

Taking ε=Cr\displaystyle\varepsilon=Cr and

ξ(X)={0forXΩBρ,lnρln|X|lnρlnrforXBρBr,1forBr.\xi(X)=\left\{\begin{array}[]{ll}0&\text{for}\ \ X\in\Omega\setminus B_{\rho},\\ \frac{\ln\rho-\ln|X|}{\ln\rho-\ln r}&\text{for}~~X\in B_{\rho}\setminus B_{r},\\ 1&\text{for}~B_{r}.\end{array}\right.

It follows from (3.37) that

Bρ{0<ψεξ}λ2|ψ|2dXε2Bρ{ψ>εξ}|ξ|2𝑑X+ΛBρ|min{ψ,εξ}|2𝑑XCr2Bρ{ψ>εξ}|ξ|2𝑑X+Cρ2r2Cr2lnρlnr+Cρ2r2.\begin{array}[]{rl}\int_{B_{\rho}\cap\{0<\psi\leq\varepsilon\xi\}}\lambda^{2}-|\nabla\psi|^{2}dX\leq&\varepsilon^{2}\int_{B_{\rho}\cap\{\psi>\varepsilon\xi\}}|\nabla\xi|^{2}dX+\Lambda\int_{B_{\rho}}|\min\{\psi,\varepsilon\xi\}|^{2}dX\\ \leq&Cr^{2}\int_{B_{\rho}\cap\{\psi>\varepsilon\xi\}}|\nabla\xi|^{2}dX+C\rho^{2}r^{2}\\ \leq&\frac{Cr^{2}}{\ln\rho-\ln r}+C\rho^{2}r^{2}.\end{array} (3.38)

Since ψCr=εξ\displaystyle\psi\leq Cr=\varepsilon\xi in Br\displaystyle B_{r}, it is easy to check that

Bρ{0<ψεξ}max{λ2|ψ|2,0}𝑑XBr{0<ψεξ}max{λ2|ψ|2,0}𝑑X=Br{ψ>0}max{λ2|ψ|2,0}𝑑X.\begin{array}[]{rl}\int_{B_{\rho}\cap\{0<\psi\leq\varepsilon\xi\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dX\geq&\int_{B_{r}\cap\{0<\psi\leq\varepsilon\xi\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dX\\ =&\int_{B_{r}\cap\{\psi>0\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dX.\end{array} (3.39)

By virtue of (3.28), (3.38) and (3.39), we have

Bρ{0<ψεξ}max{λ2|ψ|2,0}𝑑X=Bρ{0<ψεξ}max{|ψ|2λ2,0}𝑑X+Bρ{0<ψεξ}λ2|ψ|2dXBρ{0<ψεξ}max{|ψ|2λ2,0}𝑑X+Cr2lnρlnr+Cρ2r2Bρ{ψ>0}max{|ψ|2λ2,0}𝑑X+Cr2lnρlnr+Cρ2r2Cρ2((ρR)α+λR)+Cr2lnρlnr+Cρ2r2,\begin{array}[]{rl}&\int_{B_{\rho}\cap\{0<\psi\leq\varepsilon\xi\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dX\\ =&\int_{B_{\rho}\cap\{0<\psi\leq\varepsilon\xi\}}\max\{|\nabla\psi|^{2}-\lambda^{2},0\}dX+\int_{B_{\rho}\cap\{0<\psi\leq\varepsilon\xi\}}\lambda^{2}-|\nabla\psi|^{2}dX\\ \leq&\int_{B_{\rho}\cap\{0<\psi\leq\varepsilon\xi\}}\max\{|\nabla\psi|^{2}-\lambda^{2},0\}dX+\frac{Cr^{2}}{\ln\rho-\ln r}+C\rho^{2}r^{2}\\ \leq&\int_{B_{\rho}\cap\{\psi>0\}}\max\{|\nabla\psi|^{2}-\lambda^{2},0\}dX+\frac{Cr^{2}}{\ln\rho-\ln r}+C\rho^{2}r^{2}\\ \leq&C\rho^{2}\left(\left(\frac{\rho}{R}\right)^{\alpha}+\lambda_{R}\right)+\frac{Cr^{2}}{\ln\rho-\ln r}+C\rho^{2}r^{2},\end{array}

where λR=supX,YBR(X0)|λ(X)λ(Y)|\displaystyle\lambda_{R}=\sup_{X,Y\in B_{R}(X_{0})}|\lambda(X)-\lambda(Y)|, which gives that

Br{ψ>0}max{λ2|ψ|2,0}𝑑XC(ρr)2((ρR)γ+λR)+Clnρr+Cρ2,\begin{array}[]{rl}\fint_{B_{r}\cap\{\psi>0\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dX\leq C\left(\frac{\rho}{r}\right)^{2}\left(\left(\frac{\rho}{R}\right)^{\gamma}+\lambda_{R}\right)+\frac{C}{\ln\frac{\rho}{r}}+C\rho^{2},\end{array} (3.40)

Taking ρ=R2\displaystyle\rho=R^{2} and r=R2(λR+Rγ)14\displaystyle r=R^{2}\left(\lambda_{R}+R^{\gamma}\right)^{\frac{1}{4}}, it follows from (3.40) that we have

Br{ψ>0}max{λ2|ψ|2,0}𝑑XC(λR+Rγ)12+Cln1λR+Rγ+CR4,\begin{array}[]{rl}\fint_{B_{r}\cap\{\psi>0\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dX\leq C\left(\lambda_{R}+R^{\gamma}\right)^{\frac{1}{2}}+\frac{C}{\ln{\frac{1}{\lambda_{R}+R^{\gamma}}}}+CR^{4},\end{array}

for sufficiently small r>0\displaystyle r>0, which gives (3.36).

With the aid of Lemma 3.6, we have

Lemma 3.7.

For any blow-up limit ψ0\displaystyle\psi_{0} of ψ\displaystyle\psi at X0Γ\displaystyle X_{0}\in\Gamma, ψ0\displaystyle\psi_{0} is a half plane solution with slope λ(X0)\displaystyle\lambda(X_{0}). Furthermore, the linear growth (3.27) holds.

Proof.

Let ψn(X)=ψ(X0+ρnX)ρn\displaystyle\psi_{n}(X)=\frac{\psi(X_{0}+\rho_{n}X)}{\rho_{n}} be a blow-up sequence with respect to BρnR(X0)\displaystyle B_{\rho_{n}R}(X_{0}) for any R>0\displaystyle R>0, and λn(X)=λ(X0+ρnX)\displaystyle\lambda_{n}(X)=\lambda(X_{0}+\rho_{n}X) with XBR(0)\displaystyle X\in B_{R}(0), it follows from Lemma 3.5 and Lemma 3.6 that

supXBR(0)|ψn(X)|=supYBρnR(X0)|ψ(Y)|supYBρnR(X0)λ(Y)+Cρnγ2λ(X0),\sup_{X\in B_{R}(0)}|\nabla\psi_{n}(X)|=\sup_{Y\in B_{\rho_{n}R}(X_{0})}|\nabla\psi(Y)|\leq\sup_{Y\in B_{\sqrt{\rho_{n}}R}(X_{0})}\lambda(Y)+C\rho_{n}^{\frac{\gamma}{2}}\rightarrow\lambda(X_{0}),

and

BR(0){ψn>0}max{λn2|ψn|2,0}𝑑X=BρnR(X0){ψ>0}max{λ2|ψ|2,0}𝑑Y0,\fint_{B_{R}(0)\cap\{\psi_{n}>0\}}\max\{\lambda_{n}^{2}-|\nabla\psi_{n}|^{2},0\}dX=\fint_{B_{\rho_{n}R}(X_{0})\cap\{\psi>0\}}\max\{\lambda^{2}-|\nabla\psi|^{2},0\}dY\rightarrow 0,

as ρn0\displaystyle\rho_{n}\rightarrow 0. Those imply that for any blow-up limit ψ0\displaystyle\psi_{0}, we have

|ψ0|=λ(X0)in{ψ0>0}.|\nabla\psi_{0}|=\lambda(X_{0})\ \ \text{in}\ \ \{\psi_{0}>0\}.

On the other hand, ψ0\displaystyle\psi_{0} is harmonic in {ψ0>0}\displaystyle\{\psi_{0}>0\}, thus ψ0\displaystyle\nabla\psi_{0} must be invariant in each connected component of {ϕ0>0}\displaystyle\{\phi_{0}>0\}. In fact, applying the operator x\displaystyle\partial_{x} and y\displaystyle\partial_{y} to 12|ψ0|2=λ2(X0)2\displaystyle\frac{1}{2}|\nabla\psi_{0}|^{2}=\frac{\lambda^{2}(X_{0})}{2}, respectively, we have

{xψ0xxψ0+yψ0xyψ0=0,xψ0xyψ0+yψ0yyψ0=0.\left\{\begin{array}[]{rl}\partial_{x}\psi_{0}\partial_{xx}\psi_{0}+\partial_{y}\psi_{0}\partial_{xy}\psi_{0}=0,\\ \partial_{x}\psi_{0}\partial_{xy}\psi_{0}+\partial_{y}\psi_{0}\partial_{yy}\psi_{0}=0.\end{array}\right.

This together with the fact Δψ0=0\displaystyle\Delta\psi_{0}=0 gives that

𝕋𝐗=(0,0,0)tr,\mathbb{T}\mathbf{X}=(0,0,0)^{tr},

where the vector 𝐗=(xxψ0,xyψ0,yyψ0)tr\displaystyle\mathbf{X}=(\partial_{xx}\psi_{0},\partial_{xy}\psi_{0},\partial_{yy}\psi_{0})^{tr} and matrix 𝕋\displaystyle\mathbb{T} as

𝕋=(101xψ0yψ000xψ0yψ0).\mathbb{T}=\left(\begin{matrix}1&0&1\\ \partial_{x}\psi_{0}&\partial_{y}\psi_{0}&0\\ 0&\partial_{x}\psi_{0}&\partial_{y}\psi_{0}\end{matrix}\right).

Furthermore, one has

det𝕋=|ψ0|2=λ2(X0)>0,\det\mathbb{T}=|\nabla\psi_{0}|^{2}=\lambda^{2}(X_{0})>0,

which implies

𝐗=(xxψ0,xyψ0,yyψ0)=(0,0,0).\mathbf{X}=(\partial_{xx}\psi_{0},\partial_{xy}\psi_{0},\partial_{yy}\psi_{0})=(0,0,0).

Thus, ψ0(X)\displaystyle\psi_{0}(X) is a linear function, there exist a unit vector ν0\displaystyle\nu_{0}, and two numbers γ10\displaystyle\gamma_{1}\geq 0 and γ20\displaystyle\gamma_{2}\geq 0, such that one has

ψ0(X)=γ1max{Xν0,0}+γ2max{Xν0,0}.\psi_{0}(X)=\gamma_{1}\max\{-X\cdot\nu_{0},0\}+\gamma_{2}\max\{X\cdot\nu_{0},0\}. (3.41)

Since ψ0\displaystyle\psi_{0} is a local minimizer, it follows from Lemma 3.7 in [2] that the set {ψ0=0}\displaystyle\{\psi_{0}=0\} has a positive measure, this gives that γ1=λ(X0)\displaystyle\gamma_{1}=\lambda(X_{0}) and γ2=0\displaystyle\gamma_{2}=0. In view of (3.41), one has

ψ0(X)=λ(X0)max{Xν0,0},\psi_{0}(X)=\lambda(X_{0})\max\{-X\cdot\nu_{0},0\},

which implies the linear growth (3.27). ∎

3.4. Flatness of the free boundary

In this subsection, we will study the regularity of the free boundary Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\}, and obtain some flatness property of the free boundary point.

First, we introduce the relevant flatness class of the free boundary point (See the definition 5.1 in [2]).

Definition 3.1.

(Flat free boundary points) Let 0<σ+,σ1\displaystyle 0<\sigma_{+},\sigma_{-}\leq 1 and δ>0\displaystyle\delta>0. We say that ψ\displaystyle\psi is of the flatness class (σ+,σ;δ)\displaystyle\mathcal{F}(\sigma_{+},\sigma_{-};\delta) in Bρ=Bρ(X0)\displaystyle B_{\rho}=B_{\rho}(X_{0}) with a unit vector ν\displaystyle\nu (see Figure 2) if

(i) ψ\displaystyle\psi is a minimizer to the variational problem (P)\displaystyle(P).

(ii) X0Γ\displaystyle X_{0}\in\Gamma and

ψ(X)=0for(XX0)νσ+ρ,\psi(X)=0\ \ \text{for}\ \ (X-X_{0})\cdot\nu\geq\sigma_{+}\rho,

and

ψ(X)λ(X0)((XX0)ν+σρ)for(XX0)νσρ.\psi(X)\geq-\lambda(X_{0})((X-X_{0})\cdot\nu+\sigma_{-}\rho)\ \ \text{for}\ \ (X-X_{0})\cdot\nu\leq-\sigma_{-}\rho.

(iii) |ψ|λ(X0)(1+δ)\displaystyle|\nabla\psi|\leq\lambda(X_{0})(1+\delta) in Bρ\displaystyle B_{\rho} and supX,YBρ|λ(X)λ(Y)|λ(X0)δ\displaystyle\sup_{X,Y\in B_{\rho}}|\lambda(X)-\lambda(Y)|\leq\lambda(X_{0})\delta.

Refer to caption
Figure 2. Flatness free boundary point

We next show that the flatness from above implies the flatness from below.

Lemma 3.8.

There exists a positive constant C=C(Λ,c0)\displaystyle C=C(\Lambda,c_{0}), such that for any small σ>0\displaystyle\sigma>0, if ψ(σ,1;δ)\displaystyle\psi\in\mathcal{F}(\sigma,1;\delta) in Bρ(X0)\displaystyle B_{\rho}(X_{0}) and δc0σ,ρc0σ\displaystyle\delta\leq c_{0}\sigma,\rho\leq c_{0}\sigma, then ψ(2σ,Cσ;δ)\displaystyle\psi\in\mathcal{F}(2\sigma,C\sigma;\delta) in Bρ2(X0)\displaystyle B_{\frac{\rho}{2}}(X_{0}) (see Figure 3).

Refer to caption
Figure 3. Flatness from below
Proof.

We divide two steps to show this lemma.

Step 1. In this step, we will show that

lim supXX0ψ(X)dist(X,B)=λ(X0),if B is a disc in {ψ=0} touching {ψ>0} at X0.\limsup_{X\rightarrow X_{0}}\frac{\psi(X)}{\text{dist}(X,B)}=\lambda(X_{0}),\ \text{if $\displaystyle B$ is a disc in $\displaystyle\{\psi=0\}$ touching $\displaystyle\partial\{\psi>0\}$ at $\displaystyle X_{0}$}. (3.42)

Denote lim supXX0ψ(X)dist(X,B)=κ\displaystyle\limsup_{X\rightarrow X_{0}}\frac{\psi(X)}{\text{dist}(X,B)}=\kappa, it suffices to show that κ=λ(X0)\displaystyle\kappa=\lambda(X_{0}). Taking a sequence {Yn}\displaystyle\{Y_{n}\} with YnX0\displaystyle Y_{n}\rightarrow X_{0}, such that ψ(Yn)>0\displaystyle\psi(Y_{n})>0, dn=dist(Yn,B)\displaystyle d_{n}=\text{dist}(Y_{n},B) and

ψ(Yn)dnκ.\frac{\psi(Y_{n})}{d_{n}}\rightarrow\kappa.

It follows from the non-degeneracy Lemma 2.6 that κ>0\displaystyle\kappa>0. Consider a blow-up sequence ψn(X)=ψ(Xn+dnX)dn\displaystyle\psi_{n}(X)=\frac{\psi(X_{n}+d_{n}X)}{d_{n}} with respect to Bdn(Xn)\displaystyle B_{d_{n}}(X_{n}), where XnB\displaystyle X_{n}\in\partial B and |XnYn|=dn\displaystyle|X_{n}-Y_{n}|=d_{n}. Taking a subsequence {ψn}\displaystyle\{\psi_{n}\} with a blow-up limit ψ0C0,1(2)\displaystyle\psi_{0}\in C^{0,1}(\mathbb{R}^{2}), such that

XnYndnνandψnψ0uniformly in any compact subset of 2.\frac{X_{n}-Y_{n}}{d_{n}}\rightarrow\nu\ \ \text{and}\ \ \psi_{n}\rightarrow\psi_{0}\ \ \text{uniformly in any compact subset of $\displaystyle\mathbb{R}^{2}$}.

It is easy to check that ψ0(ν)=κ\displaystyle\psi_{0}(-\nu)=\kappa. We next claim that

ψ0(X)=κmax{Xν,0}.\psi_{0}(X)=\kappa\max\{-X\cdot\nu,0\}. (3.43)

In fact, the definition of ψ0\displaystyle\psi_{0} gives that

ψ0(X)κXνforXν0,andψ0(X)=0forXν0.\psi_{0}(X)\leq-\kappa X\cdot\nu\ \ \text{for}\ \ X\cdot\nu\leq 0,\ \ \text{and}\ \ \psi_{0}(X)=0\ \ \text{for}\ \ X\cdot\nu\geq 0.

Since ψ0\displaystyle\psi_{0} is harmonic in {ψ0>0}\displaystyle\{\psi_{0}>0\} and ψ0(ν)=κ\displaystyle\psi_{0}(-\nu)=\kappa, the strong maximum principle gives the claim (3.43).

With the aid of the claim (3.43), it follows from Lemma 3.7 that λ(X0)=κ\displaystyle\lambda(X_{0})=\kappa.

Step 2. Without loss of generality, we assume that X0=0\displaystyle X_{0}=0 and ν=e2=(0,1)\displaystyle\nu=e_{2}=(0,1). Set ψρ(X)=ψ(ρX)ρ\displaystyle\psi_{\rho}(X)=\frac{\psi(\rho X)}{\rho}, λρ(X)=λ(ρX)\displaystyle\lambda_{\rho}(X)=\lambda(\rho X) and λ0=λ(0)\displaystyle\lambda_{0}=\lambda(0), one has

Δψρ+ρf(ρψρ)=0in{ψρ>0},and ψρF(σ,1;δ) in B1.\Delta\psi_{\rho}+\rho f(\rho\psi_{\rho})=0\ \ \text{in}\ \ \{\psi_{\rho}>0\},\ \text{and $\displaystyle\psi_{\rho}\in F(\sigma,1;\delta)$ in $\displaystyle B_{1}$.} (3.44)

Let

η(x)={0for|x|13,e9x219x2for|x|<13,\eta(x)=\left\{\begin{array}[]{ll}0&\text{for}\ \ |x|\geq\frac{1}{3},\\ e^{-\frac{9x^{2}}{1-9x^{2}}}&\text{for}\ \ |x|<\frac{1}{3},\end{array}\right.

and choose s0\displaystyle s\geq 0 be the maximum with the property

B1{ψρ>0}Dσ={XB1y<σsη(x)}.B_{1}\cap\{\psi_{\rho}>0\}\subset D_{\sigma}=\{X\in B_{1}\mid y<\sigma-s\eta(x)\}.

Thus there exists a point ZB12Dσ{ψρ>0}\displaystyle Z\in B_{\frac{1}{2}}\cap\partial D_{\sigma}\cap\partial\{\psi_{\rho}>0\}. It is easy to check that sσ\displaystyle s\leq\sigma, due to 0Γ\displaystyle 0\in\Gamma. Define a function ϕ(x,y)\displaystyle\phi(x,y) solving the problem

{Δϕ+ρf(0)=0inDσ,ϕ=0onDσB1,ϕ=λ0(1+2σ)(σy)onDσB1.\left\{\begin{array}[]{ll}&\Delta\phi+\rho f(0)=0~~~~\text{in}~~D_{\sigma},\\ &\phi=0~~~~\text{on}~~\partial D_{\sigma}\cap B_{1},\ \ \ \ \ \phi=\lambda_{0}(1+2\sigma)(\sigma-y)~~~~\text{on}~~\partial D_{\sigma}\setminus B_{1}.\end{array}\right.

Noticing that λ0(1+2σ)(σy)+λ0(1+σ)y0\displaystyle\lambda_{0}(1+2\sigma)(\sigma-y)+\lambda_{0}(1+\sigma)y\geq 0 on Dσ\displaystyle\partial D_{\sigma}, the maximum principle gives that

ϕ(X)λ0(1+σ)yinDσ.\phi(X)\geq-\lambda_{0}(1+\sigma)y\ \ \text{in}\ \ D_{\sigma}. (3.45)

Next, we will show that

νϕ(Z)λ0(1+Cσ).\partial_{-\nu}\phi(Z)\leq\lambda_{0}(1+C\sigma). (3.46)

To see this, define a function ϕ1\displaystyle\phi_{1} as follows

ϕ1(x,y)=μ2μ1(1eμ1g(x,y))inDσ,\phi_{1}(x,y)=\frac{\mu_{2}}{\mu_{1}}\left(1-e^{-\mu_{1}g(x,y)}\right)\ \ \text{in}\ \ D_{\sigma},

where g(x,y)=y+σsη(x)\displaystyle g(x,y)=-y+\sigma-s\eta(x) and the constants μ1\displaystyle\mu_{1} and μ2λ0\displaystyle\mu_{2}\geq\lambda_{0} depending on σ\displaystyle\sigma will be determined later. It is easy to check that

1|g(x,y)|1+Cσand|D2g(x,y)|Cσ.1\leq|\nabla g(x,y)|\leq 1+C\sigma\ \ \text{and}\ \ |D^{2}g(x,y)|\leq C\sigma. (3.47)

Therefore, it follows from (3.47) and ρσ0σ\displaystyle\rho\leq\sigma_{0}\sigma that

Δϕ1+ρf(0)=μ2eμ1g(Δgμ1|g|2)+ρf(0)μ2eμ1g(Cσμ1)+ρf(0)<0,\Delta\phi_{1}+\rho f(0)=\mu_{2}e^{-\mu_{1}g}(\Delta g-\mu_{1}|\nabla g|^{2})+\rho f(0)\leq\mu_{2}e^{-\mu_{1}g}(C\sigma-\mu_{1})+\rho f(0)<0,

provided that

μ1=C1σ,C1 is sufficiently large and σ is small.\mu_{1}=C_{1}\sigma,\ \ \text{$\displaystyle C_{1}$ is sufficiently large and $\displaystyle\sigma$ is small}.

On the other hand, take μ2=λ0(1+C2σ)\displaystyle\mu_{2}=\lambda_{0}(1+C_{2}\sigma), C2\displaystyle C_{2} is sufficiently large and σ\displaystyle\sigma is small, one has

ϕ1(x,y)μ2g(x,y)(1Cμ1)λ0(1+2σ)g(x,y)for any (x,y)Dσ,\phi_{1}(x,y)\geq\mu_{2}g(x,y)(1-C\mu_{1})\geq\lambda_{0}(1+2\sigma)g(x,y)\ \ \text{for any $\displaystyle(x,y)\in\partial D_{\sigma}$},

which together with the maximum principle gives that

ϕ1(X)ϕ(X)in Dσ.\phi_{1}(X)\geq\phi(X)\ \ \text{in $\displaystyle D_{\sigma}$}.

Recalling that ϕ1(Z)=ϕ(Z)=0\displaystyle\phi_{1}(Z)=\phi(Z)=0, we have

νϕ(Z)νϕ1(Z)|ϕ1(Z)|=μ2|g(Z)|λ0(1+Cσ),σ is small.\partial_{-\nu}\phi(Z)\leq\partial_{-\nu}\phi_{1}(Z)\leq|\nabla\phi_{1}(Z)|=\mu_{2}|\nabla g(Z)|\leq\lambda_{0}(1+C\sigma),\ \ \text{$\displaystyle\sigma$ is small}.

In view of ψρF(σ,1;δ)\displaystyle\psi_{\rho}\in F(\sigma,1;\delta), one has that ψρϕ\displaystyle\psi_{\rho}\leq\phi on Dσ\displaystyle\partial D_{\sigma}. The maximum principle gives that ψρϕ\displaystyle\psi_{\rho}\leq\phi in Dσ\displaystyle D_{\sigma}.

Consider the points ξB34\displaystyle\xi\in\partial B_{\frac{3}{4}} with ξ=(ξ1,ξ2)\displaystyle\xi=(\xi_{1},\xi_{2}) and ξ2<12\displaystyle\xi_{2}<-\frac{1}{2}, and define a function φξ\displaystyle\varphi_{\xi} solving the following problem

{Δφξ=0inDσB110(ξ),φξ=0onDσ,φξ=λ0yonB110(ξ).\left\{\begin{array}[]{ll}&\Delta\varphi_{\xi}=0~~~~\text{in}~~D_{\sigma}\setminus B_{\frac{1}{10}}(\xi),\\ &\varphi_{\xi}=0~~~~\text{on}~~\partial D_{\sigma},\ \ \ \ \ \varphi_{\xi}=-\lambda_{0}y~~~~\text{on}~~\partial B_{\frac{1}{10}}(\xi).\end{array}\right.

Hopf’s Lemma gives that

νφξ(Z)cλ0>0.\partial_{-\nu}\varphi_{\xi}(Z)\geq c\lambda_{0}>0. (3.48)

Suppose that

ψρ(X)ϕ(X)+γσλ0yfor any XB110(ξ),\psi_{\rho}(X)\leq\phi(X)+\gamma\sigma\lambda_{0}y\ \ \text{for any $\displaystyle X\in B_{\frac{1}{10}}(\xi)$},

for a positive constant γ>0\displaystyle\gamma>0. It should be noted that Δ(ψρϕ+γσφξ)=ρf(ρψρ)+ρf(0)0\displaystyle\Delta(\psi_{\rho}-\phi+\gamma\sigma\varphi_{\xi})=-\rho f(\rho\psi_{\rho})+\rho f(0)\geq 0 in DσB110(ξ)\displaystyle D_{\sigma}\setminus B_{\frac{1}{10}}(\xi), and the maximum principle gives that

ψρϕγσφξinDσB110(ξ).\psi_{\rho}\leq\phi-\gamma\sigma\varphi_{\xi}\ \ \ \ \text{in}\ \ D_{\sigma}\setminus B_{\frac{1}{10}}(\xi).

Therefore, we have

λ0=lim supXZψρ(X)dist(X,Dσ)νϕ(Z)γσνφξ(Z)λ0(1+Cσcγσ),\lambda_{0}=\limsup_{X\rightarrow Z}\frac{\psi_{\rho}(X)}{\text{dist}(X,\partial D_{\sigma})}\leq\partial_{-\nu}\phi(Z)-\gamma\sigma\partial_{-\nu}\varphi_{\xi}(Z)\leq\lambda_{0}(1+C\sigma-c\gamma\sigma),

where ν\displaystyle\nu is the outer normal vector and we have used (3.48). This leads a contradiction, provided that γ\displaystyle\gamma is sufficiently large.

Hence, there exists some point Xξ=(xξ,yξ)B110(ξ)\displaystyle X_{\xi}=(x_{\xi},y_{\xi})\in B_{\frac{1}{10}}(\xi), such that

ψρ(Xξ)ϕ(Xξ)+Cσλ0yξ.\psi_{\rho}(X_{\xi})\geq\phi(X_{\xi})+C\sigma\lambda_{0}y_{\xi}. (3.49)

For any XB15(ξ)\displaystyle X\in B_{\frac{1}{5}}(\xi), it follows from (3.45), (3.49) and (iii) in Definition 3.1 that

ψρ(X)ψρ(Xξ)3λ0(1+δ)10λ0(1+σ)yξ+Cσλ0yξ3λ0(1+δ)10λ0(110Cσ)>0,\psi_{\rho}(X)\geq\psi_{\rho}(X_{\xi})-\frac{3\lambda_{0}(1+\delta)}{10}\geq-\lambda_{0}(1+\sigma)y_{\xi}+C\sigma\lambda_{0}y_{\xi}-\frac{3\lambda_{0}(1+\delta)}{10}\geq\lambda_{0}\left(\frac{1}{10}-C\sigma\right)>0,

provided that σ\displaystyle\sigma is small. Therefore, we have

Δ(ϕψρ)=ρf(0)+ρf(ρψρ)andϕψρ0inB15(ξ).\Delta(\phi-\psi_{\rho})=-\rho f(0)+\rho f(\rho\psi_{\rho})\ \ \text{and}\ \ \phi-\psi_{\rho}\geq 0\ \ \text{in}\ \ B_{\frac{1}{5}}(\xi).

Thanks to Harnack’s inequality in [31], one has

(ϕψρ)(ξ)(ϕψρ)(Xξ)+Cρf(0)f(ρψρ)L(B1)Cλ0σ+Cσ2Cλ0σ,(\phi-\psi_{\rho})(\xi)\leq(\phi-\psi_{\rho})(X_{\xi})+C\rho\|f(0)-f(\rho\psi_{\rho})\|_{L^{\infty}(B_{1})}\leq C\lambda_{0}\sigma+C\sigma^{2}\leq C\lambda_{0}\sigma,

if σ\displaystyle\sigma is small, which implies that

ψρ(ξ)ϕ(ξ)Cλ0σλ0((1+σ)ξ2Cσξ2Cσ.\psi_{\rho}(\xi)\geq\phi(\xi)-C\lambda_{0}\sigma\geq\lambda_{0}(-(1+\sigma)\xi_{2}-C\sigma\geq-\xi_{2}-C\sigma. (3.50)

It follows from (3.50) and (iii)\displaystyle(iii) in Definition 3.1 that

ψρ(ξ+te2)ψρ(ξ)λ0(1+δ)tλ0((ξ2+t)Cσ),\psi_{\rho}(\xi+te_{2})\geq\psi_{\rho}(\xi)-\lambda_{0}(1+\delta)t\geq\lambda_{0}(-(\xi_{2}+t)-C\sigma),

for t>0\displaystyle t>0 and ξ+te2B12\displaystyle\xi+te_{2}\in B_{\frac{1}{2}}, which implies that ψρ\displaystyle\psi_{\rho} is of (2σ,Cσ;δ)\displaystyle\mathcal{F}(2\sigma,C\sigma;\delta) in B12\displaystyle B_{\frac{1}{2}}. This completes the proof of this lemma. ∎

In the following, denote Br=Br(0)\displaystyle B_{r}=B_{r}(0), we will investigate the blow-up limit. Thanks to Lemma 7.3 and Corollary 7.4 in [2], we have the following non-homogeneous blow-up limit and we omit the proof here.

Lemma 3.9.

(Non-homogenous blow up)Let ψn(σn,σn;δn)\displaystyle\psi_{n}\in\mathcal{F}(\sigma_{n},\sigma_{n};\delta_{n}) in Bρn\displaystyle B_{\rho_{n}} with v=e2\displaystyle v=e_{2}, σn0\displaystyle\sigma_{n}\rightarrow 0, δnσn20\displaystyle\delta_{n}\sigma^{-2}_{n}\rightarrow 0 and ρnσn10\displaystyle\rho_{n}\sigma_{n}^{-1}\rightarrow 0. For any x(1,1)\displaystyle x\in(-1,1), set

bn+(x)=sup{h(ρnx,σnρnh){ψn>0}},b^{+}_{n}(x)=\sup\{h\mid(\rho_{n}x,\sigma_{n}\rho_{n}h)\in\partial\{\psi_{n}>0\}\},

and

bn(x)=inf{h(ρnx,σnρnh){ψn>0}},b^{-}_{n}(x)=\inf\{h\mid(\rho_{n}x,\sigma_{n}\rho_{n}h)\in\partial\{\psi_{n}>0\}\},

then for a subsequence,

b(x)=lim supzx,nbn+(z)=lim infzx,nbn(z)for any x(1,1).b(x)=\limsup_{z\rightarrow x,n\rightarrow\infty}b_{n}^{+}(z)=\liminf_{z\rightarrow x,n\rightarrow\infty}b_{n}^{-}(z)\ \ \text{for any $\displaystyle x\in(-1,1)$}.

Furthermore, bn+(x)b(x)\displaystyle b_{n}^{+}(x)\rightarrow b(x) and bn(x)b(x)\displaystyle b_{n}^{-}(x)\rightarrow b(x) uniformly in (1,1)\displaystyle(-1,1), b(0)=0\displaystyle b(0)=0 and b(x)\displaystyle b(x) is a continuous function.

Furthermore, we can show that the limit function b(x)\displaystyle b(x) is a convex function.

Lemma 3.10.

b(x)\displaystyle b(x) is convex with respect to x\displaystyle x.

Proof.

Set ψn(X)=ψ(ρnX)ρn\displaystyle\psi_{n}(X)=\frac{\psi(\rho_{n}X)}{\rho_{n}}, λn(X)=λ(ρnX)\displaystyle\lambda_{n}(X)=\lambda(\rho_{n}X) and λ0=λ(0)\displaystyle\lambda_{0}=\lambda(0), one has

Δψn+ρnf(ρnψn)=0in{ψn>0}and ψn(σn,σn;δn) in B1.\Delta\psi_{n}+\rho_{n}f(\rho_{n}\psi_{n})=0\ \ \text{in}\ \ \{\psi_{n}>0\}\ \ \text{and $\displaystyle\psi_{n}\in\mathcal{F}(\sigma_{n},\sigma_{n};\delta_{n})$ in $\displaystyle B_{1}$}.

If the assertion is not true, there exists a x0(1,1)\displaystyle x_{0}\in(-1,1) and a linear function g(x)\displaystyle g(x) in Iρ=(ρ+x0,ρ+x0)(1,1)\displaystyle I_{\rho}=(-\rho+x_{0},\rho+x_{0})\subset(-1,1), such that

g(x0±ρ)>b(x0±ρ) andb(x0)>g(x0).g(x_{0}\pm\rho)>b(x_{0}\pm\rho)\ \ \text{ and}\ b(x_{0})>g(x_{0}).

Denote

𝒟=Iρ×and𝒟+(k)={(x,y)𝒟k(x)<y<g0+2},\mathcal{D}=I_{\rho}\times\mathbb{R}\ \text{and}\ \mathcal{D}^{+}(k)=\{(x,y)\in\mathcal{D}\mid k(x)<y<g_{0}+2\},

and

𝒟0(k)={(x,y)𝒟y=k(x)}and𝒟(k)={(x,y)𝒟y<k(x)},\mathcal{D}^{0}(k)=\{(x,y)\in\mathcal{D}\mid y=k(x)\}\ \text{and}\ \mathcal{D}^{-}(k)=\{(x,y)\in\mathcal{D}\mid y<k(x)\},

for continuous function k(x)\displaystyle k(x), where g0=max{g(x0ρ),g0(x0+ρ)}\displaystyle g_{0}=\max\{g(x_{0}-\rho),g_{0}(x_{0}+\rho)\}. Define a test function dε,E(X)=min{dist(X,2E)ε,1}\displaystyle d_{\varepsilon,E}(X)=\min\left\{\frac{\text{dist}(X,\mathbb{R}^{2}\setminus E)}{\varepsilon},1\right\} for a set E\displaystyle E, it is easy to check that dε,E(X)\displaystyle d_{\varepsilon,E}(X) converges to the characteristic function IE(X)\displaystyle I_{E}(X) as ε0\displaystyle\varepsilon\rightarrow 0.

It follows from Proposition 3.3 that

Ωnψndε,E(𝒟+(σng))𝑑X+Ωn{ψn>0}ρnf(ρnψn)dε,E(𝒟+(σng))𝑑X=Ωnred{ψn>0}dε,E(𝒟+(σng))λn𝑑1,\begin{array}[]{rl}&-\int_{\Omega_{n}}\nabla\psi_{n}\cdot\nabla d_{\varepsilon,E}(\mathcal{D}^{+}(\sigma_{n}g))dX+\int_{\Omega_{n}\cap\{\psi_{n}>0\}}\rho_{n}f(\rho_{n}\psi_{n})d_{\varepsilon,E}(\mathcal{D}^{+}(\sigma_{n}g))dX\\ =&\int_{\Omega_{n}\cap\partial_{red}\{\psi_{n}>0\}}d_{\varepsilon,E}(\mathcal{D}^{+}(\sigma_{n}g))\lambda_{n}d\mathcal{H}^{1},\end{array} (3.51)

where Ωn={XρnXΩ}\displaystyle\Omega_{n}=\{X\mid\rho_{n}X\in\Omega\}.

Taking ε0\displaystyle\varepsilon\rightarrow 0 in (3.51), one has

𝒟+(σng){ψn>0}ψnνd1+𝒟+(σng){ψn>0}ρnf(ρnψn)𝑑X=𝒟+(σng)red{ψn>0}λn𝑑1,\begin{array}[]{rl}&-\int_{\partial\mathcal{D}^{+}(\sigma_{n}g)\cap\{\psi_{n}>0\}}\nabla\psi_{n}\cdot\nu d\mathcal{H}^{1}+\int_{\mathcal{D}^{+}(\sigma_{n}g)\cap\{\psi_{n}>0\}}\rho_{n}f(\rho_{n}\psi_{n})dX\\ =&\int_{\mathcal{D}^{+}(\sigma_{n}g)\cap\partial_{red}\{\psi_{n}>0\}}\lambda_{n}d\mathcal{H}^{1},\end{array}

which gives that

1(𝒟+(σng)red{ψn>0})1λ0(1δn)𝒟+(σng)red{ψn>0}λn𝑑11λ0(1δn)(𝒟+(σng){ψn>0}|ψn|𝑑1+Cρn𝒟+(σng){ψn>0}𝑑X)1+δn1δn1(𝒟0(σng){ψn>0})+Cρnσnλ0(1δn).\begin{array}[]{rl}&\mathcal{H}^{1}(\mathcal{D}^{+}(\sigma_{n}g)\cap\partial_{red}\{\psi_{n}>0\})\\ \leq&\frac{1}{\lambda_{0}(1-\delta_{n})}\int_{\mathcal{D}^{+}(\sigma_{n}g)\cap\partial_{red}\{\psi_{n}>0\}}\lambda_{n}d\mathcal{H}^{1}\\ \leq&\frac{1}{\lambda_{0}(1-\delta_{n})}\left(\int_{\partial\mathcal{D}^{+}(\sigma_{n}g)\cap\{\psi_{n}>0\}}|\nabla\psi_{n}|d\mathcal{H}^{1}+C\rho_{n}\int_{\mathcal{D}^{+}(\sigma_{n}g)\cap\{\psi_{n}>0\}}dX\right)\\ \leq&\frac{1+\delta_{n}}{1-\delta_{n}}\mathcal{H}^{1}(\mathcal{D}^{0}(\sigma_{n}g)\cap\{\psi_{n}>0\})+\frac{C\rho_{n}\sigma_{n}}{\lambda_{0}(1-\delta_{n})}.\end{array} (3.52)

Denote

En={ψn>0}𝒟(σng).E_{n}=\{\psi_{n}>0\}\cup\mathcal{D}^{-}(\sigma_{n}g).

It is easy to check that En\displaystyle E_{n} has finite perimeter in 𝒟\displaystyle\mathcal{D} and

1(𝒟redEn)1(𝒟+(σng)red{ψn>0})+1(𝒟0(σng){ψn=0}).\begin{array}[]{rl}\mathcal{H}^{1}(\mathcal{D}\cap\partial_{red}E_{n})\leq\mathcal{H}^{1}(\mathcal{D}^{+}(\sigma_{n}g)\cap\partial_{red}\{\psi_{n}>0\})+\mathcal{H}^{1}(\mathcal{D}^{0}(\sigma_{n}g)\cap\{\psi_{n}=0\}).\end{array} (3.53)

By using the similar estimate in Page 136-Page 137 in [2], one has

1(𝒟redEn)1(𝒟0(σng))+cσn2.\begin{array}[]{rl}\mathcal{H}^{1}(\mathcal{D}\cap\partial_{red}E_{n})\geq\mathcal{H}^{1}(\mathcal{D}^{0}(\sigma_{n}g))+c\sigma_{n}^{2}.\end{array} (3.54)

It follows from (3.52)-(3.54) that

1(𝒟0(σng))+cσn21+δn1δn1(𝒟0(σng))+Cρnσnλ0(1δn),\begin{array}[]{rl}&\mathcal{H}^{1}(\mathcal{D}^{0}(\sigma_{n}g))+c\sigma_{n}^{2}\leq\frac{1+\delta_{n}}{1-\delta_{n}}\mathcal{H}^{1}(\mathcal{D}^{0}(\sigma_{n}g))+\frac{C\rho_{n}\sigma_{n}}{\lambda_{0}(1-\delta_{n})},\end{array} (3.55)

that is

0<c2δnσn21δn1(𝒟0(σng))+Cρnσn1λ0(1δn),0<c\leq\frac{2\delta_{n}\sigma_{n}^{-2}}{1-\delta_{n}}\mathcal{H}^{1}(\mathcal{D}^{0}(\sigma_{n}g))+\frac{C\rho_{n}\sigma_{n}^{-1}}{\lambda_{0}(1-\delta_{n})},

which contradicts to the facts ρnσn10\displaystyle\rho_{n}\sigma_{n}^{-1}\rightarrow 0 and δnσn20\displaystyle\delta_{n}\sigma^{-2}_{n}\rightarrow 0.

With the aid of Lemma 3.10, based on Lemma 7.6 in [2] and Lemma 5.6 in [6], we will show that

Lemma 3.11.

There exists a constant C\displaystyle C, such that for any x(12,12)\displaystyle x\in\left(-\frac{1}{2},\frac{1}{2}\right),

0141r2(Avr(b(x))b(x))𝑑rC,\int_{0}^{\frac{1}{4}}\frac{1}{r^{2}}(\text{Av}_{r}(b(x))-b(x))dr\leq C,

where Avr(b(x))=b(x+r)+b(xr)2\displaystyle\text{Av}_{r}(b(x))=\frac{b(x+r)+b(x-r)}{2}.

Proof.

Set ψn(X)=ψ(ρnX)ρn\displaystyle\psi_{n}(X)=\frac{\psi(\rho_{n}X)}{\rho_{n}}, λn(X)=λ(ρnX)\displaystyle\lambda_{n}(X)=\lambda(\rho_{n}X) and λ0=λ(0)\displaystyle\lambda_{0}=\lambda(0), one has

Δψn+ρnf(ρnψn)=0in{ψn>0},and ψn(σn,σn;δn) in B1.\Delta\psi_{n}+\rho_{n}f(\rho_{n}\psi_{n})=0\ \ \text{in}\ \ \{\psi_{n}>0\},\ \ \text{and $\displaystyle\psi_{n}\in\mathcal{F}(\sigma_{n},\sigma_{n};\delta_{n})$ in $\displaystyle B_{1}$}. (3.56)

It is easy to check that ψn\displaystyle\psi_{n} be the class of (8σn,8σn;δn)\displaystyle\mathcal{F}(8\sigma_{n},8\sigma_{n};\delta_{n}) in B14(x,σnbn+(x))\displaystyle B_{\frac{1}{4}}(x,\sigma_{n}b_{n}^{+}(x)), provided that n\displaystyle n is sufficiently large. Therefore, it suffices to show the lemma for x=0\displaystyle x=0, that is

0141r2(Avr(b(0)))𝑑rC.\int_{0}^{\frac{1}{4}}\frac{1}{r^{2}}(\text{Av}_{r}(b(0)))dr\leq C. (3.57)

Set

ϕn(x,y)=ψn(x,y)+λ0yσn.\phi_{n}(x,y)=\frac{\psi_{n}(x,y)+\lambda_{0}y}{\sigma_{n}}.

Recalling the flatness condition for ψn\displaystyle\psi_{n}, the free boundary of ψn\displaystyle\psi_{n} lies in the strip |y|cσn\displaystyle|y|\leq c\sigma_{n}. Since |ψn|λ0(1+δn)\displaystyle|\nabla\psi_{n}|\leq\lambda_{0}(1+\delta_{n}) and δnσn\displaystyle\delta_{n}\leq\sigma_{n}, we have

ϕnC\displaystyle\phi_{n}\leq C in B1=B1{y<0}\displaystyle B_{1}^{-}=B_{1}\cap\{y<0\}.

The flatness condition implies that ϕnC\displaystyle\phi_{n}\geq-C in B1\displaystyle B_{1}^{-}. Therefore, one has

|ϕn|CinB1.|\phi_{n}|\leq C\ \ \text{in}\ \ B_{1}^{-}. (3.58)

It is easy to check that

Δϕn=ρnσnf(ρn(σnϕnλ0y))inB1{y<σn}.\Delta\phi_{n}=\frac{\rho_{n}}{\sigma_{n}}f(\rho_{n}(\sigma_{n}\phi_{n}-\lambda_{0}y))\ \ \text{in}\ \ B_{1}^{-}\cap\{y<-\sigma_{n}\}. (3.59)

Thanks to the flatness conditions on ψn\displaystyle\psi_{n} and Lemma 3.7, there exists a subsequence still labeled by {ψn}\displaystyle\{\psi_{n}\}, such that

ψn(x,y)λ0yin C2 in any compact subset of B1.\psi_{n}(x,y)\rightarrow-\lambda_{0}y\ \ \text{in $\displaystyle C^{2}$ in any compact subset of $\displaystyle B_{1}^{-}$}.

In view of (3.58), we can conclude that

ϕnϕin C2 in compact subsets of B1.\phi_{n}\rightarrow\phi\ \ \text{in $\displaystyle C^{2}$ in compact subsets of $\displaystyle B_{1}^{-}$}. (3.60)

Since ρnσn10\displaystyle\rho_{n}\sigma_{n}^{-1}\rightarrow 0, it follows from (3.59) and (3.60) that

Δϕ=0inB1,and|ϕ|C.\Delta\phi=0\ \ \text{in}\ \ B_{1}^{-},\ \ \text{and}\ \ |\phi|\leq C. (3.61)

By virtue of ϕn(0,0)=0\displaystyle\phi_{n}(0,0)=0, we have that for any y0\displaystyle y\leq 0, there exists a ξ(y,0)\displaystyle\xi\in(y,0), such that

ϕn(0,y)ϕn(0,ξ)yy=1σn(ψn(0,ξ)y+λ0)y|ψn|λ0σn|y|λ0δnσn|y|0.\phi_{n}(0,y)\leq\frac{\partial\phi_{n}(0,\xi)}{\partial y}y=\frac{1}{\sigma_{n}}\left(\frac{\partial\psi_{n}(0,\xi)}{\partial y}+\lambda_{0}\right)y\leq\frac{|\nabla\psi_{n}|-\lambda_{0}}{\sigma_{n}}|y|\leq\frac{\lambda_{0}\delta_{n}}{\sigma_{n}}|y|\rightarrow 0. (3.62)

Then it gives that

ϕ(0,y)0for any y0.\phi(0,y)\leq 0\ \ \text{for any $\displaystyle y\leq 0$}. (3.63)

Next, we will show that

ϕ(x,0)=λ0b(x)in the sense that limy0ϕ(x,y)=λ0b(x).\phi(x,0)=\lambda_{0}b(x)\ \ \text{in the sense that $\displaystyle\lim_{y\uparrow 0}\phi(x,y)=\lambda_{0}b(x)$}. (3.64)

To obtain the fact (3.64), we first show that for any small η>0\displaystyle\eta>0 and any large constant M\displaystyle M, we have

ϕn(x,yσn)λ0b(x)uniformly for xIη=(1+η,1η) and My1,\phi_{n}(x,y\sigma_{n})\rightarrow\lambda_{0}b(x)\ \ \text{uniformly for $\displaystyle x\in I_{\eta}=(-1+\eta,1-\eta)$ and $\displaystyle-M\leq y\leq-1$}, (3.65)

as n\displaystyle n\rightarrow\infty.

On another hand, to see this, thanks to the non-homogeneous blow-up Lemma 3.9, it suffices to show that

ϕn(x,hσn)λ0bn+(x)0.\phi_{n}(x,h\sigma_{n})-\lambda_{0}b_{n}^{+}(x)\rightarrow 0. (3.66)

It follows from (3.62) that for any xIη\displaystyle x\in I_{\eta}, one has

ϕn(x,yσn)λ0bn+(x)=ϕn(x,σnbn+(x))λ0bn+(x)+ϕn(0,ξ)yσn(ybn+(x))δn(bn+(x)y)(1+M)δn0,\begin{array}[]{rl}\phi_{n}(x,y\sigma_{n})-\lambda_{0}b_{n}^{+}(x)=&\phi_{n}(x,\sigma_{n}b_{n}^{+}(x))-\lambda_{0}b_{n}^{+}(x)+\frac{\partial\phi_{n}(0,\xi)}{\partial y}\sigma_{n}(y-b_{n}^{+}(x))\\ \leq&\delta_{n}(b_{n}^{+}(x)-y)\leq(1+M)\delta_{n}\rightarrow 0,\end{array} (3.67)

as n\displaystyle n\rightarrow\infty, where we have used the fact

ϕn(x,σnbn+(x))λ0bn+(x)=ψn(x,σnbn+(x))+λ0σnbn+(x)σnλ0bn+(x)=ψn(x,σnbn+(x))σn=0.\phi_{n}(x,\sigma_{n}b_{n}^{+}(x))-\lambda_{0}b_{n}^{+}(x)=\frac{\psi_{n}(x,\sigma_{n}b_{n}^{+}(x))+\lambda_{0}\sigma_{n}b_{n}^{+}(x)}{\sigma_{n}}-\lambda_{0}b_{n}^{+}(x)=\frac{\psi_{n}(x,\sigma_{n}b_{n}^{+}(x))}{\sigma_{n}}=0.

Taking any sequence {xn}\displaystyle\{x_{n}\} with xnIη\displaystyle x_{n}\in I_{\eta} and Mhn1\displaystyle-M\leq h_{n}\leq-1, and consider ψn\displaystyle\psi_{n} in BRσn(Xn)\displaystyle B_{R\sigma_{n}}(X_{n}), where Xn=(xn,σnbn+(xn))\displaystyle X_{n}=(x_{n},\sigma_{n}b_{n}^{+}(x_{n})) is the free boundary point and R\displaystyle R is an arbitrary large constant. Denote σ~n=1RsupxIRσn(bn+(x)bn+(xn))\displaystyle\tilde{\sigma}_{n}=\frac{1}{R}\sup_{x\in I_{R\sigma_{n}}}(b_{n}^{+}(x)-b_{n}^{+}(x_{n})) with IRσn=(xnRσn,xn+Rσn)\displaystyle I_{R\sigma_{n}}=(x_{n}-R\sigma_{n},x_{n}+R\sigma_{n}). It follows from Lemma 3.9 that σ~n0\displaystyle\tilde{\sigma}_{n}\rightarrow 0. Next, we claim that

ψn(σ~n,1;C1δn)inBRσn(Xn),\psi_{n}\in\mathcal{F}(\tilde{\sigma}_{n},1;C_{1}\delta_{n})\ \ \text{in}\ \ B_{R\sigma_{n}}(X_{n}), (3.68)

where C1\displaystyle C_{1} is a constant depending on λ1\displaystyle\lambda_{1} and λ2\displaystyle\lambda_{2}, provided that n\displaystyle n is sufficiently large. In fact, the definition of bn+(x)\displaystyle b^{+}_{n}(x) gives that

 ψn(x,y)=0 for yσnbn+(xn)+σ~nRσn.\text{ $\displaystyle\psi_{n}(x,y)=0\ $ for $\displaystyle\ y\geq\sigma_{n}b_{n}^{+}(x_{n})+\tilde{\sigma}_{n}R\sigma_{n}$}.

Since |ψn(X)|λ0(1+δn)\displaystyle|\nabla\psi_{n}(X)|\leq\lambda_{0}(1+\delta_{n}) in B1(0)\displaystyle B_{1}(0) and supX,YB1(0)|λn(X)λn(Y)|λ0δn,\displaystyle\sup_{X,Y\in B_{1}(0)}|\lambda_{n}(X)-\lambda_{n}(Y)|\leq\lambda_{0}\delta_{n}, we have that

|ψn(X)|(λn(Xn)+λ0δn)(1+δn)λn(Xn)(1+C1δn)in BRσn(Xn),|\nabla\psi_{n}(X)|\leq(\lambda_{n}(X_{n})+\lambda_{0}\delta_{n})(1+\delta_{n})\leq\lambda_{n}(X_{n})\left(1+C_{1}\delta_{n}\right)\ \ \text{in $\displaystyle B_{R\sigma_{n}}(X_{n})$,}

and

supX,YBRσn(Xn)|λn(X)λn(Y)|λ0δn(λn(Xn)+λ0δn)δnλn(Xn)C1δn,\sup_{X,Y\in B_{R\sigma_{n}}(X_{n})}|\lambda_{n}(X)-\lambda_{n}(Y)|\leq\lambda_{0}\delta_{n}\leq(\lambda_{n}(X_{n})+\lambda_{0}\delta_{n})\delta_{n}\leq\lambda_{n}(X_{n})C_{1}\delta_{n},

where C1\displaystyle C_{1} is a constant depending on λ1\displaystyle\lambda_{1} and λ2\displaystyle\lambda_{2}. Therefore, we obtain the claim (3.68).

In view of the claim (3.68), by virtue of the flatness Lemma 3.8, we have

ψn(2σ¯n,Cσ¯n;C1δn)inBRσn2(Xn),\psi_{n}\in\mathcal{F}(2\bar{\sigma}_{n},C\bar{\sigma}_{n};C_{1}\delta_{n})\ \ \text{in}\ \ B_{\frac{R\sigma_{n}}{2}}(X_{n}), (3.69)

provided that σ¯n=max{σ~n,δn}\displaystyle\bar{\sigma}_{n}=\max\{\tilde{\sigma}_{n},\delta_{n}\}.

Hence, for any h\displaystyle h with |h|<R2\displaystyle|h|<\frac{R}{2}, it follows from (3.69) that

ψn(Xn+hσne2)λn(Xn)(hσn+Cσ¯nRσn2),\psi_{n}(X_{n}+h\sigma_{n}e_{2})\geq-\lambda_{n}(X_{n})\left(h\sigma_{n}+\frac{C\bar{\sigma}_{n}R\sigma_{n}}{2}\right),

that is

ϕn(Xn+hσne2)λ0bn+(xn)=ψn(Xn+hσne2)+λ0hσnσn(λ(0)λ(ρnXn))hCλ(ρnXn)σ¯nR20.\begin{array}[]{rl}\phi_{n}(X_{n}+h\sigma_{n}e_{2})-\lambda_{0}b_{n}^{+}(x_{n})=&\frac{\psi_{n}(X_{n}+h\sigma_{n}e_{2})+\lambda_{0}h\sigma_{n}}{\sigma_{n}}\\ \geq&(\lambda(0)-\lambda(\rho_{n}X_{n}))h-\frac{C\lambda(\rho_{n}X_{n})\bar{\sigma}_{n}R}{2}\rightarrow 0.\end{array}

This together with (3.67) gives (3.66).

For any ε>0\displaystyle\varepsilon>0, let φε\displaystyle\varphi_{\varepsilon} be the solution of the Dirichlet problem

{Δφε=1inB1η,φε=λ0b(x)εonB1η{y=0},φε=infB1ϕonB1η{y<0}.\left\{\begin{array}[]{ll}&\Delta\varphi_{\varepsilon}=1~~~~\text{in}~~B_{1-\eta}^{-},\\ &\varphi_{\varepsilon}=\lambda_{0}b(x)-\varepsilon~~~~\text{on}~~\partial B_{1-\eta}^{-}\cap\{y=0\},\ \ \ \ \ \varphi_{\varepsilon}=\inf_{B_{1}^{-}}\phi~~~~\text{on}~~\partial B_{1-\eta}^{-}\cap\{y<0\}.\end{array}\right.

It follows from (3.65) that

ϕn>φεon(B1η{y<Mσn}),\phi_{n}>\varphi_{\varepsilon}\ \ \text{on}\ \ \partial(B_{1-\eta}^{-}\cap\{y<-M\sigma_{n}\}), (3.70)

for any large constant M\displaystyle M (independent of η\displaystyle\eta and ε\displaystyle\varepsilon), provided that n\displaystyle n is sufficiently large (depending on ε\displaystyle\varepsilon and M\displaystyle M). Moreover,

Δ(φεϕn)=1+ρnσnf(ρn(σnϕnλ0y))>0 in B1η{y<Mσn},\Delta(\varphi_{\varepsilon}-\phi_{n})=1+\frac{\rho_{n}}{\sigma_{n}}f(\rho_{n}(\sigma_{n}\phi_{n}-\lambda_{0}y))>0\ \text{ in $\displaystyle B_{1-\eta}^{-}\cap\{y<-M\sigma_{n}\}$},

provided that n\displaystyle n is sufficiently large. In view of (3.70), the maximum principe gives that

φεϕn in B1η{y<Mσn}, if n is large enough.\varphi_{\varepsilon}\leq\phi_{n}\ \ \text{ in $\displaystyle B_{1-\eta}^{-}\cap\{y<-M\sigma_{n}\}$, if $\displaystyle n$ is large enough}. (3.71)

Taking n\displaystyle n\rightarrow\infty in (3.71), we have

φε(x,y)ϕ(x,y) in B1η.\varphi_{\varepsilon}(x,y)\leq\phi(x,y)\ \ \text{ in $\displaystyle B_{1-\eta}^{-}$}.

Consequently,

λ0b(x)2εlimy0φε(x,y)lim infy0ϕ(x,y) for x(1+2η,12η) and η>0.\lambda_{0}b(x)-2\varepsilon\leq\lim_{y\uparrow 0}\varphi_{\varepsilon}(x,y)\leq\liminf_{y\uparrow 0}\phi(x,y)\ \ \text{ for $\displaystyle x\in(-1+2\eta,1-2\eta)$ and $\displaystyle\eta>0$}. (3.72)

Similarly, by working with a solution φ~ε(x,y)\displaystyle\tilde{\varphi}_{\varepsilon}(x,y) to the Dirichlet problem

{Δφ~ε=1inB1η,φ~ε=λ0b(x)+εonB1η{y=0},φ~ε=supB1ϕonB1η{y<0},\left\{\begin{array}[]{ll}&\Delta\tilde{\varphi}_{\varepsilon}=-1~~~~\text{in}~~B_{1-\eta}^{-},\\ &\tilde{\varphi}_{\varepsilon}=\lambda_{0}b(x)+\varepsilon~~~~\text{on}~~\partial B_{1-\eta}^{-}\cap\{y=0\},\ \ \ \ \ \tilde{\varphi}_{\varepsilon}=\sup_{B_{1}^{-}}\phi~~~~\text{on}~~\partial B_{1-\eta}^{-}\cap\{y<0\},\end{array}\right.

for any ε>0\displaystyle\varepsilon>0, we can obtain that

λ0b(x)+2εlim supy0ϕ(x,y) for x(1+2η,12η).\lambda_{0}b(x)+2\varepsilon\geq\limsup_{y\uparrow 0}\phi(x,y)\ \ \text{ for $\displaystyle x\in(-1+2\eta,1-2\eta)$}. (3.73)

By virtue of the arbitrariness of ε\displaystyle\varepsilon, it follows from (3.72) and (3.73) that we obtain the claim (3.64).

In view of Lemma 3.10, (3.58), (3.61),(3.63) and (3.64), we can apply Lemma 5.5 in [6] to obtain (3.57). ∎

With the aid of Lemma 3.11, it follows from Lemma 7.7 - Lemma 7.9 in [2] that the flatness from below implies the flatness from above (see Figure 4).

Refer to caption
Figure 4. Flatness from above
Lemma 3.12.

For any θ>0\displaystyle\theta>0, there exist a large constant C\displaystyle C, cθ=c(θ)\displaystyle c_{\theta}=c(\theta) and a small σθ=σ(θ)\displaystyle\sigma_{\theta}=\sigma(\theta), such that if

ψ(σ,σ;δ)in Bρ in direction ν with σσθ,δσθσ2,ρσθσ,\psi\in\mathcal{F}(\sigma,\sigma;\delta)\ \ \text{in $\displaystyle B_{\rho}$ in direction $\displaystyle\nu$ with $\displaystyle\sigma\leq\sigma_{\theta},\delta\leq\sigma_{\theta}\sigma^{2},\rho\leq\sigma_{\theta}\sigma$},

then

ψ(θσ,1;δ)in Bρ¯ in direction ν¯ for some ρ¯ and ν¯ with cθρρ¯θρ and |νν¯|Cσ.\psi\in\mathcal{F}(\theta\sigma,1;\delta)\ \ \text{in $\displaystyle B_{\bar{\rho}}$ in direction $\displaystyle\bar{\nu}$ for some $\displaystyle\bar{\rho}$ and $\displaystyle\bar{\nu}$ with $\displaystyle c_{\theta}\rho\leq\bar{\rho}\leq\theta\rho$ and $\displaystyle|\nu-\bar{\nu}|\leq C\sigma$}.

By virtue of Lemma 3.8 and Lemma 3.12, we will show that

Lemma 3.13.

Let ω(s)\displaystyle\omega(s) be a monotone increasing and continuous function with ω(0)=0\displaystyle\omega(0)=0. For any θ(0,1)\displaystyle\theta\in(0,1), there exists a small σ0>0\displaystyle\sigma_{0}>0, such that if

ψ(σ,1;δ)in Bρ in direction ν with σσ0,δσ0σ2,ρσ0σ,\psi\in\mathcal{F}(\sigma,1;\delta)\ \ \text{in $\displaystyle B_{\rho}$ in direction $\displaystyle\nu$ with $\displaystyle\sigma\leq\sigma_{0},\ \delta\leq\sigma_{0}\sigma^{2},\ \rho\leq\sigma_{0}\sigma$},

and supX,YBsρ|λ(X)λ(Y)|ω(s)λ(0)δ\displaystyle\sup_{X,Y\in B_{s\rho}}|\lambda(X)-\lambda(Y)|\leq\omega(s)\lambda(0)\delta for s[0,1]\displaystyle s\in[0,1], then there exist a large constant C\displaystyle C and cθ=c(θ,ω)\displaystyle c_{\theta}=c(\theta,\omega), such that

ψ(θσ,θσ;θ2δ)in Bρ¯ in direction ν¯\psi\in\mathcal{F}(\theta\sigma,\theta\sigma;\theta^{2}\delta)\ \ \text{in $\displaystyle B_{\bar{\rho}}$ in direction $\displaystyle\bar{\nu}$}

for some ρ¯\displaystyle\bar{\rho} and ν¯\displaystyle\bar{\nu}, where cθρρ¯14ρ\displaystyle c_{\theta}\rho\leq\bar{\rho}\leq\frac{1}{4}\rho and |νν¯|Cσ\displaystyle|\nu-\bar{\nu}|\leq C\sigma.

Proof.

Step 1. By virtue of Lemma 3.8, one has

ψ(C0σ,C0σ;δ)in Bρ2 in direction ν, where C0>2.\psi\in\mathcal{F}(C_{0}\sigma,C_{0}\sigma;\delta)\ \ \text{in $\displaystyle B_{\frac{\rho}{2}}$ in direction $\displaystyle\nu$, where $\displaystyle C_{0}>2$.}

Then for some small θ1(0,1/2]\displaystyle\theta_{1}\in\left(0,1/2\right] to be determined later, it follows from Lemma 3.12 that

ψ(C0θ1σ,1;δ)in Bρ1 in direction ν1 for some ρ1 and ν1,\psi\in\mathcal{F}(C_{0}\theta_{1}\sigma,1;\delta)\ \ \text{in $\displaystyle B_{\rho_{1}}$ in direction $\displaystyle\nu_{1}$ for some $\displaystyle\rho_{1}$ and $\displaystyle\nu_{1}$},

where

cθ1ρ2ρ1θ1ρ2 and |νν1|Cσ.\text{$\displaystyle c_{\theta_{1}}\frac{\rho}{2}\leq\rho_{1}\leq\theta_{1}\frac{\rho}{2}$ and $\displaystyle|\nu-\nu_{1}|\leq C\sigma$}.

To improve the value of δ\displaystyle\delta, we define

U(X)=max{|ψ(X)|supXB3ρ1λ(X),0}inB2ρ1{ψ>0}.U(X)=\max\left\{|\nabla\psi(X)|-\sup_{X\in B_{3\rho_{1}}}\lambda(X),0\right\}\ \ \text{in}\ \ B_{2\rho_{1}}\cap\{\psi>0\}.

It is easy to check that

ΔU=|D2ψ||ψ|2kψjψijψikψ|ψ|3f(ψ)|ψ|0inB2ρ1{ψ>0}.\Delta U=\frac{|D^{2}\psi||\nabla\psi|^{2}-\partial_{k}\psi\partial_{j}\psi\partial_{ij}\psi\partial_{ik}\psi}{|\nabla\psi|^{3}}-f^{\prime}(\psi)|\nabla\psi|\geq 0\ \ \text{in}\ \ B_{2\rho_{1}}\cap\{\psi>0\}.

For subharmonic function U\displaystyle U, we can use the similar arguments in P141 in [2] and obtain that there exists a c(0,1)\displaystyle c\in(0,1), such that

U(X)=max{|ψ(X)|supXB3ρ1λ(X),0}(1c)λ(0)δinBρ1,U(X)=\max\left\{|\nabla\psi(X)|-\sup_{X\in B_{3\rho_{1}}}\lambda(X),0\right\}\leq(1-c)\lambda(0)\delta\ \ \text{in}\ \ B_{\rho_{1}},

which implies that

|ψ(X)|supXB3ρ1λ(X)+λ(0)(1c)δλ(0)+λ(0)(1c+w(2θ1))δinBρ1.|\nabla\psi(X)|\leq\sup_{X\in B_{3\rho_{1}}}\lambda(X)+\lambda(0)(1-c)\delta\leq\lambda(0)+\lambda(0)(1-c+w(2\theta_{1}))\delta\ \ \text{in}\ \ B_{\rho_{1}}.

Denote

θ0=1c2.\theta_{0}=\sqrt{1-\frac{c}{2}}.

Taking θ1\displaystyle\theta_{1} be sufficiently small, such that θ12θ0<1\displaystyle\frac{\theta_{1}}{2\theta_{0}}<1 and C0θ1<θ0\displaystyle C_{0}\theta_{1}<\theta_{0}, then the continuity of ω(s)\displaystyle\omega(s) gives that

ω(2θ1)c2andω(θ12)<θ02.\omega(2\theta_{1})\leq\frac{c}{2}\ \ \text{and}\ \ \omega\left(\frac{\theta_{1}}{2}\right)<\theta_{0}^{2}.

Furthermore,

|ψ(X)|λ(0)+λ(0)(1c2)δλ(0)(1+θ02δ)inBρ1,|\nabla\psi(X)|\leq\lambda(0)+\lambda(0)\left(1-\frac{c}{2}\right)\delta\leq\lambda(0)(1+\theta_{0}^{2}\delta)\ \ \text{in}\ \ B_{\rho_{1}},

and

supX,YBρ1|λ(X)λ(Y)|supX,YBθ12ρ|λ(X)λ(Y)|ω(θ12)λ(0)δθ02λ(0)δ.\sup_{X,Y\in B_{\rho_{1}}}|\lambda(X)-\lambda(Y)|\leq\sup_{X,Y\in B_{\frac{\theta_{1}}{2}\rho}}|\lambda(X)-\lambda(Y)|\leq\omega\left(\frac{\theta_{1}}{2}\right)\lambda(0)\delta\leq\theta_{0}^{2}\lambda(0)\delta.

Thus we have

ψ(θ0σ,1;θ02δ)in Bρ1 in direction ν1.\psi\in\mathcal{F}(\theta_{0}\sigma,1;\theta_{0}^{2}\delta)\ \ \text{in $\displaystyle B_{\rho_{1}}$ in direction $\displaystyle\nu_{1}$}.

On the other hand, we have

supX,YBsρ1|λ(X)λ(Y)|ω1(s)θ02λ(0)δandω1(s)=θ02ω(θ12s)θ02ω(θ12)<1.\sup_{X,Y\in B_{s\rho_{1}}}|\lambda(X)-\lambda(Y)|\leq\omega_{1}(s)\theta_{0}^{2}\lambda(0)\delta\ \ \text{and}\ \ \omega_{1}(s)=\theta_{0}^{-2}\omega\left(\frac{\theta_{1}}{2}s\right)\leq\theta_{0}^{-2}\omega\left(\frac{\theta_{1}}{2}\right)<1.

Notice that

ρ1θ1ρ2(θ12θ0σ0)θ0σ<σ0(θ0σ)andθ02δσ0(θ0σ)2,\rho_{1}\leq\theta_{1}\frac{\rho}{2}\leq\left(\frac{\theta_{1}}{2\theta_{0}}\sigma_{0}\right)\theta_{0}\sigma<\sigma_{0}(\theta_{0}\sigma)\ \ \text{and}\ \ \theta^{2}_{0}\delta\leq\sigma_{0}(\theta_{0}\sigma)^{2},

due to θ12θ0<1\displaystyle\frac{\theta_{1}}{2\theta_{0}}<1 and ρσ0σ\displaystyle\rho\leq\sigma_{0}\sigma.

Step 2. Repeating the arguments in Step 1, and choosing sufficiently small σ0>0\displaystyle\sigma_{0}>0, we obtain

ψ(θ02σ,1;θ04δ)in Bρ2 in direction ν2,\psi\in\mathcal{F}(\theta^{2}_{0}\sigma,1;\theta_{0}^{4}\delta)\ \ \text{in $\displaystyle B_{\rho_{2}}$ in direction $\displaystyle\nu_{2}$},

for some ρ2\displaystyle\rho_{2} and ν2\displaystyle\nu_{2}, where

cθ1ρ12ρ2θ1ρ12 and |ν2ν1|Cθ0σ.\text{$\displaystyle c_{\theta_{1}}\frac{\rho_{1}}{2}\leq\rho_{2}\leq\theta_{1}\frac{\rho_{1}}{2}$ and $\displaystyle|\nu_{2}-\nu_{1}|\leq C\theta_{0}\sigma$}.

Similarly, we can repeat the arguments in Step 1 for a finite number n\displaystyle n, and we choose the constant σ0\displaystyle\sigma_{0} being small enough in the statement for each step, thus

ψ(θ0nσ,1;θ02nδ)in Bρn in direction νn,\psi\in\mathcal{F}(\theta^{n}_{0}\sigma,1;\theta_{0}^{2n}\delta)\ \ \text{in $\displaystyle B_{\rho_{n}}$ in direction $\displaystyle\nu_{n}$}, (3.74)

for some ρn\displaystyle\rho_{n} and νn\displaystyle\nu_{n}, where

cθ1ρn12ρnθ1ρn12 and |νnνn1|Cθ0n1σ.\text{$\displaystyle c_{\theta_{1}}\frac{\rho_{n-1}}{2}\leq\rho_{n}\leq\theta_{1}\frac{\rho_{n-1}}{2}$ and $\displaystyle|\nu_{n}-\nu_{n-1}|\leq C\theta^{n-1}_{0}\sigma$}. (3.75)

Applying Lemma 3.8 again, we have

ψ(2θ0nσ,Cθ0nσ;θ02nδ)in Bρn2 in direction νn.\psi\in\mathcal{F}(2\theta_{0}^{n}\sigma,C\theta_{0}^{n}\sigma;\theta_{0}^{2n}\delta)\ \ \text{in $\displaystyle B_{\frac{\rho_{n}}{2}}$ in direction $\displaystyle\nu_{n}$.}

Choosing N=N(θ)\displaystyle N=N(\theta) as the smallest positive integer such that

max{2θ0N,Cθ0N}θ,\max\left\{2\theta_{0}^{N},C\theta_{0}^{N}\right\}\leq\theta,

where the constant C\displaystyle C is obtained in Lemma 3.8. It follows from (3.75) that

cθ1N2NρρNθ1N2Nρ and |νNν|Cσ1θ0.\text{$\displaystyle\frac{c^{N}_{\theta_{1}}}{2^{N}}\rho\leq\rho_{N}\leq\frac{\theta_{1}^{N}}{2^{N}}\rho$ and $\displaystyle|\nu_{N}-\nu|\leq\frac{C\sigma}{1-\theta_{0}}$}. (3.76)

Then we have

ψ(θσ,θσ;θ2δ)in Bρ¯ in direction ν¯,\psi\in\mathcal{F}(\theta\sigma,\theta\sigma;\theta^{2}\delta)\ \ \text{in $\displaystyle B_{\bar{\rho}}$ in direction $\displaystyle\bar{\nu}$,}

where ρ¯=ρN2\displaystyle\bar{\rho}=\frac{\rho_{N}}{2} and ν¯=νN\displaystyle\bar{\nu}=\nu_{N}. Denote cθ=cθ1N2N(θ)+1\displaystyle c_{\theta}=\frac{c^{N}_{\theta_{1}}}{2^{N(\theta)+1}}, it follows from (3.76) that

cθρρ¯θ1N2N+1ρρ4 and |νν¯|Cσ1θ0Cσ.\text{$\displaystyle c_{\theta}\rho\leq\bar{\rho}\leq\frac{\theta_{1}^{N}}{2^{N+1}}\rho\leq\frac{\rho}{4}$ and $\displaystyle|\nu-\bar{\nu}|\leq\frac{C\sigma}{1-\theta_{0}}\leq C\sigma$}.

3.5. The regularity of the free boundary

In Lemma 3.13, we show that the flatness in Bρ\displaystyle B_{\rho} is improved in a smaller disc Bρ¯\displaystyle B_{\bar{\rho}}. Based on the results in the previous subsection, we will investigate the regularity of the free boundary in this subsection.

By virtue of Theorem 8.1 in [2], the flatness of free boundary implies immediately the C1,α\displaystyle C^{1,\alpha}-smoothness of the free boundary, which is the main result in this paper. We omit the proof and present the result as follows.

Theorem 3.14.

For any compact subset D\displaystyle D of Ω\displaystyle\Omega, there exists β>0,σ0>0,γ0>0\displaystyle\beta>0,\ \sigma_{0}>0,\ \gamma_{0}>0 and C>0\displaystyle C>0, such that if

ψ(σ,1;)in Bρ(X0) in direction ν\psi\in\mathcal{F}(\sigma,1;\infty)\ \ \text{in $\displaystyle B_{\rho}(X_{0})$ in direction $\displaystyle\nu$} (3.77)

where X0{ψ>0}\displaystyle X_{0}\in\partial\{\psi>0\}, Bρ(X0)D\displaystyle B_{\rho}(X_{0})\subset D and

 σσ0andργ0σ2β,\text{ $\displaystyle\sigma\leq\sigma_{0}\ \ \text{and}\ \ \rho\leq\gamma_{0}\sigma^{\frac{2}{\beta}}$},

then

Bρ4(X0){ψ>0}is a C1,α curvefor someα>0.B_{\frac{\rho}{4}}(X_{0})\cap\partial\{\psi>0\}\ \ \text{is a $\displaystyle C^{1,\alpha}$ curve}\ \ \text{for some}\ \alpha>0.

Namely, a graph in direction ν\displaystyle\nu of a C1,α\displaystyle C^{1,\alpha} function, for any X1\displaystyle X_{1} and X2\displaystyle X_{2} on this curve,

|ν(X1)ν(X2)|Cσ|X1X2ρ|α.|\nu(X_{1})-\nu(X_{2})|\leq C\sigma\left|\frac{X_{1}-X_{2}}{\rho}\right|^{\alpha}.

In particular, if ν=e2\displaystyle\nu=e_{2}, then Bρ4(X0){ψ>0}\displaystyle B_{\frac{\rho}{4}}(X_{0})\cap\partial\{\psi>0\} can be written in the form y=g(x)C1,α\displaystyle y=g(x)\in C^{1,\alpha}; Furthermore,

|g(x1)g(x2)|Cσ|X1X2ρ|α.|g^{\prime}(x_{1})-g^{\prime}(x_{2})|\leq C\sigma\left|\frac{X_{1}-X_{2}}{\rho}\right|^{\alpha}.

Moreover, the free boundary possesses higher regularity, provided that the functions λ(X)\displaystyle\lambda(X) and the vorticity strength function f(t)\displaystyle f(t) possess higher regularity. With the aid of Lemma 3.7 and Theorem 3.14, by using the similar arguments in Theorem 3.12 and Corollary 3.13 in [24], we have

Theorem 3.15.

(1) If λ(X)\displaystyle\lambda(X) is Ck,β\displaystyle C^{k,\beta} and f(t)\displaystyle f(t) is C1,β\displaystyle C^{1,\beta}, for k=0,1,2\displaystyle k=0,1,2, then the free boundary Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\} is Ck+1,α\displaystyle C^{k+1,\alpha} locally in Ω\displaystyle\Omega with α(0,β)\displaystyle\alpha\in(0,\beta).
(2) If λ(X)\displaystyle\lambda(X) is Ck,β\displaystyle C^{k,\beta} and f(t)\displaystyle f(t) is Ck1,β\displaystyle C^{k-1,\beta}, for k=3,4,\displaystyle k=3,4,\mathellipsis, then the free boundary Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\} is Ck+1,α\displaystyle C^{k+1,\alpha} locally in Ω\displaystyle\Omega with α(0,β)\displaystyle\alpha\in(0,\beta).
(3) If λ(X)\displaystyle\lambda(X) and f(t)\displaystyle f(t) are analytic, then the free boundary Ω{ψ>0}\displaystyle\Omega\cap\partial\{\psi>0\} is locally analytic in Ω\displaystyle\Omega.

Proof.

For any X0Ω{ψ>0}\displaystyle X_{0}\in\Omega\cap\partial\{\psi>0\}, let ψn(X)=ψ(X0+ρnX)ρn\displaystyle\psi_{n}(X)=\frac{\psi(X_{0}+\rho_{n}X)}{\rho_{n}} be a blow up sequence with respect to the discs Bρn(X0)\displaystyle B_{\rho_{n}}(X_{0}), it follows from Lemma 3.7 that there exists a unit vector ν0\displaystyle\nu_{0}, such that

ψn(X)ψ0(X)=λ(X0)max{Xν0,0}in any compact subset of 2.\psi_{n}(X)\rightarrow\psi_{0}(X)=\lambda(X_{0})\max\{-X\cdot\nu_{0},0\}\ \ \text{in any compact subset of $\displaystyle\mathbb{R}^{2}$}.

With the aid of Lemma 3.4, we have

ψ(X0+ρnX)=λ(X0)ρnmax{Xν0,0}+o(ρn)inB1(0),\psi(X_{0}+\rho_{n}X)=\lambda(X_{0})\rho_{n}\max\{-X\cdot\nu_{0},0\}+o(\rho_{n})\ \ \text{in}\ \ B_{1}(0),

which implies that there exists a sequence {σn}\displaystyle\{\sigma_{n}\} with σn0\displaystyle\sigma_{n}\rightarrow 0, such that

ψF(σn,1,)in Bρn(X0) in direction ν0σnσ0 and ρnγ0σn2β,\psi\in F(\sigma_{n},1,\infty)\ \ \text{in $\displaystyle B_{\rho_{n}}(X_{0})$ in direction $\displaystyle\nu_{0}$, $\displaystyle\sigma_{n}\leq\sigma_{0}$ and $\displaystyle\rho_{n}\leq\gamma_{0}\sigma_{n}^{\frac{2}{\beta}}$},

provided that n\displaystyle n is sufficiently large. Applying Theorem 3.14, we have that Bρn4(X0){ψ>0}\displaystyle B_{\frac{\rho_{n}}{4}}(X_{0})\cap\partial\{\psi>0\} is C1,α\displaystyle C^{1,\alpha}.

Next, we can take a C1,α\displaystyle C^{1,\alpha} transformation to flatten the free boundary. Then reflect ψ\displaystyle\psi to the full neighborhood of the free boundary, applying the Schauder estimates for the elliptic equations in divergence form in Section 9 in [1], we can obtain the C1,α\displaystyle C^{1,\alpha} regularity of ψ\displaystyle\psi up to the free boundary.

It follows from Proposition 2.2 that

limε0+Ω{ψ>ε}(|ψ|2λ2(X)F(ψ))ηνε𝑑S=0.\lim_{\varepsilon\rightarrow 0^{+}}\int_{\Omega\cap\partial\{\psi>\varepsilon\}}(|\nabla\psi|^{2}-\lambda^{2}(X)-F(\psi))\eta\cdot\nu_{\varepsilon}dS=0.

Since ψ\displaystyle\psi is C1,α\displaystyle C^{1,\alpha} up to the free boundary and F(ψ)=0\displaystyle F(\psi)=0 on the free boundary, we have

|ψ(X)|=λ(X)on the free boundary.|\nabla\psi(X)|=\lambda(X)\ \ \text{on the free boundary}.

Without loss of generality, we assume that the outward normal direction to {ψ>0}\displaystyle\partial\{\psi>0\} is in the direction of the positive y\displaystyle y-axis. Extend ψ\displaystyle\psi and f\displaystyle f as C1,α\displaystyle C^{1,\alpha} functions into a full neighborhood of 0{ψ>0}\displaystyle 0\in\partial\{\psi>0\}. In view of |ψ(0)|=λ(0)λ1>0\displaystyle|\nabla\psi(0)|=\lambda(0)\geq\lambda_{1}>0, we have that

ψy(0)<0.\psi_{y}(0)<0. (3.78)

Define a mapping as follows,

S=TX=(s,t)(x,ψ(x,y)),X=(x,y).S=TX=(s,t)\triangleq(x,\psi(x,y)),\ \ X=(x,y).

In view of (3.78), it is easy to check that

det(SX)=ψy(x,y)<0in a neighborhood of 0.\text{det}\left(\frac{\partial S}{\partial X}\right)=\psi_{y}(x,y)<0\ \ \text{in a neighborhood of $\displaystyle 0$}.

And thus the mapping T\displaystyle T is a local diffeomorphism near 0\displaystyle 0.

Construct a function ϕ(S)\displaystyle\phi(S) as follows,

ϕ(S)=y.\phi(S)=y.

Therefore, the free boundary Γ\displaystyle\Gamma is transformed into t=0\displaystyle t=0, and we have

(XS)=(SX)1=(10ψxψy1ψy).\left(\frac{\partial X}{\partial S}\right)=\left(\frac{\partial S}{\partial X}\right)^{-1}=\left(\begin{matrix}1&0\\ -\frac{\psi_{x}}{\psi_{y}}&\frac{1}{\psi_{y}}\end{matrix}\right).

Consequently, one has

ϕs=ys=ψxψy,ϕt=yt=1ψy,tx=ψx=ϕsϕtandty=ψy=1ϕt.\phi_{s}=\frac{\partial y}{\partial s}=-\frac{\psi_{x}}{\psi_{y}},\ \ \phi_{t}=\frac{\partial y}{\partial t}=\frac{1}{\psi_{y}},\ \ \frac{\partial t}{\partial x}=\psi_{x}=-\frac{\phi_{s}}{\phi_{t}}\ \ \text{and}\ \ \frac{\partial t}{\partial y}=\psi_{y}=\frac{1}{\phi_{t}}. (3.79)

It follows from (3.79) that

𝒬ϕ=s(ϕsϕt)+t(1+ϕs22ϕt2)+f(t)=0in the neighborhood of 0.\mathcal{Q}\phi=\partial_{s}\left(-\frac{\phi_{s}}{\phi_{t}}\right)+\partial_{t}\left(\frac{1+\phi^{2}_{s}}{2\phi^{2}_{t}}\right)+f(t)=0\ \ \text{in the neighborhood of $\displaystyle 0$}.

It is easy to check that 𝒬ϕ=0\displaystyle\mathcal{Q}\phi=0 is a quasilinear elliptic equation in a neighborhood E\displaystyle E of 0\displaystyle 0. Furthermore, ϕ\displaystyle\phi satisfies the Neumann type boundary condition as follows,

{𝒬ϕ=0inD,ϕt1+ϕs2=1λ(s)onD¯{t=0},\left\{\begin{array}[]{ll}\mathcal{Q}\phi=0&\text{in}\ \ \ D,\\ \frac{\phi_{t}}{\sqrt{1+\phi_{s}^{2}}}=\frac{1}{\lambda(s)}&\text{on}\ \ \bar{D}\cap\{t=0\},\end{array}\right. (3.80)

where D=E{t>0}\displaystyle D=E\cap\{t>0\}.

Noting that ψ\displaystyle\psi is in C1,α\displaystyle C^{1,\alpha} near 0\displaystyle 0, which implies that the coefficients of the operator 𝒬\displaystyle\mathcal{Q} are in Cα\displaystyle C^{\alpha}. By using the elliptic regularity in Section 9 in [1], we obtain that ϕ(S)\displaystyle\phi(S) is in C2,α\displaystyle C^{2,\alpha} near 0\displaystyle 0. Furthermore, the free boundary Γ\displaystyle\Gamma can be denoted by y=ϕ(s,0)=ϕ(x,0)\displaystyle y=\phi(s,0)=\phi(x,0), and thus the free boundary Γ\displaystyle\Gamma is C2,α\displaystyle C^{2,\alpha} near 0\displaystyle 0.

Applying the Schauder estimates for elliptic equations in [1], we can obtain the C2,α\displaystyle C^{2,\alpha} regularity of ψ\displaystyle\psi up to the free boundary Γ\displaystyle\Gamma. Using the above arguments again, we can conclude that the free boundary Γ\displaystyle\Gamma is C3,α\displaystyle C^{3,\alpha} near 0\displaystyle 0, provided that λ(X)C1,β\displaystyle\lambda(X)\in C^{1,\beta} and fC1,β\displaystyle f\in C^{1,\beta}.

By bootstrap argument, we can obtain the higher regularity of the free boundary Γ\displaystyle\Gamma.

Finally, if f(s)\displaystyle f(s) is analytic and λ(X)\displaystyle\lambda(X) is analytic, by virtue of the results of Section 6.7 in [28], we can conclude that ϕ\displaystyle\phi is analytic. Hence, we obtain the analyticity of the free boundary Γ\displaystyle\Gamma.

Remark 3.1.

The results in this paper can be extended for the n\displaystyle n-dimensional case (n3\displaystyle n\geq 3), and Theorem 3.14 replaced by Corollary 3.11 in [24].

4. Applications: steady ideal jet flow and cavitational flow with general vorticity

As an important application of the mathematical theory of the free boundary problem (1.1), we will investigate the well-posedness of the steady incompressible inviscid fluid with free streamline. There are at least two classical hydrodynamical problems of two-dimensional steady flows with non-trivial vorticity can be described mathematically as a free boundary problem (1.1) for a semilinear elliptic equation, the one is the two-dimensional incompressible inviscid jet flow problem and another one is the two-dimensional incompressible inviscid cavitational flow problem. For a classical example, we will investigate briefly the existence and uniqueness of the two-dimensional incompressible inviscid jet flow issuing from a given semi-infinitely long nozzle in this section, and state the similar results on the incompressible cavitational flow problem.

In the previous sections, the technique of variational method has provided some solutions to the free boundary problem (1.1). The regularity of the free boundary and the regularity of the derivative to the solution ψ\displaystyle\psi up to it, follows from Theorem 3.15. However, the common topological properties of the free boundary between Ω{ψ>0}\displaystyle\Omega\cap\{\psi>0\} and Ω{ψ=0}\displaystyle\Omega\cap\{\psi=0\} are still unknown. Of course, it is an important and natural question to ask if the free boundary is smooth enough to provide a classical solution of the steady incompressible jet flow problem under consideration. The main aim of this section is to establish the existence and uniqueness of the steady incompressible jet flow issuing from symmetric semi-infinitely long nozzle via the mathematical theory on the free boundary problem established before.

4.1. Statement of the physical problem and main results

The problem we address is that the flow of an incompressible, inviscid fluid issues from a given symmetric semi-infinitely long nozzle and emerges as a jet with two symmetric free boundaries (see Figure 5). The flow is assumed to be both steady and irrotational. The free boundary Γ\displaystyle\Gamma initiates smoothly at the end point of the nozzle wall and extends to infinity in downstream, where the jet flow tends to some uniform flow.

Refer to caption
Figure 5. Symmetric jet flow problem

Denote N={y=g(x),x(,0]}\displaystyle N=\{y=g(x),x\in(-\infty,0]\} the upper nozzle wall of the semi-infinitely long plane symmetric nozzle, which satisfies that

g(x)C2,α((,0])\displaystyle g(x)\in C^{2,\alpha}((-\infty,0]), limxg(x)=H\displaystyle\lim_{x\rightarrow-\infty}g(x)=H and g(0)=a=minx0g(x)\displaystyle g(0)=a=\min_{x\leq 0}g(x). (4.1)

Let T={y=0}\displaystyle T=\{y=0\} be the symmetric axis, and A=(0,a)\displaystyle A=(0,a) be the end point of the nozzle wall.

The problem of incompressible jet flow with general vorticity can be formulated as the free boundary problem of finding a domain Ω0\displaystyle\Omega_{0} in (x,y)\displaystyle(x,y)-plane, whose boundary consists of the upper nozzle wall N\displaystyle N, the symmetric axis T\displaystyle T, and a priori unknown curve Γ\displaystyle\Gamma expressed by y=k(x)\displaystyle y=k(x) for x0\displaystyle x\geq 0 with

k(0)=g(0)=aandk(0)=g(0).k(0)=g(0)=a\ \ \ \text{and}\ \ \ k^{\prime}(0)=g^{\prime}(0). (4.2)

The condition (4.2) is the so-called continuous fit condition and smooth fit condition, respectively.

A vector (u,v,p)\displaystyle(u,v,p) representing the horizontal velocity, the vertical velocity and the pressure of the flow in Ω0\displaystyle\Omega_{0}, which belongs to (C1,α(Ω0)C0(Ω¯0))3\displaystyle\left(C^{1,\alpha}(\Omega_{0})\cap C^{0}(\bar{\Omega}_{0})\right)^{3} and satisfies the steady incompressible Euler equations

{xu+yv=0,uxu+vyu+xp=0,uxv+vyv+yp=0,\left\{\begin{array}[]{ll}&\partial_{x}u+\partial_{y}v=0,\\ &u\partial_{x}u+v\partial_{y}u+\partial_{x}p=0,\\ &u\partial_{x}v+v\partial_{y}v+\partial_{y}p=0,\end{array}\right. (4.3)

and the slip boundary condition

(u,v)n=0onNT,(u,v)\cdot\vec{n}=0\ \ \ \text{on}\ \ \ N\cup T, (4.4)

where n\displaystyle\vec{n} is the normal direction of the boundary.

The free boundary Γ\displaystyle\Gamma is assumed to be a material surface of the incompressible fluid, and then the velocity still satisfies the slip boundary condition (4.4) on Γ\displaystyle\Gamma. Moreover, the classical assumption (neglecting the effects of surface tension) is constant pressure condition, namely,

p=patmonΓ.p=p_{atm}\ \ \ \text{on}\ \ \ \Gamma.

Here, we denote patm\displaystyle p_{atm} the constant atmosphere pressure.

In the upstream, we assume that

u(x,y)u0(y)andv(x,y)0asx,u(x,y)\rightarrow u_{0}(y)\ \ \text{and}\ \ v(x,y)\rightarrow 0\ \ \text{as}\ \ x\rightarrow-\infty, (4.5)

and denote pin\displaystyle p_{in} the constant champer pressure in the inlet of the nozzle. Here, the variation and amplitude of the horizontal velocity u0(y)\displaystyle u_{0}(y) are arbitrary, and it implies that the vorticity of the jet flow is arbitrary in the upstream.

Remark 4.1.

Once we impose the horizontal velocity u0(y)\displaystyle u_{0}(y) in the inlet of the nozzle, the mass flux of incoming flow is determined by

Q=0Hu0(y)𝑑y.Q=\int_{0}^{H}u_{0}(y)dy.

Before we state the well-posedness results on the incompressible jet problem, it should be noted that there are two invariant quantities for the steady incompressible inviscid fluid, i.e.,

(u,v)ω=0,(u,v)\cdot\nabla\omega=0, (4.6)

and

(u,v)(12(u2+v2)+p)=0,(u,v)\cdot\nabla\left(\frac{1}{2}(u^{2}+v^{2})+p\right)=0, (4.7)

where ω=vxuy\displaystyle\omega=v_{x}-u_{y} denotes the vorticity of the fluid in two dimensions. In particular, the relation (4.7) gives that quantity 12(u2+v2)+p\displaystyle\frac{1}{2}(u^{2}+v^{2})+p called Bernoulli’s function remains invariant along the each streamline, which tells us two facts,
(1) the speed remains a constant denoted as λ\displaystyle\lambda along the free boundary Γ\displaystyle\Gamma.
(2) along the upper nozzle wall and the free boundary Γ\displaystyle\Gamma,

λ=u02(H)+2(pinpatm)\lambda=\sqrt{u^{2}_{0}(H)+2(p_{in}-p_{atm})}

as long as the continuous fit condition of the free boundary holds, where pin\displaystyle p_{in} denotes the uniform pressure in the upstream. Here, the quantity pdiff=pinpatm\displaystyle p_{diff}=p_{in}-p_{atm} is nothing but the pressure difference between the upstream and the downstream. It should be noted that the quantity pdiff\displaystyle p_{diff} is an undermined parameter here, and we will show that the appropriate choice pdiff\displaystyle p_{diff} is guaranteed by the continuous fit condition.

The incompressible jet problem is stated as follows.

The incompressible jet flow problem. Given a symmetric semi-infinitely long nozzle N\displaystyle N, a horizontal velocity u0(y)\displaystyle u_{0}(y) in the inlet and a constant atmosphere pressure patm\displaystyle p_{atm} on the free surface, does there exist a unique symmetric incompressible inviscid jet flow issuing from the nozzle N\displaystyle N, and the free boundary Γ\displaystyle\Gamma initiates smoothly at the endpoint A\displaystyle A.

Moreover, we introduce the definition of a solution to the incompressible jet flow problem in the following.

Definition 4.1.

(A solution to the incompressible jet flow problem).
Given an atmosphere pressure patm\displaystyle p_{atm} and an incoming velocity u0(y)\displaystyle u_{0}(y), a vector (u,v,p,Γ)\displaystyle(u,v,p,\Gamma) is called a solution to the incompressible jet flow problem, provided that
(1) The free boundary Γ\displaystyle\Gamma can be expressed by a C1\displaystyle C^{1}-smooth function y=k(x)>0\displaystyle y=k(x)>0 for any x[0,+)\displaystyle x\in[0,+\infty), and there exists an appropriate pressure difference pdiff\displaystyle p_{diff}, such that Γ\displaystyle\Gamma satisfies the continuous and smooth fit condition (4.2).
(2) (u,v,p)(C1,α(Ω0)C(Ω¯0))3\displaystyle(u,v,p)\in\left(C^{1,\alpha}(\Omega_{0})\cap C(\bar{\Omega}_{0})\right)^{3} solves the Euler system (4.3), and satisfies the boundary condition (4.4).
(3) There exists a unique positive constant h(0,a]\displaystyle h\in(0,a] such that

k(x)handk(x)0as x+.k(x)\rightarrow h\ \ \text{and}\ \ k^{\prime}(x)\rightarrow 0\ \ \text{as $\displaystyle x\rightarrow+\infty$}.

where h\displaystyle h is the asymptotic height of the free boundary in downstream.
(4) p=patm\displaystyle p=p_{atm} on Γ\displaystyle\Gamma.
(5) (u,v)(u0(y),0)\displaystyle(u,v)\rightarrow(u_{0}(y),0) uniformly for y(0,H)\displaystyle y\in(0,H), as x\displaystyle x\rightarrow-\infty.

The main results on the existence and uniqueness of the incompressible jet flow problem read as follows.

Theorem 4.1.

Suppose that the nozzle wall N\displaystyle N satisfies the assumption condition (4.1). Assume that the horizontal velocity in upstream u0(y)C2,β([0,H])\displaystyle u_{0}(y)\in C^{2,\beta}([0,H]) satisfies that

u0(y)>0,u0(0)=0,andu0′′(y)0for any y[0,H].u_{0}(y)>0,\ u_{0}^{\prime}(0)=0,\ \ \text{and}\ \ u_{0}^{\prime\prime}(y)\geq 0\ \ \text{for any $\displaystyle y\in[0,H]$}. (4.8)

Then, there exist a unique difference pressure pdiff0\displaystyle p_{diff}\geq 0 and a unique solution (u,v,p,Γ)\displaystyle(u,v,p,\Gamma) to the incompressible jet flow problem, such that
(1) The jet flow satisfies the following asymptotic behavior in the far fields,

(u,v,p)(u0(y),0,pin),u(0,u0(y)),v0,p0,(u,v,p)\rightarrow(u_{0}(y),0,p_{in}),\ \nabla u\rightarrow(0,u_{0}^{\prime}(y)),~~\nabla v\rightarrow 0,~~\nabla p\rightarrow 0,

uniformly in any compact subset of (0,H)\displaystyle(0,H), as x\displaystyle x\rightarrow-\infty, and

(u,v,p)(u1(y),0,patm),u(0,u1(y)),v0,p0,(u,v,p)\rightarrow(u_{1}(y),0,p_{atm}),\ \nabla u\rightarrow(0,u_{1}^{\prime}(y)),~~\nabla v\rightarrow 0,~~\nabla p\rightarrow 0,

uniformly in any compact subset of (0,h)\displaystyle(0,h), as x+,\displaystyle x\rightarrow+\infty, where pin=pdiff+patm\displaystyle p_{in}=p_{diff}+p_{atm}, and u1(y)\displaystyle u_{1}(y) and h\displaystyle h are uniquely determined by u0(y)\displaystyle u_{0}(y), pin\displaystyle p_{in} and patm\displaystyle p_{atm}.
(2) u>0\displaystyle u>0 in Ω¯0\displaystyle\bar{\Omega}_{0}.
(3) hk(x)H¯=maxx0g(x)\displaystyle h\leq k(x)\leq\bar{H}=\max_{x\leq 0}g(x) for any x[0,)\displaystyle x\in[0,\infty).

Remark 4.2.

The assumption u0(0)=0\displaystyle u^{\prime}_{0}(0)=0 in (4.8) follows from the symmetry of the incompressible jet flow. However, if we consider an asymmetric incompressible jet flow, the condition u0(0)=0\displaystyle u^{\prime}_{0}(0)=0 can be instead of u0(H)0\displaystyle u^{\prime}_{0}(H)\geq 0.

Remark 4.3.

To obtain the continuous fit condition, we choose the difference pressure pdiff\displaystyle p_{diff} as a parameter, and then show that there exists a unique pdiff\displaystyle p_{diff}, such that the free boundary Γ\displaystyle\Gamma satisfies the continuous fit condition in (4.2).

Similarly, we can obtain the well-posedness results on the incompressible cavitational flow problem. We will give the statement of the physical problem and the well-posedness results as follows.

Given a two-dimensional obstacle as

N:y=g(x)C2,α((b,0]),g(b)=0and g(0)=a=maxbx0g(x).N:y=g(x)\in C^{2,\alpha}((-b,0]),\ g(-b)=0\ \text{and $\displaystyle g(0)=a=\max_{-b\leq x\leq 0}g(x)$}. (4.9)

and denote T0={(x,0)<x<+}\displaystyle T_{0}=\{(x,0)\mid-\infty<x<+\infty\} the asymmetric axis and T1={(x,H)<x<+}\displaystyle T_{1}=\{(x,H)\mid-\infty<x<+\infty\} (see Figure 6).

Refer to caption
Figure 6. Symmetric cavitational flow

The incompressible cavitational flow problem. Given a two-dimensional symmetric obstacle N\displaystyle N, atmosphere pressure patm\displaystyle p_{atm} and the horizontal velocity u0(y)\displaystyle u_{0}(y) in the upstream, does there exist a unique incompressible symmetric inviscid cavitational flow past the given obstacle, and the free boundary Γ\displaystyle\Gamma initiates smoothly at the corner A\displaystyle A of the obstacle?

Definition 4.2.

(A solution to the incompressible cavitational flow problem).
A vector (u,v,p,Γ)\displaystyle(u,v,p,\Gamma) is called a solution to the incompressible cavitational flow problem, provided that
(1) The free boundary Γ\displaystyle\Gamma can be expressed by a C1\displaystyle C^{1}-smooth function y=k(x)>0\displaystyle y=k(x)>0 for any x[0,+)\displaystyle x\in[0,+\infty), such that NΓ\displaystyle N\cup\Gamma is C1\displaystyle C^{1}, namely,

g(0)=k(0)=aandg(0)=k(0).g(0)=k(0)=a~~\text{and}~~g^{\prime}(0)=k^{\prime}(0). (4.10)

(2) (u,v,p)(C1,α(Ω0)C(Ω¯0))3\displaystyle(u,v,p)\in\left(C^{1,\alpha}(\Omega_{0})\cap C(\bar{\Omega}_{0})\right)^{3} solves the Euler system (4.3), and satisfies the boundary condition (4.6), where Ω0\displaystyle\Omega_{0} is the flow field bounded by N,T0,T1\displaystyle N,T_{0},T_{1} and Γ\displaystyle\Gamma.
(3) There exists a positive constant h(a,H)\displaystyle h\in(a,H) such that

k(x)handk(x)0asx+.k(x)\rightarrow h\ \ \text{and}\ \ k^{\prime}(x)\rightarrow 0\ \ \text{as}\ \ x\rightarrow+\infty.

where h\displaystyle h is the asymptotic height of the free boundary in downstream.
(4) p=patm\displaystyle p=p_{atm} on Γ\displaystyle\Gamma.
(5) (u,v)(u0(y),0)\displaystyle(u,v)\rightarrow(u_{0}(y),0) uniformly for y(0,H)\displaystyle y\in(0,H), as x\displaystyle x\rightarrow-\infty.

We give the results as follows.

Theorem 4.2.

Suppose that the solid wall N\displaystyle N satisfies the assumption (4.9). Given an atmosphere pressure patm\displaystyle p_{atm} and u0(y)C2,β([0,H])\displaystyle u_{0}(y)\in C^{2,\beta}([0,H]), which satisfies that

u0(y)>0,u0(0)=0andu0′′(y)0for any y[0,H].u_{0}(y)>0,\ \ u_{0}^{\prime}(0)=0\ \ \text{and}\ \ u_{0}^{\prime\prime}(y)\geq 0\ \ \text{for any $\displaystyle y\in[0,H]$}. (4.11)

Then, there exist a unique difference pressure pdiff=puppatm\displaystyle p_{diff}=p_{up}-p_{atm} and a unique solution (u,v,p,Γ)\displaystyle(u,v,p,\Gamma) to the cavitational flow problem, where pin\displaystyle p_{in} denotes the pressure in the inlet of the channel. Furthermore,
(1) The cavitational flow satisfies the following asymptotic behavior in the far fields,

(u,v,p)(u0(y),0,pin),u(0,u0(y)),v0,p0,(u,v,p)\rightarrow(u_{0}(y),0,p_{in}),\ \nabla u\rightarrow(0,u_{0}^{\prime}(y)),~~\nabla v\rightarrow 0,~~\nabla p\rightarrow 0,

uniformly in any compact subset of (0,H)\displaystyle(0,H), as x\displaystyle x\rightarrow-\infty, and

(u,v,p)(u1(y),0,patm),u(0,u1(y)),v0,p0,(u,v,p)\rightarrow(u_{1}(y),0,p_{atm}),\ \nabla u\rightarrow(0,u_{1}^{\prime}(y)),~~\nabla v\rightarrow 0,~~\nabla p\rightarrow 0,

uniformly in any compact subset of (h,H)\displaystyle(h,H), as x+,\displaystyle x\rightarrow+\infty, where u1(y)\displaystyle u_{1}(y) and h\displaystyle h are uniquely determined by u0(y)\displaystyle u_{0}(y), pin\displaystyle p_{in} and patm\displaystyle p_{atm}.
(2) u>0\displaystyle u>0 in Ω¯0\displaystyle\bar{\Omega}_{0}.
(3) k(x)>0\displaystyle k(x)>0 for any x>0\displaystyle x>0.

4.2. Reformulation of the free boundary problem

In the following, we will reformulate the original physical problem into a free boundary problem of a semilinear elliptic equation. The similar idea has been adapted in the compressible subsonic flows with general vorticity in an infinitely long nozzle in [17, 18, 35].

First, it follows from the continuity equation in the incompressible Euler system (4.3) that there exists a stream function ψ\displaystyle\psi, such that

u=ψyandv=ψx.u=\psi_{y}\ \ ~~\text{and}~~\ \ v=-\psi_{x}. (4.12)

Second, suppose that the streamlines are well-defined in the whole flow fluid, and thus the incompressible jet flow problem can be solved along the each streamline as follows. On another hand, the positivity of the horizontal velocity of the jet will be verified later, which gives the well-definedness of the streamlines in the whole flow fluid.

Refer to caption
Figure 7. The streamline in the fluid field

Denote Ω={<x0,0<y<g(x)}{x>0,y>0}\displaystyle\Omega=\{-\infty<x\leq 0,0<y<g(x)\}\cup\{x>0,y>0\} the possible flow field (see Figure 7) and Ω0=Ω{0<ψ<Q}\displaystyle\Omega_{0}=\Omega\cap\{0<\psi<Q\} the flow field. For any point (x,y)Ω0\displaystyle(x,y)\in\Omega_{0}, it can be pulled back along one streamline to the point (,κ(ψ(x,y)))\displaystyle(-\infty,\kappa(\psi(x,y))) in the inlet. Thus

ψ(x,y)=0κ(ψ(x,y))u0(s)𝑑s,\psi(x,y)=\int_{0}^{\kappa(\psi(x,y))}u_{0}(s)ds,

which implies that

{1=u0(κ(ψ))κ(ψ),κ(0)=0.\left\{\begin{array}[]{ll}&1=u_{0}(\kappa(\psi))\kappa^{\prime}(\psi),\\ &\kappa(0)=0.\end{array}\right. (4.13)

It’s easy to see that the function κ(t)\displaystyle\kappa(t) can be solved uniquely by (4.13) provided that u0(y)\displaystyle u_{0}(y) is C2,β\displaystyle C^{2,\beta}-smooth. Meanwhile, it’s clear that κ(Q)=H\displaystyle\kappa(Q)=H. Due to the fact that the vorticity ω\displaystyle\omega is invariable on each streamline, we obtain the governing equation to the stream function,

Δψ=ω=u0(κ(ψ))f0(ψ)in the fluid field Ω0.\begin{array}[]{rl}-\Delta\psi=\omega=-u^{\prime}_{0}(\kappa(\psi))\triangleq f_{0}(\psi)\ \ \ \ \text{in the fluid field $\displaystyle\Omega_{0}$}.\end{array}

Furthermore, it is not difficult to check the following facts

(1) f0(t) is C1,β,(2) f0(0)=0,f0(Q)=u0(H)0,(3) f0(t)=u0′′(κ(t))u0(κ(t))0,\begin{array}[]{rl}&\text{(1) $\displaystyle f_{0}(t)$ is $\displaystyle C^{1,\beta}$,}\\ &\text{(2) $\displaystyle f_{0}(0)=0,f_{0}(Q)=-u_{0}^{\prime}(H)\leq 0$},\\ &\text{(3) $\displaystyle f_{0}^{\prime}(t)=-\frac{u_{0}^{\prime\prime}(\kappa(t))}{u_{0}(\kappa(t))}\leq 0$},\end{array} (4.14)

provided that u0(y)\displaystyle u_{0}(y) satisfies the assumption (4.8).

On another hand, we can impose the Dirichlet boundary conditions,

ψ=QonNI0,andψ=0onT,\psi=Q~~\text{on}~~N\cup I_{0},~~\text{and}~\psi=0~\text{on}~T,

where I0={(0,y)ya}\displaystyle I_{0}=\{(0,y)\mid y\geq a\}. Moreover, the free boundary can be defined as

Γ=Ω{x>0}{ψ<Q}.\Gamma=\Omega\cap\{x>0\}\cap\partial{\{\psi<Q\}}. (4.15)

The constant pressure condition on the free boundary together with the Bernoulli’s law gives that the speed remains a constant on Γ\displaystyle\Gamma, denote λ\displaystyle\lambda the constant speed, i.e.,

|ψ|=λ=u02(H)+2pdiffon Γ.|\nabla\psi|=\lambda=\sqrt{u^{2}_{0}(H)+2p_{diff}}\ \ \text{on $\displaystyle\Gamma$}.

Clearly, the constant λ\displaystyle\lambda is determined uniquely by pdiff0\displaystyle p_{diff}\geq 0.

Therefore, we formulate the following free boundary problem of the stream function that

{Δψ=f0(ψ)inΩ{ψ<Q},ψ=QonNΓ,ψ=0onT,|ψ|=λonΓ.\left\{\begin{array}[]{ll}&-\Delta\psi=f_{0}(\psi)~~~~\ \ \ \text{in}~~\Omega\cap\{\psi<Q\},\\ &\psi=Q~~~~\ \ \ \text{on}~~N\cup\Gamma,\ \ \ \ \ \psi=0~~~~\text{on}~~T,\\ &|\nabla\psi|=\lambda~~~~\ \ \ \text{on}~~\Gamma.\end{array}\right. (4.16)

4.3. Uniqueness of asymptotic behavior in downstream

Next, we will show that the asymptotic behavior of the jet flow in downstream can be determined by the state (u0(y),0)\displaystyle(u_{0}(y),0) of the incoming flow and the difference pressure pdiff\displaystyle p_{diff}.

For any initiate point (,s)\displaystyle(-\infty,s) in the inlet, the well-definedness of the streamlines implies that there exists a unique point denoted (0,χ(s;pdiff))\displaystyle(0,\chi(s;p_{diff})) in the downstream for any s[0,H]\displaystyle s\in[0,H] (see Figure 8), which can be pulled back to the imposed initiate point (,s)\displaystyle(-\infty,s) along a streamline.

Refer to caption
Figure 8. The streamline in Ω0\displaystyle\Omega_{0}

Denote u1(y)\displaystyle u_{1}(y) the horizontal velocity in the downstream, the conservation of mass and Bernoulli’s law gives that

0su0(t)𝑑t=0χ(s;pdiff)u1(t)𝑑t,\int_{0}^{s}u_{0}(t)dt=\int_{0}^{\chi(s;p_{diff})}u_{1}(t)dt,

and

u02(s)2+pdiff=u12(χ(s;pdiff))2in the fluid field.\frac{u^{2}_{0}(s)}{2}+p_{diff}=\frac{u^{2}_{1}(\chi(s;p_{diff}))}{2}\ \ \text{in the fluid field}.

Here, the difference pressure pdiff0\displaystyle p_{diff}\geq 0 between the inlet and the downstream is regarded as a parameter. These also give the initial value problem to the function χ(s;pdiff)\displaystyle\chi(s;p_{diff}) as follows,

{dχ(s;pdiff)ds=u0(s)u02(s)+2pdiff>0,χ(0;pdiff)=0,\left\{\begin{array}[]{ll}&\frac{d\chi(s;p_{diff})}{ds}=\frac{u_{0}(s)}{\sqrt{u_{0}^{2}(s)+2p_{diff}}}>0,\\ &\chi(0;p_{diff})=0,\end{array}\right. (4.17)

for any s[0,H]\displaystyle s\in[0,H] and pdiff0\displaystyle p_{diff}\geq 0. Thus,

χ(s;pdiff)=0su0(t)u02(t)+2pdiff𝑑ts.\chi(s;p_{diff})=\int_{0}^{s}\frac{u_{0}(t)}{\sqrt{u_{0}^{2}(t)+2p_{diff}}}dt\leq s. (4.18)

It is easy to check that χ(s;pdiff)\displaystyle\chi(s;p_{diff}) is strictly decreasing with respect to the parameter pdiff\displaystyle p_{diff}. In particular, the asymptotic height of the jet in downstream is in fact h=χ(H;pdiff)\displaystyle h=\chi(H;p_{diff}).

Noting that dχ(s;pdiff)ds>0\displaystyle\frac{d\chi(s;p_{diff})}{ds}>0 for any s[0,H]\displaystyle s\in[0,H] and pdiff0\displaystyle p_{diff}\geq 0, thus there exists an inverse function χ1(t;pdiff)\displaystyle\chi^{-1}(t;p_{diff}), such that s=χ1(t;pdiff)\displaystyle s=\chi^{-1}(t;p_{diff}) for any t[0,h]\displaystyle t\in[0,h]. Then the horizontal velocity u1(y)\displaystyle u_{1}(y) in downstream can be solved uniquely by

u1(t)=u02(χ1(t;pdiff))+2pdifffor any t[0,h],u_{1}(t)=\sqrt{u_{0}^{2}(\chi^{-1}(t;p_{diff}))+2p_{diff}}\ \ \text{for any $\displaystyle t\in[0,h]$}, (4.19)

once χ(s;pdiff)\displaystyle\chi(s;p_{diff}) is solved by the initiate value problem (4.17) for any pdiff0\displaystyle p_{diff}\geq 0.

Remark 4.4.

For any pdiff0\displaystyle p_{diff}\geq 0, it is easy to check that

dχ(H;pdiff)dpdiff<0 andλ=u02(H)+2pdiffu0(H)λ00.\text{$\displaystyle\frac{d\chi(H;p_{diff})}{dp_{diff}}<0$ and}\ \ \lambda=\sqrt{u^{2}_{0}(H)+2p_{diff}}\geq u_{0}(H)\triangleq\lambda_{0}\geq 0.

Moreover, the asymptotic height h=χ(H;pdiff)=χ(H;λ2u02(H)2)=h(λ)hλ\displaystyle h=\chi(H;p_{diff})=\chi\left(H;\frac{\lambda^{2}-u^{2}_{0}(H)}{2}\right)=h(\lambda)\triangleq h_{\lambda} is strictly increasing with respect to λ\displaystyle\lambda, and hλ0=H\displaystyle h_{\lambda_{0}}=H.

4.4. The variational approach

To solve the free boundary problem (4.16), we introduce a variational problem with a parameter λλ0\displaystyle\lambda\geq\lambda_{0}. For any L>H¯=maxx0g(x)\displaystyle L>\bar{H}=\max_{x\leq 0}g(x), denote ΩL\displaystyle\Omega_{L} the truncated domain (see Figure 9),

ΩL=Ω{(x,y)L<x<L,y<L},DL=ΩL{x>0},NL=N{x>L},\Omega_{L}=\Omega\cap\{(x,y)\mid-L<x<L,y<L\},\ \ \ \ D_{L}=\Omega_{L}\cap\{x>0\},\ \ \ N_{L}=N\cap\{x>-L\},

and

σL={(L,y)|0yg(L)}andσL={(L,y)|0yL},\sigma_{-L}=\{(-L,y)|~0\leq y\leq g(-L)\}\ \ \ \text{and}\ \ \ \ \sigma_{L}=\{(L,y)|~0\leq y\leq L\},

and

TL=T{LxL},I0,L={(0,y)ayL}andlL=BL/2((L/2,L)){yL}.T_{L}=T\cap\{-L\leq x\leq L\},\ I_{0,L}=\{(0,y)\mid a\leq y\leq L\}\ \ \text{and}\ \ l_{L}=\partial B_{L/2}\left(\left(L/2,L\right)\right)\cap\{y\geq L\}.
Refer to caption
Figure 9. The truncated domain ΩL\displaystyle\Omega_{L}

Since, a priorily, one does not know whether the stream function ψ\displaystyle\psi satisfies that 0ψQ\displaystyle 0\leq\psi\leq Q. Hence, we need extend the function f0(t)\displaystyle f_{0}(t) as follows

f~0(t)={f0(Q)+f0(Q)2iftQ+1,f0(Q)+f0(Q)(tQ(tQ)22)ifQtQ+1,f0(t)if0tQ,f0(0)(t+t22)if1t0,f0(0)2ift1.\tilde{f}_{0}(t)=\left\{\begin{array}[]{ll}f_{0}(Q)+\frac{f_{0}^{\prime}(Q)}{2}&\text{if}~~t\geq Q+1,\\ f_{0}(Q)+f_{0}^{\prime}(Q)\left(t-Q-\frac{(t-Q)^{2}}{2}\right)&\text{if}~~Q\leq t\leq Q+1,\\ f_{0}(t)&\text{if}~~0\leq t\leq Q,\\ f_{0}^{\prime}(0)\left(t+\frac{t^{2}}{2}\right)&\text{if}~~-1\leq t\leq 0,\\ -\frac{f_{0}^{\prime}(0)}{2}&\text{if}~~t\leq-1.\end{array}\right. (4.20)

Set

F0(t)=2tQf~0(t)𝑑s and f(t)=f~0(Qt).\text{$\displaystyle F_{0}(t)=2\int_{t}^{Q}\tilde{f}_{0}(t)ds$\ \ \ and \ \ \ $\displaystyle f(t)=-\tilde{f}_{0}(Q-t)$}. (4.21)

For y[0,g(L)]\displaystyle y\in[0,g(-L)], define a function ΨL(y)\displaystyle\Psi_{-L}(y), which satisfies that

ΨL′′(y)=f~0(ΨL(y)),ΨL(0)=0,ΨL(g(L))=Q.\Psi_{-L}^{\prime\prime}(y)=-\tilde{f}_{0}(\Psi_{-L}(y)),\ \Psi_{-L}(0)=0,\ \Psi_{-L}(g(-L))=Q.

By virtue of (4.14), it is easy to check that

0<ΨL(y)<QandΨL(y)0for any y(0,g(L)).0<\Psi_{-L}(y)<Q\ \ \text{and}\ \ \Psi^{\prime}_{-L}(y)\geq 0\ \ \text{for any $\displaystyle y\in(0,g(-L))$.}

Firstly, we give the following functional that

Jλ,L(ψ)=ΩL(|ψ|2+F0(ψ)+λ2I{ψ<Q}DL)𝑑X,J_{\lambda,L}(\psi)=\int_{\Omega_{L}}\left(|\nabla\psi|^{2}+F_{0}(\psi)+\lambda^{2}I_{\{\psi<Q\}\cap D_{L}}\right)dX,

where the admissible set is defined as follows

Kλ,L={ψHloc1(2)ψ=0lies belowT,ψ=Qlies aboveNI0,LlL,ψ=ΨLonσL,ψ=min{Ψλ,Q}onσL},\begin{array}[]{rl}K_{\lambda,L}=\{\psi\in H^{1}_{loc}(\mathbb{R}^{2})\mid&\psi=0~~\text{lies below}~T,\psi=Q~~\text{lies above}~N\cup I_{0,L}\cup l_{L},\\ &\psi=\Psi_{-L}~~~\text{on}~~\sigma_{-L},~~~\psi=\min\{\Psi_{\lambda},Q\}~~~\text{on}~~\sigma_{L}\},\end{array}

and

Ψλ(y)=0yu1(s)𝑑s.\Psi_{\lambda}(y)=\int_{0}^{y}u_{1}(s)ds. (4.22)

Truncated variational problem (Pλ,L)\displaystyle(P_{\lambda,L}): For any L>H¯\displaystyle L>\bar{H} and λλ0\displaystyle\lambda\geq\lambda_{0}, find a ψλ,LKλ,L\displaystyle\psi_{\lambda,L}\in K_{\lambda,L} such that

Jλ,L(ψλ,L)=minψKλ,LJλ,L(ψ).J_{\lambda,L}(\psi_{\lambda,L})=\min_{\psi\in K_{\lambda,L}}J_{\lambda,L}(\psi).
Remark 4.5.

Set ϕ=Qψ\displaystyle\phi=Q-\psi, by virtue of (4.14) and (4.21), the variational problem (Pλ,L)\displaystyle(P_{\lambda,L}) can be transformed into the variational problem in Section 2, and f(t)\displaystyle f(t) satisfies the condition (2.1). Thus, we can use the results in Section 2 and Section 3 in the following.

Proposition 4.3.

For any λλ0\displaystyle\lambda\geq\lambda_{0} and L>H¯\displaystyle L>\bar{H}, there exists a minimizer ψλ,L\displaystyle\psi_{\lambda,L} to the variational problem (Pλ,L)\displaystyle(P_{\lambda,L}) with 0ψλ,L(x,y)Q\displaystyle 0\leq\psi_{\lambda,L}(x,y)\leq Q in ΩL\displaystyle\Omega_{L}.

Proof.

The existence of the minimizer to the variational problem Pλ,L\displaystyle P_{\lambda,L} follows from Lemma 2.1, and denote ψλ,L\displaystyle\psi_{\lambda,L} the minimizer to the variational problem (Pλ,L)\displaystyle(P_{\lambda,L}).

Denote ψ=ψλ,L\displaystyle\psi=\psi_{\lambda,L} and ψε=ψεmin{ψ,0}\displaystyle\psi^{\varepsilon}=\psi-\varepsilon\min\{\psi,0\} for any ε(0,1)\displaystyle\varepsilon\in(0,1). Thus, ψεKλ,L\displaystyle\psi^{\varepsilon}\in K_{\lambda,L} and ψεψ\displaystyle\psi^{\varepsilon}\geq\psi. Moreover, ψ<Q\displaystyle\psi<Q if and only if ψε<Q\displaystyle\psi^{\varepsilon}<Q in ΩL\displaystyle\Omega_{L}, and we have

0Jλ,L(ψε)Jλ,L(ψ)ΩL{ψ<0}((1ε)21)|ψ|2εF0((1ε)ψ)ψdX0,\begin{array}[]{rl}0\leq&J_{\lambda,L}(\psi^{\varepsilon})-J_{\lambda,L}(\psi)\\ \leq&\int_{\Omega_{L}\cap\{\psi<0\}}((1-\varepsilon)^{2}-1)|\nabla\psi|^{2}-\varepsilon F^{\prime}_{0}((1-\varepsilon)\psi)\psi dX\\ \leq&0,\end{array}

due to F(t)0\displaystyle F^{\prime}(t)\leq 0 for t0\displaystyle t\leq 0, which implies that

ψ(x,y)0 in ΩL.\text{$\displaystyle\psi(x,y)\geq 0$ in $\displaystyle\Omega_{L}$}.

Similarly, we can show that

ψ(x,y)Q in ΩL.\text{$\displaystyle\psi(x,y)\leq Q$ in $\displaystyle\Omega_{L}$}.

Since 0ψλ,LQ\displaystyle 0\leq\psi_{\lambda,L}\leq Q in ΩL\displaystyle\Omega_{L}, we can remove the truncation of the function f~0(t)\displaystyle\tilde{f}_{0}(t) in (4.20). Denote the free boundary of ψλ,L\displaystyle\psi_{\lambda,L} as

Γλ,L=DL{ψλ,L<Q}.\Gamma_{\lambda,L}=D_{L}\cap\partial\{\psi_{\lambda,L}<Q\}.

With the aid of Lemma 2.1 and Theorem 3.15, we have

Proposition 4.4.

The minimizer ψλ,L\displaystyle\psi_{\lambda,L} satisfies that
(1) ψλ,LC0,1(ΩL)\displaystyle\psi_{\lambda,L}\in C^{0,1}(\Omega_{L}).
(2) Δψλ,L+f0(ψλ,L)=0\displaystyle\Delta\psi_{\lambda,L}+f_{0}(\psi_{\lambda,L})=0 in ΩL{ψλ,L<Q}\displaystyle\Omega_{L}\cap\{\psi_{\lambda,L}<Q\} and Δψλ,L+f0(ψλ,L)0\displaystyle\Delta\psi_{\lambda,L}+f_{0}(\psi_{\lambda,L})\leq 0 in ΩL\displaystyle\Omega_{L} in the weak sense. Furthermore, ψλ,L>0\displaystyle\psi_{\lambda,L}>0 in ΩL\displaystyle\Omega_{L} and ψλ,LC2,α(G)\displaystyle\psi_{\lambda,L}\in C^{2,\alpha}(G) for any compact subset G\displaystyle G of ΩL{ψλ,L<Q}\displaystyle\Omega_{L}\cap\{\psi_{\lambda,L}<Q\}, α(0,1)\displaystyle\alpha\in(0,1).
(3) The free boundary of the minimizer ψλ,L\displaystyle\psi_{\lambda,L} is C3,α\displaystyle C^{3,\alpha}.
(4) |ψλ,L|=λ\displaystyle|\nabla\psi_{\lambda,L}|=\lambda on the free boundary Γλ,L\displaystyle\Gamma_{\lambda,L}. Moreover, |ψλ,L|λ\displaystyle|\nabla\psi_{\lambda,L}|\geq\lambda on the segment lI0,L{ψλ,L<Q}\displaystyle l\in I_{0,L}\cap\partial\{\psi_{\lambda,L}<Q\} and |ψλ,L|λ\displaystyle|\nabla\psi_{\lambda,L}|\leq\lambda on the segment l{x=0,0<y<a}{ψλ,L<Q}\displaystyle l\in\{x=0,0<y<a\}\cap\partial\{\psi_{\lambda,L}<Q\}.

The Statement (1) in Proposition 4.4 gives the Lipschitz continuity of the minimizer in the interior of ΩL\displaystyle\Omega_{L}. Next, we will show that the minimizer ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) is also Lipschitz continuous near the boundary of ΩL\displaystyle\Omega_{L}.

Lemma 4.5.

ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) is Lipschitz continuous in every compact subsect of Ω¯L\displaystyle\bar{\Omega}_{L} that does not contain the point A\displaystyle A.

Proof.

We first consider the Lipschitz continuity of ψλ,L\displaystyle\psi_{\lambda,L} near the boundary DL\displaystyle\partial D_{L}. Denote ϕ(X)=Qψλ,L(X)\displaystyle\phi(X)=Q-\psi_{\lambda,L}(X) and d(X)=dist(X,Γλ,L)\displaystyle d(X)=\text{dist}(X,\Gamma_{\lambda,L}) and d1(X)=dist(X,DL)\displaystyle d_{1}(X)=\text{dist}(X,\partial D_{L}) for any XDL\displaystyle X\in D_{L}. We consider the following two cases.

Case 1. d(X)d1(X)\displaystyle d(X)\leq d_{1}(X), we can obtain the Lipschitz continuity of ϕ\displaystyle\phi at X\displaystyle X by using the similar arguments in the proof Theorem 2.4.

Case 2. d(X)>d1(X)\displaystyle d(X)>d_{1}(X). Without loss of generality, we assume that d1(X)=|XX0|\displaystyle d_{1}(X)=|X-X_{0}| for X0=(0,y)I0,L\displaystyle X_{0}=(0,y)\in I_{0,L} and X=(x,y)DL\displaystyle X=(x,y)\in D_{L}. Set s0=min{12,ya}\displaystyle s_{0}=\min\left\{\frac{1}{2},y-a\right\} and Bs0=Bs0(X0)\displaystyle B_{s_{0}}=B_{s_{0}}(X_{0}). Let φ\displaystyle\varphi be a solution to the following boundary problem,

{Δφ+f(φ)=0inBs0{x>0},φ=0onBs0{x=0},φ=QonBs0{x>0}.\left\{\begin{array}[]{ll}&\Delta\varphi+f(\varphi)=0\ \ \text{in}\ B_{s_{0}}\cap\{x>0\},\\ &\varphi=0\ \text{on}\ B_{s_{0}}\cap\{x=0\},\ \ \varphi=Q\ \text{on}\ \partial B_{s_{0}}\cap\{x>0\}.\end{array}\right.

Since ϕφ\displaystyle\phi\leq\varphi on (Bs0{x>0})\displaystyle\partial(B_{s_{0}}\cap\{x>0\}), the maximum principle gives that

ϕφin Bs0{x>0}.\phi\leq\varphi\ \ \text{in $\displaystyle B_{s_{0}}\cap\{x>0\}$}.

Denote φ~(X~)=φ(X0+s0X~)\displaystyle\tilde{\varphi}(\tilde{X})=\varphi(X_{0}+s_{0}\tilde{X}) for X~=(x~,y~)\displaystyle\tilde{X}=(\tilde{x},\tilde{y}), and Δφ~+s02f(φ~)=0\displaystyle\Delta\tilde{\varphi}+s_{0}^{2}f(\tilde{\varphi})=0 in B1(0){x~>0}\displaystyle B_{1}(0)\cap\{\tilde{x}>0\}. By using the elliptic estimates in [26], we have

φ~(X~)Cx~in B12(0){x~>0},\tilde{\varphi}(\tilde{X})\leq C\tilde{x}\ \ \text{in $\displaystyle B_{\frac{1}{2}}(0)\cap\{\tilde{x}>0\}$},

which implies that

ϕ(X)φ(X)Cxs0in Bs02(X0){x>0}.\phi(X)\leq\varphi(X)\leq C\frac{x}{s_{0}}\ \ \text{in $\displaystyle B_{\frac{s_{0}}{2}}(X_{0})\cap\{x>0\}$}. (4.23)

If r=x<s04\displaystyle r=x<\frac{s_{0}}{4}, we have

ϕ>0andΔϕ+f(ϕ)=0 in Br(X)Bs02{x>0}.\phi>0\ \ \text{and}\ \ \Delta\phi+f(\phi)=0\ \ \text{ in $\displaystyle B_{r}(X)\subset B_{\frac{s_{0}}{2}}\cap\{x>0\}$}.

Denote ϕ~(X~)=ϕ(X+rX~)r\displaystyle\tilde{\phi}(\tilde{X})=\frac{\phi(X+r\tilde{X})}{r} for X~=(x~,y~)\displaystyle\tilde{X}=(\tilde{x},\tilde{y}), it follows from (4.23) that

0<ϕ~<Cs0andΔϕ~+rf(rϕ~)=0 in B1(0).0<\tilde{\phi}<\frac{C}{s_{0}}\ \ \text{and}\ \ \Delta\tilde{\phi}+rf(r\tilde{\phi})=0\ \ \text{ in $\displaystyle B_{1}(0)$}.

By using the elliptic estimates in [26], we have

|ϕ(X)|=|ϕ~(0)|C(ϕ~L(B1(0))+rf(rϕ~)L(B1(0)))Cs0+Cs0.|\nabla\phi(X)|=|\nabla\tilde{\phi}(0)|\leq C(\|\tilde{\phi}\|_{L^{\infty}(B_{1}(0))}+r\|f(r\tilde{\phi})\|_{L^{\infty}(B_{1}(0))})\leq\frac{C}{s_{0}}+Cs_{0}.

If r=xs04\displaystyle r=x\geq\frac{s_{0}}{4}, the elliptic estimate in [26] gives that |ϕ(X)|C\displaystyle|\nabla\phi(X)|\leq C.

Since Δψλ,L+f0(ψλ,L)=0\displaystyle\Delta\psi_{\lambda,L}+f_{0}(\psi_{\lambda,L})=0 in ΩLDL\displaystyle\Omega_{L}\setminus D_{L}, the Lipschitz continuity of ψλ,L\displaystyle\psi_{\lambda,L} near (ΩLDL)\displaystyle\partial(\Omega_{L}\setminus D_{L}) can be obtained by using elliptic regularity.

4.5. The uniqueness, monotonicity and free boundary of the minimizer

We first give the uniqueness of the minimizer ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) for any given λλ0\displaystyle\lambda\geq\lambda_{0} and L>H¯\displaystyle L>\bar{H}, and show that ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) is monotone increasing with respect to y\displaystyle y.

Once we have the regularity of the free boundary Γλ,L\displaystyle\Gamma_{\lambda,L} at hand, we will establish some topological properties of the free boundary, provided that some special geometric conditions on the solid boundaries are assumed. For example, it’s desired that the free boundary is x\displaystyle x-graph, as long as we assume that the nozzle wall N\displaystyle N is a x\displaystyle x-graph and the horizontal velocity of the incoming flow is positive.

To obtain this topological property of the free boundary, we will establish the monotonicity of the minimizer ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) with respect to y\displaystyle y first.

Lemma 4.6.

The minimizer of the truncated variational problem (Pλ,L)\displaystyle(P_{\lambda,L}) is unique and ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) is monotone increasing with respect to y\displaystyle y.

Proof.

Assume that ψλ,L\displaystyle\psi_{\lambda,L} and ψ~λ,L\displaystyle\tilde{\psi}_{\lambda,L} are two minimizers to the truncated variational problem (Pλ,L)\displaystyle(P_{\lambda,L}). Set ψλ,Lε(x,y)=ψλ,L(x,yε)\displaystyle\psi_{\lambda,L}^{\varepsilon}(x,y)=\psi_{\lambda,L}(x,y-\varepsilon) in ΩLε={(x,y)(x,yε)ΩL}\displaystyle\Omega_{L}^{\varepsilon}=\{(x,y)\mid(x,y-\varepsilon)\in\Omega_{L}\} for any ε>0\displaystyle\varepsilon>0.

Noticing that ψλ,Lε\displaystyle\psi_{\lambda,L}^{\varepsilon} is a minimizer to the functional

Jλ,Lε(ψ)=ΩLε(|ψ|2+F0(ψ)+λ2I{ψ<Q}DLε)𝑑XJ_{\lambda,L}^{\varepsilon}(\psi)=\int_{\Omega^{\varepsilon}_{L}}\left(|\nabla\psi|^{2}+F_{0}(\psi)+\lambda^{2}I_{\{\psi<Q\}\cap D^{\varepsilon}_{L}}\right)dX

and the admissible set Kλ,Lε={ψψ(x,yε)Kλ,L}\displaystyle K^{\varepsilon}_{\lambda,L}=\{\psi\mid\psi(x,y-\varepsilon)\in K_{\lambda,L}\}, where DLε=ΩLε{x>0}\displaystyle D_{L}^{\varepsilon}=\Omega_{L}^{\varepsilon}\cap\{x>0\}. Extend ψλ,Lε=0\displaystyle\psi_{\lambda,L}^{\varepsilon}=0 in ΩLΩLε\displaystyle\Omega_{L}\setminus\Omega^{\varepsilon}_{L} and ψ~λ,L=Q\displaystyle\tilde{\psi}_{\lambda,L}=Q in ΩLεΩL\displaystyle\Omega^{\varepsilon}_{L}\setminus\Omega_{L}.

Denote ψ~=ψ~λ,L\displaystyle\tilde{\psi}=\tilde{\psi}_{\lambda,L} and ψε=ψλ,Lε\displaystyle\psi^{\varepsilon}=\psi_{\lambda,L}^{\varepsilon} for simplicity, it follows from Proposition 4.3 that

ϕ1=min{ψε,ψ~}Kλ,Lεandϕ2=max{ψε,ψ~}Kλ,L,\phi_{1}=\min\{\psi^{\varepsilon},\tilde{\psi}\}\in K^{\varepsilon}_{\lambda,L}\ \ \text{and}\ \ \phi_{2}=\max\{\psi^{\varepsilon},\tilde{\psi}\}\in K_{\lambda,L},

and

ΩL{ψε>ψ~}=ΩLε{ψε>ψ~}.\Omega_{L}\cap\{\psi^{\varepsilon}>\tilde{\psi}\}=\Omega_{L}^{\varepsilon}\cap\{\psi^{\varepsilon}>\tilde{\psi}\}. (4.24)

Next, we will show that

Jλ,Lε(ϕ1)=Jλ,Lε(ψε)andJλ,L(ϕ2)=Jλ,L(ψ~).J_{\lambda,L}^{\varepsilon}(\phi_{1})=J_{\lambda,L}^{\varepsilon}(\psi^{\varepsilon})\ \ \text{and}\ \ J_{\lambda,L}(\phi_{2})=J_{\lambda,L}(\tilde{\psi}). (4.25)

Since ψ~\displaystyle\tilde{\psi} and ψε\displaystyle\psi^{\varepsilon} are minimizers, it suffices to show that

Jλ,Lε(ϕ1)+Jλ,L(ϕ2)=Jλ,L(ψ~)+Jλ,Lε(ψε).J_{\lambda,L}^{\varepsilon}(\phi_{1})+J_{\lambda,L}(\phi_{2})=J_{\lambda,L}(\tilde{\psi})+J_{\lambda,L}^{\varepsilon}(\psi^{\varepsilon}). (4.26)

It follows from (4.24) that

ΩLε|ϕ1|2𝑑X+ΩL|ϕ2|2𝑑X=ΩLε{ψεψ~}|ψε|2𝑑X+ΩLε{ψε>ψ~}|ψ~|2𝑑X+ΩL{ψεψ~}|ψ~|2𝑑X+ΩL{ψε>ψ~}|ψε|2𝑑X=ΩLε{ψεψ~}|ψε|2𝑑X+ΩL{ψε>ψ~}|ψ~|2𝑑X+ΩL{ψεψ~}|ψ~|2𝑑X+ΩLε{ψε>ψ~}|ψε|2𝑑X=ΩL|ψ~|2𝑑X+ΩLε|ψε|2𝑑X.\begin{array}[]{rl}&\int_{\Omega_{L}^{\varepsilon}}|\nabla\phi_{1}|^{2}dX+\int_{\Omega_{L}}|\nabla\phi_{2}|^{2}dX\\ =&\int_{\Omega_{L}^{\varepsilon}\cap\{\psi^{\varepsilon}\leq\tilde{\psi}\}}|\nabla\psi^{\varepsilon}|^{2}dX+\int_{\Omega_{L}^{\varepsilon}\cap\{\psi^{\varepsilon}>\tilde{\psi}\}}|\nabla\tilde{\psi}|^{2}dX\\ &+\int_{\Omega_{L}\cap\{\psi^{\varepsilon}\leq\tilde{\psi}\}}|\nabla\tilde{\psi}|^{2}dX+\int_{\Omega_{L}\cap\{\psi^{\varepsilon}>\tilde{\psi}\}}|\nabla\psi^{\varepsilon}|^{2}dX\\ =&\int_{\Omega_{L}^{\varepsilon}\cap\{\psi^{\varepsilon}\leq\tilde{\psi}\}}|\nabla\psi^{\varepsilon}|^{2}dX+\int_{\Omega_{L}\cap\{\psi^{\varepsilon}>\tilde{\psi}\}}|\nabla\tilde{\psi}|^{2}dX\\ &+\int_{\Omega_{L}\cap\{\psi^{\varepsilon}\leq\tilde{\psi}\}}|\nabla\tilde{\psi}|^{2}dX+\int_{\Omega_{L}^{\varepsilon}\cap\{\psi^{\varepsilon}>\tilde{\psi}\}}|\nabla\psi^{\varepsilon}|^{2}dX\\ =&\int_{\Omega_{L}}|\nabla\tilde{\psi}|^{2}dX+\int_{\Omega_{L}^{\varepsilon}}|\nabla\psi^{\varepsilon}|^{2}dX.\end{array} (4.27)

Similarly, we can verify that

ΩLεF0(ϕ1)𝑑X+ΩLF0(ϕ2)𝑑X=ΩLεF0(ψε)𝑑X+ΩLF0(ψ~)𝑑X,\int_{\Omega_{L}^{\varepsilon}}F_{0}(\phi_{1})dX+\int_{\Omega_{L}}F_{0}(\phi_{2})dX=\int_{\Omega_{L}^{\varepsilon}}F_{0}(\psi^{\varepsilon})dX+\int_{\Omega_{L}}F_{0}(\tilde{\psi})dX,

and

DLεI{ϕ1<Q}𝑑X+DLI{ϕ2<Q}𝑑X=DLεI{ψε<Q}𝑑X+DLI{ψ~<Q}𝑑X,\begin{array}[]{rl}\int_{D_{L}^{\varepsilon}}I_{\{\phi_{1}<Q\}}dX+\int_{D_{L}}I_{\{\phi_{2}<Q\}}dX=\int_{D_{L}^{\varepsilon}}I_{\{\psi^{\varepsilon}<Q\}}dX+\int_{D_{L}}I_{\{\tilde{\psi}<Q\}}dX,\end{array}

which together with (4.27) yield (4.26).

Next, we claim that

if ψε(X0)=ψ~(X0)<Q for X0ΩL, then either ψεψ~ or ψεψ~ in Br(X0),\text{if $\displaystyle\psi^{\varepsilon}(X_{0})=\tilde{\psi}(X_{0})<Q$ for $\displaystyle X_{0}\in\Omega_{L}$, then either $\displaystyle\psi^{\varepsilon}\geq\tilde{\psi}$ or $\displaystyle\psi^{\varepsilon}\leq\tilde{\psi}$ in $\displaystyle B_{r}(X_{0})$}, (4.28)

for small r>0\displaystyle r>0. Suppose that the claim (4.28) is not true, the continuity of ψ~\displaystyle\tilde{\psi} and ψε\displaystyle\psi^{\varepsilon} give that 0<ψ~<Q\displaystyle 0<\tilde{\psi}<Q and 0<ψε<Q\displaystyle 0<\psi^{\varepsilon}<Q in Br(X0)\displaystyle B_{r}(X_{0}) for small r>0\displaystyle r>0, then we have that ϕ2\displaystyle\phi_{2} is not a solution of Δϕ2+f0(ϕ2)=0\displaystyle\Delta\phi_{2}+f_{0}(\phi_{2})=0 in Br(X0)\displaystyle B_{r}(X_{0}). In fact, if not, the maximum principle gives that max{ψ~,ψε}=ϕ2=ψ~\displaystyle\max\{\tilde{\psi},\psi^{\varepsilon}\}=\phi_{2}=\tilde{\psi} in Br(X0)\displaystyle B_{r}(X_{0}), due to ϕ2(X0)=ψ~(X0)\displaystyle\phi_{2}(X_{0})=\tilde{\psi}(X_{0}). This contradicts to our assumptions.

Let ϕ\displaystyle\phi be the solution of the following boundary value problem

{Δϕ+f0(ϕ)=0inBr(X0),ϕ=ϕ2onBr(X0).\left\{\begin{array}[]{ll}\Delta\phi+f_{0}(\phi)=0&\text{in}\ B_{r}(X_{0}),\\ \phi=\phi_{2}&\text{on}\ \partial B_{r}(X_{0}).\end{array}\right.

Thus ϕϕ2\displaystyle\phi\neq\phi_{2} in Br(X0)\displaystyle B_{r}(X_{0}), it is easy to check that

Br(X0)|ϕ|2+F0(ϕ)dXBr(X0)|ϕ2|2+F0(ϕ2)dXBr(X0)|(ϕϕ2)|2𝑑X<0.\begin{array}[]{rl}&\int_{B_{r}(X_{0})}|\nabla\phi|^{2}+F_{0}(\phi)dX-\int_{B_{r}(X_{0})}|\nabla\phi_{2}|^{2}+F_{0}(\phi_{2})dX\\ \leq&-\int_{B_{r}(X_{0})}|\nabla(\phi-\phi_{2})|^{2}dX<0.\end{array} (4.29)

Extend ϕ=ϕ2\displaystyle\phi=\phi_{2} in ΩLBr(X0)\displaystyle\Omega_{L}\setminus B_{r}(X_{0}), it follows from (4.25) and (4.29) that

Jλ,L(ϕ)<Jλ,L(ϕ2)=Jλ,L(ψ~),J_{\lambda,L}(\phi)<J_{\lambda,L}(\phi_{2})=J_{\lambda,L}(\tilde{\psi}),

which contradicts to the minimality of ψ~\displaystyle\tilde{\psi}.

Since ψ~>ψε\displaystyle\tilde{\psi}>\psi^{\varepsilon} near TL\displaystyle T_{L}, it follows from (4.28) that ψ~ψε\displaystyle\tilde{\psi}\geq\psi^{\varepsilon} in the connected component Ω0\displaystyle\Omega_{0} of {ψ~<Q}\displaystyle\{\tilde{\psi}<Q\} which contains an ΩL\displaystyle\Omega_{L}-neighborhood of TL\displaystyle T_{L}. In view of the boundary value of ψ~\displaystyle\tilde{\psi}, we have that {ψ~<Q}ΩL\displaystyle\{\tilde{\psi}<Q\}\cap\partial\Omega_{L} is a connected arc. The maximum principle gives that any component of {ψ~<Q}\displaystyle\{\tilde{\psi}<Q\} must touch the boundary {ψ~<Q}ΩL\displaystyle\{\tilde{\psi}<Q\}\cap\partial\Omega_{L}. Therefore, the domain ΩL{ψ~<Q}\displaystyle\Omega_{L}\cap\{\tilde{\psi}<Q\} is connected and

ψ(x,yε)=ψε(x,y)ψ~(x,y)inΩL.\psi(x,y-\varepsilon)=\psi^{\varepsilon}(x,y)\leq\tilde{\psi}(x,y)\ \text{in}~~\Omega_{L}. (4.30)

Similarly, we can show that

ψ(x,y+ε)ψ~(x,y)inΩL.\psi(x,y+\varepsilon)\geq\tilde{\psi}(x,y)~~\text{in}~~\Omega_{L}. (4.31)

Taking ε0\displaystyle\varepsilon\rightarrow 0 in (4.30) and (4.31), which yields that

ψ(x,y)=ψ~(x,y)inΩL.\psi(x,y)=\tilde{\psi}(x,y)~~\text{in}~~\Omega_{L}.

In particular, taking ψ=ψ~\displaystyle\psi=\tilde{\psi} in (4.30), we have that the minimizer ψ=ψλ,L(x,y)\displaystyle\psi=\psi_{\lambda,L}(x,y) is monotone increasing with respect to y\displaystyle y.

The monotonicity of ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) with respect to y\displaystyle y implies that the free boundary Γλ\displaystyle\Gamma_{\lambda} is a x\displaystyle x-graph, and then there exists a function y=kλ,L(x)\displaystyle y=k_{\lambda,L}(x) for x(0,L)\displaystyle x\in(0,L), such that

{ψλ,L<Q}DL={(x,y)DL0<y<kλ,L(x),0<x<L}.\{\psi_{\lambda,L}<Q\}\cap D_{L}=\{(x,y)\in D_{L}\mid 0<y<k_{\lambda,L}(x),0<x<L\}.

To obtain the continuity of kλ,L(x)\displaystyle k_{\lambda,L}(x), we first give the following non-oscillation lemma.

Lemma 4.7.

(Non-oscillation Lemma) Suppose that there exist some constants α1,α2\displaystyle\alpha_{1},\alpha_{2} with α1<α2\displaystyle\alpha_{1}<\alpha_{2} and a domain DDL{ψλ,L<Q}\displaystyle D\subset D_{L}\cap\{\psi_{\lambda,L}<Q\} with dist(A,D)>c0\displaystyle(A,D)>c_{0} for some c0>0\displaystyle c_{0}>0, which is bounded by two disjointed arcs γ1,γ2\displaystyle\gamma_{1},\gamma_{2} (γ1,γ2Γλ,L)\displaystyle\gamma_{1},\gamma_{2}\subset\Gamma_{\lambda,L}), the lines {y=α1}\displaystyle\{y=\alpha_{1}\} and {y=α2}\displaystyle\{y=\alpha_{2}\}. Denote (βi,α1)\displaystyle(\beta_{i},\alpha_{1}) and (ηi,α2)\displaystyle(\eta_{i},\alpha_{2}) the endpoints of γi\displaystyle\gamma_{i} for i=1,2\displaystyle i=1,2. Then there exists a constant C>0\displaystyle C>0, such that

α2α1Cmax{|β1β2|,|η1η2|}.\alpha_{2}-\alpha_{1}\leq C\max\{|\beta_{1}-\beta_{2}|,|\eta_{1}-\eta_{2}|\}.
Proof.

Denote h=max{|β1β2|,|η1η2|}\displaystyle h=\max\left\{|\beta_{1}-\beta_{2}|,|\eta_{1}-\eta_{2}|\right\} and ψ=ψλ,L\displaystyle\psi=\psi_{\lambda,L}. Since Δψ=f0(ψ)\displaystyle\Delta\psi=-f_{0}(\psi) in D\displaystyle D, we have

Dψν𝑑S=Df0(ψ)𝑑XDf0(Q)+CQdXCh(α2α1)Ch,\int_{\partial D}\frac{\partial\psi}{\partial\nu}dS=-\int_{D}f_{0}(\psi)dX\leq\int_{D}-f_{0}(Q)+CQdX\leq Ch(\alpha_{2}-\alpha_{1})\leq Ch,

which implies that

γ1γ2λ𝑑S=γ1γ2ψν𝑑SD({y=α1}{y=α2})|ψx|𝑑y+ChCh,\int_{\gamma_{1}\cup\gamma_{2}}\lambda dS=\int_{\gamma_{1}\cup\gamma_{2}}\frac{\partial\psi}{\partial\nu}dS\leq\int_{\partial D\cap(\{y=\alpha_{1}\}\cup\{y=\alpha_{2}\})}\left|\frac{\partial\psi}{\partial x}\right|dy+Ch\leq Ch, (4.32)

where we have used the Lipschitz continuity of ψ\displaystyle\psi due to Lemma 4.5.

On the other hand, one gets

γ1γ2λ𝑑S2λ(α2α1),\int_{\gamma_{1}\cup\gamma_{2}}\lambda dS\geq 2\lambda(\alpha_{2}-\alpha_{1}),

which together with (4.32) gives that

α2α1Ch.\alpha_{2}-\alpha_{1}\leq Ch.

With the aid of Lemma 4.7, we have

Lemma 4.8.

kλ,L(x)\displaystyle k_{\lambda,L}(x) is a continuous function in [0,L]\displaystyle[0,L].

Proof.

By using the non-oscillation Lemma 4.7, it follows from the similar arguments in Lemma 5.4 in [5] that kλ,L(x)\displaystyle k_{\lambda,L}(x) has at most one limit point as x0\displaystyle x\downarrow 0 and xL\displaystyle x\uparrow L. Moreover, kλ,L(x)\displaystyle k_{\lambda,L}(x) has most one limit point as xx0\displaystyle x\uparrow x_{0} or as xx0\displaystyle x\downarrow x_{0} for any x0(0,L)\displaystyle x_{0}\in(0,L). It suffices to show that

kλ,L(x+0)=kλ,L(x0)=kλ,L(x)for any x(0,L).k_{\lambda,L}(x+0)=k_{\lambda,L}(x-0)=k_{\lambda,L}(x)\ \ \text{for any $\displaystyle x\in(0,L)$}.

If not, without loss of generality, we assume that there exists a point x0(0,L)\displaystyle x_{0}\in(0,L), such that kλ,L(x0)<kλ,L(x0+0)\displaystyle k_{\lambda,L}(x_{0})<k_{\lambda,L}(x_{0}+0). Denote Iδ={(x0,y)kλ,L(x0)+δ<y<kλ,L(x0)+3δ}\displaystyle I_{\delta}=\{(x_{0},y)\mid k_{\lambda,L}(x_{0})+\delta<y<k_{\lambda,L}(x_{0})+3\delta\} with δ=kλ,L(x0+0)kλ,L(x0)4\displaystyle\delta=\frac{k_{\lambda,L}(x_{0}+0)-k_{\lambda,L}(x_{0})}{4}. The monotonicity of ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) with respect to y\displaystyle y and the Lipschitz continuity of ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) give that

ψλ,L=QonIδ,Δψλ,L+f0(ψλ,L)=0 and u=ψλ,Ly0 in Eδ,ε,\psi_{\lambda,L}=Q\ \ \text{on}\ I_{\delta},\ \ \text{$\displaystyle\Delta\psi_{\lambda,L}+f_{0}(\psi_{\lambda,L})=0$ and $\displaystyle u=\frac{\partial\psi_{\lambda,L}}{\partial y}\geq 0$ in $\displaystyle E_{\delta,\varepsilon}$,}

where Eδ,ε={(x,y)x0<x<x0+ε,kλ,L(x0)+δ<y<kλ,L(x0)+3δ}\displaystyle E_{\delta,\varepsilon}=\{(x,y)\mid x_{0}<x<x_{0}+\varepsilon,k_{\lambda,L}(x_{0})+\delta<y<k_{\lambda,L}(x_{0})+3\delta\} for small ε>0\displaystyle\varepsilon>0. Thus, Iδ\displaystyle I_{\delta} is a part of the free boundary Γλ,L\displaystyle\Gamma_{\lambda,L}, it follows from (4) in Proposition 4.4 that

ψλ,L(x0+0,y)x=|ψλ,L|=λonIδ.\frac{\partial\psi_{\lambda,L}(x_{0}+0,y)}{\partial x}=|\nabla\psi_{\lambda,L}|=\lambda\ \ \text{on}\ \ I_{\delta}. (4.33)

It is easy to check that

Δu+b(X)u=0inEδ,εandb(X)=f0(ψλ,L(X))0.\Delta u+b(X)u=0\ \ \text{in}\ \ E_{\delta,\varepsilon}\ \ \text{and}\ \ b(X)=f^{\prime}_{0}(\psi_{\lambda,L}(X))\leq 0.

The strong maximum principle gives that u>0\displaystyle u>0 in Eδ,ε\displaystyle E_{\delta,\varepsilon}. In view of u=0\displaystyle u=0 on Iδ\displaystyle I_{\delta}, thanks to Hopf’s lemma, we have

2ψλ,Lxy=ux>0onIδ.\frac{\partial^{2}\psi_{\lambda,L}}{\partial x\partial y}=\frac{\partial u}{\partial x}>0\ \ \text{on}\ \ I_{\delta}. (4.34)

On the other hand, it follows from (4.33) that

2ψλ,Lyx=0onIδ,\frac{\partial^{2}\psi_{\lambda,L}}{\partial y\partial x}=0\ \ \text{on}\ \ I_{\delta},

which contradicts to (4.34).

Hence, we obtain the continuity of the function kλ,L(x)\displaystyle k_{\lambda,L}(x).

Next, we will introduce the bounded gradient lemma for ψλ,L\displaystyle\psi_{\lambda,L} near the boundary of DL\displaystyle D_{L}.

Lemma 4.9.

For any X0DL\displaystyle X_{0}\in\partial D_{L}, if BR(X0)Γλ,L\displaystyle B_{R}(X_{0})\cap\Gamma_{\lambda,L}\neq\varnothing for any R>0\displaystyle R>0, then for small r>0\displaystyle r>0, there exists a constant C\displaystyle C independent of Q\displaystyle Q and r\displaystyle r, such that

|ψλ,L(X)|Cin Br(X0)ΩL.|\nabla\psi_{\lambda,L}(X)|\leq C\ \ \text{in $\displaystyle B_{r}(X_{0})\cap\Omega_{L}$}. (4.35)
Proof.

We first consider the case X0=A\displaystyle X_{0}=A. It suffices to show that

|ψλ,L(X)|Cin ΩL{ρ<|XA|<2ρ},|\nabla\psi_{\lambda,L}(X)|\leq C\ \ \text{in $\displaystyle\Omega_{L}\cap\{\rho<|X-A|<2\rho\}$},

for any small ρ>0\displaystyle\rho>0.

Denote ψρ(X~)=1ρψλ,L(A+ρX~)\displaystyle\psi_{\rho}(\tilde{X})=\frac{1}{\rho}\psi_{\lambda,L}(A+\rho\tilde{X}), D1={X~12<|X~|<52,A+ρX~ΩL}\displaystyle D_{1}=\left\{\tilde{X}\mid\frac{1}{2}<|\tilde{X}|<\frac{5}{2},A+\rho\tilde{X}\in\Omega_{L}\right\} and D2={X~1<|X~|<2,A+ρX~ΩL}\displaystyle D_{2}=\left\{\tilde{X}\mid 1<|\tilde{X}|<2,A+\rho\tilde{X}\in\Omega_{L}\right\}. It follows from Proposition 4.4 that

0ψρQρin D1ψρ=Qρ and ψρν=λ on the free boundary.0\leq\psi_{\rho}\leq\frac{Q}{\rho}\ \ \text{in $\displaystyle D_{1}$, $\displaystyle\psi_{\rho}=\frac{Q}{\rho}$ and $\displaystyle\frac{\partial\psi_{\rho}}{\partial\nu}=\lambda$ on the free boundary}.

Denote ϕρ=Qρψρ\displaystyle\phi_{\rho}=\frac{Q}{\rho}-\psi_{\rho}, it is easy to check that

0ϕρQρandΔϕρ+ρf(ρϕρ)=0 in D1{ϕρ>0}.0\leq\phi_{\rho}\leq\frac{Q}{\rho}\ \ \text{and}\ \ \Delta\phi_{\rho}+\rho f(\rho\phi_{\rho})=0\ \ \text{ in $\displaystyle D_{1}\cap\{\phi_{\rho}>0\}$}.

Obviously, D2D1\displaystyle D_{2}\subset D_{1}. Since the boundary D2{X~A+ρX~ΩL}\displaystyle\partial D_{2}\cap\{\tilde{X}\mid A+\rho\tilde{X}\in\partial\Omega_{L}\} is C2,α\displaystyle C^{2,\alpha}, and ϕρ=0\displaystyle\phi_{\rho}=0 on D2{X~A+ρX~ΩL}\displaystyle\partial D_{2}\cap\{\tilde{X}\mid A+\rho\tilde{X}\in\partial\Omega_{L}\}, then the Harnack’s inequality is still valid up to the boundary D2{X~A+ρX~ΩL}\displaystyle\partial D_{2}\cap\{\tilde{X}\mid A+\rho\tilde{X}\in\partial\Omega_{L}\}. By using the similar arguments in Theorem 2.4 and Lemma 4.5, we can obtain that

|ϕρ(X~)|Cin D2,|\nabla\phi_{\rho}(\tilde{X})|\leq C\ \ \text{in $\displaystyle D_{2}$},

where C\displaystyle C is a constant independent of Qρ\displaystyle\frac{Q}{\rho}. This implies the estimate (4.35).

For the other case X0DL\displaystyle X_{0}\in\partial D_{L} with BR(X0)Γλ,L\displaystyle B_{R}(X_{0})\cap\Gamma_{\lambda,L}\neq\varnothing for any R>0\displaystyle R>0, we can show that the estimate (4.35) is still valid by using the above arguments.

For the point X0=(x0,y0)(DLΓ¯λ,L)\displaystyle X_{0}=(x_{0},y_{0})\in(\partial D_{L}\cap\bar{\Gamma}_{\lambda,L}), there is a possible case that Br(X0)Γλ,L{x<x0}\displaystyle B_{r}(X_{0})\cap\Gamma_{\lambda,L}\cap\{x<x_{0}\}\neq\varnothing and Br(X0)Γλ,L{x>x0}\displaystyle B_{r}(X_{0})\cap\Gamma_{\lambda,L}\cap\{x>x_{0}\}\neq\varnothing for any r>0\displaystyle r>0, and thus for the C1\displaystyle C^{1}-regularity of ψλ,L\displaystyle\psi_{\lambda,L} and Γλ,L\displaystyle\Gamma_{\lambda,L} at X0\displaystyle X_{0}, here we can not use the smooth fit condition in Theorem 6.1 and Lemma 6.4 in [7] directly. (see Figure 10)

Therefore, we will estimate the C1\displaystyle C^{1}-regularity of ψλ,L\displaystyle\psi_{\lambda,L} and Γλ,L\displaystyle\Gamma_{\lambda,L} near the free boundary DLΓ¯λ,L\displaystyle\partial D_{L}\cap\bar{\Gamma}_{\lambda,L} as follows.

Refer to caption
Figure 10. The regularity near the free boundary
Proposition 4.10.

For any X0=(x0,y0)DL\displaystyle X_{0}=(x_{0},y_{0})\in\partial D_{L}, if Br(X0)Γλ,L\displaystyle B_{r}(X_{0})\cap\Gamma_{\lambda,L}\neq\varnothing for any r>0\displaystyle r>0, then we have

ψλ,L(X)λν0asXX0,XΩL{ψλ,L<Q},\nabla\psi_{\lambda,L}(X)\rightarrow\lambda\nu_{0}\ \ \text{as}\ \ \ X\rightarrow X_{0},\ \ X\in\Omega_{L}\cap\{\psi_{\lambda,L}<Q\},

where ν0\displaystyle\nu_{0} is the outer normal vector to DL{ψλ,L<Q}\displaystyle\partial D_{L}\cap\partial\{\psi_{\lambda,L}<Q\} at X0\displaystyle X_{0}. Moreover, kλ,L(x)\displaystyle k_{\lambda,L}(x) is C1\displaystyle C^{1}-smooth at X=X0\displaystyle X=X_{0}.

Proof.

Without loss of generality, we assume X0=(L2,3L2)\displaystyle X_{0}=\left(\frac{L}{2},\frac{3L}{2}\right) and ν0=(0,1)\displaystyle\nu_{0}=(0,1).

Suppose XnX0DL\displaystyle X_{n}\rightarrow X_{0}\in\partial D_{L} with XnDL\displaystyle X_{n}\in\partial D_{L}, and ϕ0\displaystyle\phi_{0} is a blow-up limit of ϕn(X)=Qψλ,L(Xn+rnX)rn\displaystyle\phi_{n}(X)=\frac{Q-\psi_{\lambda,L}(X_{n}+r_{n}X)}{r_{n}}, where rn=|XnX0|\displaystyle r_{n}=|X_{n}-X_{0}| and XBR(0)\displaystyle X\in B_{R}(0) for any R>0\displaystyle R>0. By virtue of the results of Subsection 3.2, there exists a blow-up limit ϕ0\displaystyle\phi_{0}, such that

ϕ0(x,y)=0for any y0, and Δϕ0=0in {ϕ0>0}.\phi_{0}(x,y)=0\ \ \text{for any $\displaystyle y\geq 0$, and }\ \Delta\phi_{0}=0\ \ \text{in $\displaystyle\{\phi_{0}>0\}$}. (4.36)

It is easy to check that B2(0){ϕn>0}\displaystyle B_{2}(0)\cap\partial\{\phi_{n}>0\}\neq\varnothing, we next claim that

ϕ0(X)=λmin{yy1,0}in 2 for some y10.\phi_{0}(X)=-\lambda\min\{y-y_{1},0\}\ \ \text{in $\displaystyle\mathbb{R}^{2}$ for some $\displaystyle y_{1}\leq 0$}. (4.37)

Consider the complex z\displaystyle z-plane with z=x+iy\displaystyle z=x+iy. Let l\displaystyle l be a straight line in z\displaystyle z-plane with the direction (0,1)\displaystyle(0,1), and passing through the origin. Set

𝒟λ={z|z|λ}and𝒞λ=𝒟λ.\mathcal{D}_{\lambda}=\{z\mid|z|\leq\lambda\}\ \ \text{and}\ \ \mathcal{C}_{\lambda}=\partial\mathcal{D}_{\lambda}.

Since the blow-up limit ϕ0\displaystyle\phi_{0} is still a harmonic function, we can use the similar arguments in the proof of Lemma 6.2 in [7] to show that

lim supϕ(X)>0,XX0dist(ψλ,L(X),𝒟λl)=0.\limsup_{\phi(X)>0,X\rightarrow X_{0}}\text{dist}(\nabla\psi_{\lambda,L}(X),\mathcal{D}_{\lambda}\cup l)=0. (4.38)

Next, we will show that

|ϕ0(X)|λ.|\nabla\phi_{0}(X)|\leq\lambda. (4.39)

Suppose not, it follows from (4.38) that

ϕ0(X0)l\displaystyle\nabla\phi_{0}(X_{0})\in l and |ϕ0(X0)|>λ\displaystyle|\nabla\phi_{0}(X_{0})|>\lambda for some X0{ϕ0>0}\displaystyle X_{0}\in\{\phi_{0}>0\}. (4.40)

For the harmonic function ϕ0\displaystyle\phi_{0}, one has

if ϕ0\displaystyle\nabla\phi_{0}\neqconst in E\displaystyle E, then Xϕ0(X)\displaystyle X\rightarrow\nabla\phi_{0}(X) is an open mapping in E\displaystyle E

for any compact subsect E\displaystyle E of {ϕ0>0}\displaystyle\{\phi_{0}>0\}. This together with (4.40) implies that

ϕ0=const in {ϕ0>0}.\text{$\displaystyle\nabla\phi_{0}=$const in $\displaystyle\{\phi_{0}>0\}$}. (4.41)

Since ϕ0(X)=0\displaystyle\phi_{0}(X)=0 for any y0\displaystyle y\geq 0, it follows from (4.41) that ϕ0\displaystyle\phi_{0} is linear in E\displaystyle E respect to y\displaystyle y, and E\displaystyle\partial E is a straight line of the form {y=y1}\displaystyle\{y=y_{1}\} with y10\displaystyle y_{1}\leq 0. Thus, there exists a constant γ>λ\displaystyle\gamma>\lambda, such that

ϕ0(X)=γ(yy1)in{y<y1},y10.\phi_{0}(X)=-\gamma(y-y_{1})\ \ \ \text{in}\ \ \{y<y_{1}\},\ \ \ y_{1}\leq 0.

We next consider the following two cases for y1\displaystyle y_{1}.

Case 1. y1<0\displaystyle y_{1}<0. In view of Lemma 3.4, ϕ0\displaystyle\phi_{0} is a local minimizer in {y<0}\displaystyle\{y<0\} and {ϕ0>0}{y<0}\displaystyle\partial\{\phi_{0}>0\}\subset\{y<0\}, and Theorem 2.5 in [2] gives that |ϕ0|=λ=γ\displaystyle|\nabla\phi_{0}|=\lambda=\gamma, which contradicts to γ>λ\displaystyle\gamma>\lambda.

Case 2. y1=0\displaystyle y_{1}=0. Since ϕnϕ0\displaystyle\phi_{n}\rightarrow\phi_{0} uniformly in any compact subset of 2\displaystyle\mathbb{R}^{2}, for any R>0\displaystyle R>0 and ε>0\displaystyle\varepsilon>0, there exists a N=N(ε,R)\displaystyle N=N(\varepsilon,R), such that

ϕn>0inBR(0){y<ε},forn>N.\phi_{n}>0\ \ \text{in}\ \ B_{R}(0)\cap\{y<-\varepsilon\},\ \ \text{for}\ \ n>N.

Noticing that there are free boundary points Zn=(sn,tn)\displaystyle Z_{n}=(s_{n},t_{n}) of ϕn\displaystyle\phi_{n} in B2(0)\displaystyle B_{2}(0), the assumption y1=0\displaystyle y_{1}=0 implies that tn0\displaystyle t_{n}\rightarrow 0. Without loss of generality, we assume that sn0\displaystyle s_{n}\rightarrow 0. Therefore,

|sn|<ε4and|tn|<ε4for sufficiently large n.|s_{n}|<\frac{\varepsilon}{4}\ \ \ \text{and}\ \ \ |t_{n}|<\frac{\varepsilon}{4}\ \ \ \text{for sufficiently large $\displaystyle n$}.

Introduce a function hδ,n(x)\displaystyle h_{\delta,n}(x) as follows

hδ,n(x)=2ε+δηn(x),for someδ>0,h_{\delta,n}(x)=-2\varepsilon+\delta\eta_{n}(x),\ \ \ \text{for some}\ \ \delta>0,

where

ηn(x)={e9|xτn|219|xτn|2for|xτn|<13,0for|xτn|13,\eta_{n}(x)=\left\{\begin{array}[]{ll}e^{-\frac{9|x-\tau_{n}|^{2}}{1-9|x-\tau_{n}|^{2}}}&\text{for}\ \ |x-\tau_{n}|<\frac{1}{3},\\ 0&\text{for}\ \ |x-\tau_{n}|\geq\frac{1}{3},\end{array}\right.

for a sequence τn0\displaystyle\tau_{n}\rightarrow 0. Denote the domain Eδ,n=B1(0){y<hδ,n(x)}\displaystyle E_{\delta,n}=B_{1}(0)\cap\{y<h_{\delta,n}(x)\}, and choose a largest δ=δn\displaystyle\delta=\delta_{n}, such that ϕn>0\displaystyle\phi_{n}>0 in Eδn,n\displaystyle E_{\delta_{n},n} and B1(0)Eδn,n\displaystyle B_{1}(0)\cap\partial E_{\delta_{n},n} contains a free boundary point Z~n=(s~n,t~n)\displaystyle\tilde{Z}_{n}=(\tilde{s}_{n},\tilde{t}_{n}) of ϕn\displaystyle\phi_{n}. The τn(εn,εn)\displaystyle\tau_{n}\in\left(-\frac{\varepsilon}{n},\frac{\varepsilon}{n}\right) can be chosen such that δn<2ε\displaystyle\delta_{n}<2\varepsilon and y~n=hδn,n(x~n)\displaystyle\tilde{y}_{n}=h_{\delta_{n},n}(\tilde{x}_{n}).

Let ωn\displaystyle\omega_{n} be the solution of the following Dirichlet problem

{Δωn+rnf(rnωn)=0inEδn,n,ωn=0onEδn,nB12(0),ωn=ζϕnonEδn,n(B1(0)B12(0)),ωn=ϕnonEδn,nB1(0),\left\{\begin{array}[]{ll}&\Delta\omega_{n}+r_{n}f(r_{n}\omega_{n})=0\ \ \ \ \text{in}\ E_{\delta_{n},n},\\ &\omega_{n}=0\ \ \ \text{on}\ \partial E_{\delta_{n},n}\cap B_{\frac{1}{2}}(0),\ \omega_{n}=\zeta\phi_{n}\ \text{on}\ \partial E_{\delta_{n},n}\cap(B_{1}(0)\setminus B_{\frac{1}{2}}(0)),\\ &\omega_{n}=\phi_{n}\ \ \ \text{on}\ \partial E_{\delta_{n},n}\cap\partial B_{1}(0),\end{array}\right.

where ζ(X)=min{max{2|X|1,0},1}\displaystyle\zeta(X)=\min\left\{\max\left\{2|X|-1,0\right\},1\right\}. Obviously, ϕnωn\displaystyle\phi_{n}\geq\omega_{n} on Eδn,n\displaystyle\partial E_{\delta_{n},n} and Δϕn+rnf(rnϕn)=0\displaystyle\Delta\phi_{n}+r_{n}f(r_{n}\phi_{n})=0 in Eδn,n\displaystyle E_{\delta_{n},n}, the strong maximum principle gives that ωn<ϕn\displaystyle\omega_{n}<\phi_{n} in Eδn,n\displaystyle E_{\delta_{n},n}. Thanks to Hopf’s Lemma, one has

|ϕnνn|>|ωnνn|atZ~n,\left|\frac{\partial\phi_{n}}{\partial\nu_{n}}\right|>\left|\frac{\partial\omega_{n}}{\partial\nu_{n}}\right|\ \ \ \text{at}\ \ \tilde{Z}_{n}, (4.42)

where νn\displaystyle\nu_{n} is the inner normal vector to Eδn,n\displaystyle E_{\delta_{n},n} at Z~n\displaystyle\tilde{Z}_{n}.

The definition of hδn,n\displaystyle h_{\delta_{n},n} gives that hδn,nC2,αCε\displaystyle\|h_{\delta_{n},n}\|_{C^{2,\alpha}}\leq C\varepsilon, where the constant C\displaystyle C is independent of ε\displaystyle\varepsilon and n\displaystyle n. Applying the regularity theory for the semilinear elliptic equation yields that

|ωnϕ0|C1ε+C2rnatZ~n,|\nabla\omega_{n}-\nabla\phi_{0}|\leq C_{1}\varepsilon+C_{2}r_{n}\ \ \ \text{at}\ \ \ \tilde{Z}_{n},

where the constants C1,C2\displaystyle C_{1},C_{2} are independent of ε\displaystyle\varepsilon and n\displaystyle n, which together with (4.42) implies that

λ=|ϕn(Z~n)|>|ωn(Z~n)||ϕ0(Z~n)|C1εC2rn=γC1εC2rn.\lambda=|\phi_{n}(\tilde{Z}_{n})|>|\nabla\omega_{n}(\tilde{Z}_{n})|\geq|\nabla\phi_{0}(\tilde{Z}_{n})|-C_{1}\varepsilon-C_{2}r_{n}=\gamma-C_{1}\varepsilon-C_{2}r_{n}.

This leads a contradiction to the assumption γ>λ\displaystyle\gamma>\lambda, provided that ε\displaystyle\varepsilon is small enough and n\displaystyle n is sufficiently large.

Hence, we complete the proof of (4.39).

Next, we will show the claim (4.37). Since ϕ0\displaystyle\phi_{0} is harmonic in {ϕ0>0}\displaystyle\{\phi_{0}>0\}, it follows from Lemma 7.2 in [7] that

qν+κq=0on the free boundary of ϕ0,\frac{\partial q}{\partial\nu}+\kappa q=0\ \ \text{on the free boundary of $\displaystyle\phi_{0}$}, (4.43)

where q=|ϕ0|\displaystyle q=|\nabla\phi_{0}|, ν\displaystyle\nu is the outer normal vector and the curvature κ>0\displaystyle\kappa>0 if the streamline is concave to the fluid. Due to |ϕ0|λ\displaystyle|\nabla\phi_{0}|\leq\lambda and |ϕ0|=λ\displaystyle|\nabla\phi_{0}|=\lambda on the free boundary {ϕ0>0}\displaystyle\partial\{\phi_{0}>0\}, it follows from (4.43) that {ϕ0>0}\displaystyle\partial\{\phi_{0}>0\} is convex to the domain {ϕ0>0}\displaystyle\{\phi_{0}>0\}, and the set {ϕ0=0}\displaystyle\{\phi_{0}=0\} is convex.

If ϕ0>0\displaystyle\phi_{0}>0 in {y<0}\displaystyle\{y<0\}. Denote ω=xϕ0\displaystyle\omega=\partial_{x}\phi_{0}, it is easy to see that ω\displaystyle\omega is harmonic in {ϕ0>0}\displaystyle\{\phi_{0}>0\}. The fact ϕ0=0\displaystyle\phi_{0}=0 on {y=0}\displaystyle\{y=0\} implies that ω=0\displaystyle\omega=0. In view of |ϕ0|λ\displaystyle|\nabla\phi_{0}|\leq\lambda, one has

limr+m(r)r=0where r=|X| and m(r)=max|X|=r|ω(X)|.\lim_{r\rightarrow+\infty}\frac{m(r)}{r}=0\ \ \ \text{where $\displaystyle r=|X|$ and $\displaystyle m(r)=\max_{|X|=r}|\omega(X)|$}.

Therefore, thanks to Phragmén-Lindelöf Theorem (see also Theorem 1.1 in [25]), we have

ω=xϕ0=0in{y<0},\omega=\partial_{x}\phi_{0}=0\ \ \text{in}\ \ \{y<0\},

which implies that ϕ0(X)\displaystyle\phi_{0}(X) is a function of y\displaystyle y and

ϕ0(y)=γyin{y<0}.\phi_{0}(y)=-\gamma y\ \ \ \text{in}\ \ \{y<0\}.

It follows from Theorem 2.5 in [2] that γ=λ\displaystyle\gamma=\lambda, and thus the claim (4.37) holds.

If there exists a free boundary point X~=(x~,y~)\displaystyle\tilde{X}=(\tilde{x},\tilde{y}) of ϕ0\displaystyle\phi_{0} with y~<0\displaystyle\tilde{y}<0, denote Γ0\displaystyle\Gamma_{0} the maximal free boundary arc containing X~\displaystyle\tilde{X}. We next consider the following two cases.

Case 1. The free boundary Γ0\displaystyle\Gamma_{0} is a straight line l\displaystyle l. The fact ϕ0=0\displaystyle\phi_{0}=0 in {y0}\displaystyle\{y\geq 0\} implies that the straight line l\displaystyle l must be parallel to the x\displaystyle x-axis. Moreover, ϕ0>0\displaystyle\phi_{0}>0 below l\displaystyle l, due to that ϕ0(X)\displaystyle\phi_{0}(X) is monotone decreasing with respect to y\displaystyle y. Thus one has

l:{y=y1}andϕ0(y)=λmin{yy1,0}.l:\{y=y_{1}\}\ \ \ \text{and}\ \ \ \phi_{0}(y)=-\lambda\min\{y-y_{1},0\}.

Case 2. The free boundary Γ0\displaystyle\Gamma_{0} is not a straight line. We next consider the following two subcases (see Figure 11).

Refer to caption
Figure 11. Case 2

Subcase 2.1. ϕ0>0\displaystyle\phi_{0}>0 above Γ0\displaystyle\Gamma_{0}. Since ϕ0(X)\displaystyle\phi_{0}(X) is monotone decreasing with respect to y\displaystyle y, we can rule out this subcase.

Subcase 2.2. ϕ0>0\displaystyle\phi_{0}>0 below Γ0\displaystyle\Gamma_{0}. The convexity of the free boundary Γ0\displaystyle\Gamma_{0} implies that Γ0\displaystyle\Gamma_{0} must intersect the x\displaystyle x-axis, denote P\displaystyle P the intersection point. Moreover, the convexity of Γ0\displaystyle\Gamma_{0} and monotonicity of ϕ0\displaystyle\phi_{0} with respect to y\displaystyle y imply that the angle θ\displaystyle\theta between Γ\displaystyle\Gamma and x\displaystyle x-axis at point P\displaystyle P lies in [π2,π)\displaystyle\left[\frac{\pi}{2},\pi\right). Applying the similar arguments in the proof of (2.30) in [17], one has

ϕ0(X)C|XP|2πθ+π+εnear the intersection point P,\phi_{0}(X)\leq C|X-P|^{\frac{2\pi}{\theta+\pi+\varepsilon}}\ \ \ \text{near the intersection point $\displaystyle P$}, (4.44)

for small ε(0,πθ4)\displaystyle\varepsilon\in\left(0,\frac{\pi-\theta}{4}\right). Denote r=|XP|\displaystyle r=|X-P|, using the non-degeneracy Lemma 3.4 and Remark 3.5 in [3], we have

1rBr(P)ϕ0𝑑Scλ.\frac{1}{r}\fint_{\partial B_{r}(P)}\phi_{0}dS\geq c_{*}\lambda. (4.45)

On the other hand, it follows from (4.44) and (4.45) that

cλ1rBr(P)ϕ0𝑑SCr2πθ+π+εr=Crπθεπ+θ+ε,c_{*}\lambda\leq\frac{1}{r}\fint_{\partial B_{r}(P)}\phi_{0}dS\leq C\frac{r^{\frac{2\pi}{\theta+\pi+\varepsilon}}}{r}=Cr^{\frac{\pi-\theta-\varepsilon}{\pi+\theta+\varepsilon}},

which leads a contradiction, provided that r\displaystyle r is small enough.

Hence, the proof of the claim (4.37) is done.

With the aid of the claim (4.37), we will show that

ψλ,L(X)λ(0,1)asXX0,XΩL{ψλ,L<Q}.\nabla\psi_{\lambda,L}(X)\rightarrow\lambda(0,1)\ \ \text{as}\ \ \ X\rightarrow X_{0},\ \ X\in\Omega_{L}\cap\{\psi_{\lambda,L}<Q\}. (4.46)

For any sequence of points Xn=(xn,yn)ΩL{ψλ,L<Q}\displaystyle X_{n}=(x_{n},y_{n})\in\Omega_{L}\cap\{\psi_{\lambda,L}<Q\} with XnX0\displaystyle X_{n}\rightarrow X_{0}, let X~n=(x~n,y~n)DL\displaystyle\tilde{X}_{n}=(\tilde{x}_{n},\tilde{y}_{n})\in\partial D_{L} be the nearest point to Xn\displaystyle X_{n}. Denote

rn=|XnX~n|anddn=dist(Xn,Γλ,L).r_{n}=|X_{n}-\tilde{X}_{n}|\ \ \ \text{and}\ \ \ d_{n}=\text{dist}(X_{n},\Gamma_{\lambda,L}).

Next, we consider the following two cases.

Case 1. There exists a constant C0>1\displaystyle C_{0}>1, such that rn<C0dn\displaystyle r_{n}<C_{0}d_{n}. Define a blow-up sequence

ϕn(X)=Qψλ,L(X~n+C0dnX)C0dnandZn=XnX~nC0dn.\phi_{n}(X)=\frac{Q-\psi_{\lambda,L}(\tilde{X}_{n}+C_{0}d_{n}X)}{C_{0}d_{n}}\ \ \ \text{and}\ \ \ Z_{n}=\frac{X_{n}-\tilde{X}_{n}}{C_{0}d_{n}}.

It is easy to check that B2(0)\displaystyle B_{2}(0) contains some free boundary points of ϕn\displaystyle\phi_{n} and B2c0(Zn){ϕn>0}=\displaystyle B_{2c_{0}}(Z_{n})\cap\partial\{\phi_{n}>0\}=\varnothing, which does not intersect the free boundary of ϕn\displaystyle\phi_{n}, where c0=14C0\displaystyle c_{0}=\frac{1}{4C_{0}}. Moreover,

Δϕn+C0dnf(C0xnϕn)=0in B2c0(Zn){ϕn>0}.\Delta\phi_{n}+C_{0}d_{n}f(C_{0}x_{n}\phi_{n})=0\ \text{in $\displaystyle B_{2c_{0}}(Z_{n})\cap\{\phi_{n}>0\}$}. (4.47)

There exist a blow-up limit ϕ0\displaystyle\phi_{0} and a point Z0\displaystyle Z_{0} with |Z0|1\displaystyle|Z_{0}|\leq 1, such that

ϕnϕ0in any compact subsets of 2 and ZnZ0.\phi_{n}\rightarrow\phi_{0}\ \ \text{in any compact subsets of $\displaystyle\mathbb{R}^{2}$ and $\displaystyle Z_{n}\rightarrow Z_{0}$.}

Due to the fact that B2c0(Z0)\displaystyle B_{2c_{0}}(Z_{0}) does not intersect the free boundary of ϕ0\displaystyle\phi_{0} and DL\displaystyle\partial D_{L} is C2\displaystyle C^{2}-smooth near X0\displaystyle X_{0}, one has

ϕnC1,α in Bc0(Zn){ϕn>0} uniformly with respect to n.\text{$\displaystyle\phi_{n}\in C^{1,\alpha}$ in $\displaystyle B_{c_{0}}(Z_{n})\cap\{\phi_{n}>0\}$ uniformly with respect to $\displaystyle n$}.

Then we have

ψλ,L(Xn)=ϕn(Zn)ϕ0(Z0), as n,\text{$\displaystyle-\nabla\psi_{\lambda,L}(X_{n})=\nabla\phi_{n}(Z_{n})\rightarrow\nabla\phi_{0}(Z_{0})$, as $\displaystyle n\rightarrow\infty$}, (4.48)

which together with (4.37) gives that (4.46) holds.

Case 2. rndn+\displaystyle\frac{r_{n}}{d_{n}}\rightarrow+\infty. Define a blow-up sequence

ϕn(X)=Qψλ,L(X~n+rnX)rnandYn=XnX~nrn.\phi_{n}(X)=\frac{Q-\psi_{\lambda,L}(\tilde{X}_{n}+r_{n}X)}{r_{n}}\ \ \ \text{and}\ \ \ Y_{n}=\frac{X_{n}-\tilde{X}_{n}}{r_{n}}.

Obviously, B2(0)\displaystyle B_{2}(0) contains the free boundary points of ϕn\displaystyle\phi_{n} for sufficiently large n\displaystyle n. Then there exist a blow-up limit ϕ0\displaystyle\phi_{0} and a point Y0\displaystyle Y_{0} with |Y0|=1\displaystyle|Y_{0}|=1, such that

ϕnϕ0=λmin{yy1,0}in any compact subsets of 2 and YnY0,\phi_{n}\rightarrow\phi_{0}=-\lambda\min\{y-y_{1},0\}\ \ \text{in any compact subsets of $\displaystyle\mathbb{R}^{2}$ and $\displaystyle Y_{n}\rightarrow Y_{0}$,}

where y10\displaystyle y_{1}\leq 0. Since the free boundary {ϕn>0}{ϕ0>0}\displaystyle\partial\{\phi_{n}>0\}\rightarrow\partial\{\phi_{0}>0\} locally in the Hausdorff distance (see Lemma 3.4) and {ϕ0>0}\displaystyle\partial\{\phi_{0}>0\} is a straight line, we can conclude that the free boundary of ϕn\displaystyle\phi_{n} satisfies the flatness condition at (0,y1)\displaystyle(0,y_{1}) in the direction (0,1)\displaystyle(0,1), for sufficiently large n\displaystyle n. By virtue of Theorem 3.15, one has the free boundary of ϕn\displaystyle\phi_{n} is C1,α\displaystyle C^{1,\alpha} (0<α<1)\displaystyle(0<\alpha<1) uniformly in n\displaystyle n. The elliptic regularity gives that ϕn\displaystyle\phi_{n} is uniformly C1,α\displaystyle C^{1,\alpha} in BR(Yn){ϕn>0}\displaystyle B_{R}(Y_{n})\cap\{\phi_{n}>0\} for some R>0\displaystyle R>0. Thus we have

ψλ,L(Xn)=ϕn(Yn)ϕ0(Y0)=λ(0,1), as n.\text{$\displaystyle-\nabla\psi_{\lambda,L}(X_{n})=\nabla\phi_{n}(Y_{n})\rightarrow\nabla\phi_{0}(Y_{0})=-\lambda(0,1)$, as $\displaystyle n\rightarrow\infty$}.

Since the blow-up limit ϕ0\displaystyle\phi_{0} is a harmonic function, recalling the fact (4.46), along the similar arguments on the smooth fit condition (see Section 11 in Chapter 3 in [23]), we can show that

kλ,L(L2)=0.k^{\prime}_{\lambda,L}\left(\frac{L}{2}\right)=0.

Hence, we complete the proof of the proposition. ∎

4.6. The continuous fit condition for free boundary

In this subsection, we will check the continuous fit condition of the free boundary Γλ,L\displaystyle\Gamma_{\lambda,L} at A\displaystyle A, namely, for any L>H¯=maxx0g(x)\displaystyle L>\bar{H}=\max_{x\leq 0}g(x), there exists a λLλ0\displaystyle\lambda_{L}\geq\lambda_{0}, such that

kλL,L(0)=g(0)=a.k_{\lambda_{L},L}(0)=g(0)=a. (4.49)

To obtain the continuous fit condition (4.49), we first establish the continuous dependence of ψλ,L\displaystyle\psi_{\lambda,L} and kλ,L(0)\displaystyle k_{\lambda,L}(0) with respect to the parameter λ\displaystyle\lambda.

Lemma 4.11.

If λnλ\displaystyle\lambda_{n}\rightarrow\lambda with λnλ0\displaystyle\lambda_{n}\geq\lambda_{0}, then

ψλn,Lψλ,Lweakly   inH1(ΩL)and uniformly inΩL,\psi_{\lambda_{n},L}\rightarrow\psi_{\lambda,L}\ \ \text{weakly~~ in}~~H^{1}(\Omega_{L})~~\text{and uniformly in}~\Omega_{L},

and

kλn,L(0)kλ,L(0)forkλ,L(0)<L.k_{\lambda_{n},L}(0)\rightarrow k_{\lambda,L}(0)\ \ \text{for}~~k_{\lambda,L}(0)<L. (4.50)
Proof.

It follows from the similar arguments in the proof of Lemma 3.4, there exists a subsequence {ψλn,L}\displaystyle\{\psi_{\lambda_{n},L}\} and a ψ0C0,1(ΩL)\displaystyle\psi_{0}\in C^{0,1}(\Omega_{L}), such that

ψλn,L(X)ψ0(X)uniformly in ΩL.\psi_{\lambda_{n},L}(X)\rightarrow\psi_{0}(X)\ \text{uniformly in $\displaystyle\Omega_{L}$}.

Furthermore, ψ0\displaystyle\psi_{0} is a minimizer to the variational problem (Pλ,L)\displaystyle(P_{\lambda,L}). The uniqueness of the minimizer to the variational problem (Pλ,L)\displaystyle(P_{\lambda,L}) gives that ψ0=ψλ,L\displaystyle\psi_{0}=\psi_{\lambda,L}.

Suppose that the assertion (4.50) is not true, then there exists a sequence {kλn,L(0)}\displaystyle\{k_{\lambda_{n},L}(0)\}, such that

kλn,L(0)kλ,L(0)+δas n+,δ0.k_{\lambda_{n},L}(0)\rightarrow k_{\lambda,L}(0)+\delta\ \ \text{as $\displaystyle n\rightarrow+\infty$},\ \ \delta\neq 0.

We consider the following three cases and derive a contradiction.

Case 1. δ<0\displaystyle\delta<0. The monotonicity of ψλ,L(x,y)\displaystyle\psi_{\lambda,L}(x,y) with respect to y\displaystyle y gives that kλ,L(0)+δa\displaystyle k_{\lambda,L}(0)+\delta\geq a.

Next, we claim that

ψλ,L(0+0,y)x=λonIε,\frac{\partial\psi_{\lambda,L}(0+0,y)}{\partial x}=-\lambda\ \ \ \text{on}\ \ I_{\varepsilon}, (4.51)

where Iε={(0,y)kλ,L(0)4ε<y<kλ,L(0)2ε}\displaystyle I_{\varepsilon}=\{(0,y)\mid k_{\lambda,L}(0)-4\varepsilon<y<k_{\lambda,L}(0)-2\varepsilon\} with ε=δ6\displaystyle\varepsilon=-\frac{\delta}{6}.

It follows from the statement (4) in Proposition 4.4 that

ψλ,L(0+0,y)x=|ψλ,L|λonIε.-\frac{\partial\psi_{\lambda,L}(0+0,y)}{\partial x}=|\nabla\psi_{\lambda,L}|\geq\lambda\ \ \text{on}\ \ I_{\varepsilon}. (4.52)

To obtain the claim (4.51), it suffices to show that

ψλ,L(0+0,y)xλonIε.\frac{\partial\psi_{\lambda,L}(0+0,y)}{\partial x}\geq-\lambda\ \ \ \text{on}\ \ I_{\varepsilon}. (4.53)

Denote ϕn=Qψλn,L\displaystyle\phi_{n}=Q-\psi_{\lambda_{n},L}, ϕ=Qψλ,L\displaystyle\phi=Q-\psi_{\lambda,L} and

Uε,τ={(x,y)τ<x<τ,kλ,L(0)5ε<y<kλ,L(0)ε}U_{\varepsilon,\tau}=\{(x,y)\mid-\tau<x<\tau,k_{\lambda,L}(0)-5\varepsilon<y<k_{\lambda,L}(0)-\varepsilon\}

for small τ>0\displaystyle\tau>0, Proposition 4.4 gives that the free boundary Uε,τ{ϕn>0}\displaystyle U_{\varepsilon,\tau}\cap\partial\{\phi_{n}>0\} is C3,α\displaystyle C^{3,\alpha}, then we have

Uε{ϕn>0}Uε,τ{ϕ>0}\displaystyle U_{\varepsilon}\cap\partial\{\phi_{n}>0\}\rightarrow U_{\varepsilon,\tau}\cap\partial\{\phi>0\} in C1,α\displaystyle C^{1,\alpha} for some α(0,1)\displaystyle\alpha\in(0,1).

For any fixed X1=(0,y1)Uε,τ{ϕ>0}\displaystyle X_{1}=(0,y_{1})\in U_{\varepsilon,\tau}\cap\partial\{\phi>0\}, then there exists a sequence XnUε\displaystyle X_{n}\in U_{\varepsilon} with ϕn(Xn)=0\displaystyle\phi_{n}(X_{n})=0, such that XnX1\displaystyle X_{n}\rightarrow X_{1} as n+\displaystyle n\rightarrow+\infty. In fact, suppose that there exists a small r>0\displaystyle r>0, such that ϕn>0\displaystyle\phi_{n}>0 and Δϕn+f(ϕn)=0\displaystyle\Delta\phi_{n}+f(\phi_{n})=0 in Br(X1)\displaystyle B_{r}(X_{1}), which gives that Δϕ+f(ϕ)=0\displaystyle\Delta\phi+f(\phi)=0 in Br(X1)\displaystyle B_{r}(X_{1}) and ϕ(X1)=0\displaystyle\phi(X_{1})=0. The maximum principle implies that ϕ0\displaystyle\phi\equiv 0 in Br(X1)\displaystyle B_{r}(X_{1}), which contradicts to X1Uε,τ{ϕ>0}\displaystyle X_{1}\in U_{\varepsilon,\tau}\cap\partial\{\phi>0\}.

Take a small r>0\displaystyle r>0 with EnBr(X1)Uε,τ\displaystyle E_{n}\subset B_{r}(X_{1})\subset U_{\varepsilon,\tau}, such that

En=Br(X1){x>εn},ϕn>0inEnandXnEn,E_{n}=B_{r}(X_{1})\cap\{x>\varepsilon_{n}\},\ \phi_{n}>0\ \text{in}\ E_{n}\ \text{and}\ X_{n}\notin E_{n}, (4.54)

where εn0\displaystyle\varepsilon_{n}\downarrow 0 as n+\displaystyle n\rightarrow+\infty. Define a function hs,n(y)\displaystyle h_{s,n}(y) as follows

hs,n(y)=εnsη(2(yy1)r),s>0,h_{s,n}(y)=\varepsilon_{n}-s\eta\left(\frac{2(y-y_{1})}{r}\right),\ \ s>0,

where

η(y)={ey21y2for|y|<10for|y|1.\eta(y)=\left\{\begin{array}[]{ll}e^{-\frac{y^{2}}{1-y^{2}}}&\text{for}\ \ |y|<1\\ 0&\text{for}\ |y|\geq 1.\end{array}\right. (4.55)

Denote the domain Es,n=Br(X1){x>hs,n(y)}\displaystyle E_{s,n}=B_{r}(X_{1})\cap\{x>h_{s,n}(y)\}. It is easy to check that E0,n=En\displaystyle E_{0,n}=E_{n}. Let s=snεn\displaystyle s=s_{n}\leq\varepsilon_{n} to be the largest one, such that ϕn>0\displaystyle\phi_{n}>0 in Esn,n\displaystyle E_{s_{n},n}, and X~n=(x~n,y~n)(Br(X1)Esn,n)Γλn,L\displaystyle\tilde{X}_{n}=(\tilde{x}_{n},\tilde{y}_{n})\in(B_{r}(X_{1})\cap\partial E_{s_{n},n})\cap\Gamma_{\lambda_{n},L}. Furthermore,

x~n=hsn,n(y~n)andsn0asn+.\tilde{x}_{n}=h_{s_{n},n}(\tilde{y}_{n})\ \ \text{and}\ \ s_{n}\rightarrow 0\ \ \text{as}\ \ n\rightarrow+\infty.

Let φn\displaystyle\varphi_{n} be the solution of the following Dirichlet problem

{Δφn+f(φn)=0inEsn,n,φn=0onEsn,nBr2(X1),φn=ζϕnonEsn,n(Br(X1)Br2(X1)),φn=ϕnonEsn,nBr(X1),\left\{\begin{array}[]{ll}&\Delta\varphi_{n}+f(\varphi_{n})=0\ \ \text{in}\ E_{s_{n},n},\\ &\varphi_{n}=0\ \text{on}\ \partial E_{s_{n},n}\cap B_{\frac{r}{2}}(X_{1}),\ \varphi_{n}=\zeta\phi_{n}\ \text{on}\ \partial E_{s_{n},n}\cap(B_{r}(X_{1})\setminus B_{\frac{r}{2}}(X_{1})),\\ &\varphi_{n}=\phi_{n}\ \text{on}\ \partial E_{s_{n},n}\cap\partial B_{r}(X_{1}),\end{array}\right.

where ζ(X)=min{max{2|XX1|rr,0},1}\displaystyle\zeta(X)=\min\left\{\max\left\{\frac{2|X-X_{1}|-r}{r},0\right\},1\right\}. It is clear that φnϕn\displaystyle\varphi_{n}\leq\phi_{n} on Esn,n\displaystyle\partial E_{s_{n},n}, which gives that

λn=ϕn(X~n)νnφn(X~n)νn,\lambda_{n}=\frac{\partial\phi_{n}(\tilde{X}_{n})}{\partial\nu_{n}}\geq\frac{\partial\varphi_{n}(\tilde{X}_{n})}{\partial\nu_{n}}, (4.56)

where νn\displaystyle\nu_{n} is the inner normal vector to Esn,n\displaystyle\partial E_{s_{n},n} at X~n\displaystyle\tilde{X}_{n}.

Taking εn0\displaystyle\varepsilon_{n}\rightarrow 0 implies that

hsn,n(y)0inC1,βasn+.h_{s_{n},n}(y)\rightarrow 0\ \ \text{in}\ \ C^{1,\beta}\ \ \text{as}\ \ n\rightarrow+\infty.

Thanks to the standard estimates of the solutions of the semilinear elliptic equation, we conclude that φn\displaystyle\varphi_{n} in Esn,nBr2(X1)\displaystyle E_{s_{n},n}\cap B_{\frac{r}{2}}(X_{1}) converges to ϕ\displaystyle\phi in {ϕ>0}Br2(X1)\displaystyle\{\phi>0\}\cap B_{\frac{r}{2}}(X_{1}) in C1,β\displaystyle C^{1,\beta}-sense. Denote X~nX~{ϕ>0}Br2(X1).\displaystyle\tilde{X}_{n}\rightarrow\tilde{X}\in\partial\{\phi>0\}\cap B_{\frac{r}{2}}(X_{1}). It follows from (4.56) that

φn(X~n)νϕ(X~)νandϕ(X~)νλ.\frac{\partial\varphi_{n}(\tilde{X}_{n})}{\partial\nu}\rightarrow\frac{\partial\phi(\tilde{X})}{\partial\nu}\ \ \text{and}\ \ \frac{\partial\phi(\tilde{X})}{\partial\nu}\leq\lambda. (4.57)

Taking r0\displaystyle r\rightarrow 0 yields that X~X1\displaystyle\tilde{X}\rightarrow X_{1} and

λϕ(X~)νϕ(X1)ν=ψλ,L(X1)x,X1Uε,τ{ϕ>0},\lambda\geq\frac{\partial\phi(\tilde{X})}{\partial\nu}\rightarrow\frac{\partial\phi(X_{1})}{\partial\nu}=-\frac{\partial\psi_{\lambda,L}(X_{1})}{\partial x},\ \ X_{1}\in U_{\varepsilon,\tau}\cap\partial\{\phi>0\},

which together with (4.53) gives that the claim (4.51) holds.

With the aid of the fact (4.51), we can obtain a contradiction by using the similar arguments in the proof of Lemma 4.8.

Case 2. δ>0\displaystyle\delta>0 and kλ,L(0)<a\displaystyle k_{\lambda,L}(0)<a. Similar to the proof of the claim (4.51) in Case 1, one has

ψλ,L(00,y)x=λonIε,\frac{\partial\psi_{\lambda,L}(0-0,y)}{\partial x}=\lambda\ \ \ \text{on}\ \ I_{\varepsilon}, (4.58)

where Iε={(0,y)kλ,L(0)+ε<y<kλ,L(0)+2ε}\displaystyle I_{\varepsilon}=\{(0,y)\mid k_{\lambda,L}(0)+\varepsilon<y<k_{\lambda,L}(0)+2\varepsilon\} with ε=min{δ,akλ,L(0)}3\displaystyle\varepsilon=\frac{\min\{\delta,a-k_{\lambda,L}(0)\}}{3}, and we can obtain a contradiction.

Case 3. δ>0\displaystyle\delta>0 and kλ,L(0)a\displaystyle k_{\lambda,L}(0)\geq a. Since ψλ,L=Q\displaystyle\psi_{\lambda,L}=Q on I0,L\displaystyle I_{0,L}, it follows from the statement (4) in Proposition 4.4 that

ψλn,L(0+0,y)xλnonIδ,\frac{\partial\psi_{\lambda_{n},L}(0+0,y)}{\partial x}\geq\lambda_{n}\ \ \text{on}\ \ I_{\delta},

for sufficiently large n\displaystyle n, where Iδ={(0,y)kλ,L(0)+δ4<y<kλ,L(0)+3δ4}\displaystyle I_{\delta}=\left\{(0,y)\mid k_{\lambda,L}(0)+\frac{\delta}{4}<y<k_{\lambda,L}(0)+\frac{3\delta}{4}\right\}.

Let En\displaystyle E_{n} be a domain bounded by x=0\displaystyle x=0, y=kλn,L(x)\displaystyle y=k_{\lambda_{n},L}(x), y=kλ,L(0)+δ4\displaystyle y=k_{\lambda,L}(0)+\frac{\delta}{4} and y=kλ,L(0)+3δ4\displaystyle y=k_{\lambda,L}(0)+\frac{3\delta}{4}. Moreover,

xn=max{xkλn,L(x)=kλ,L(0)+δ4}0asn+.x_{n}=\max\left\{x\mid k_{\lambda_{n},L}(x)=k_{\lambda,L}(0)+\frac{\delta}{4}\right\}\rightarrow 0\ \ \text{as}\ \ n\rightarrow+\infty.

Thanks to the non-oscillation Lemma 4.7 for ψλn,L\displaystyle\psi_{\lambda_{n},L} in En\displaystyle E_{n}, there exists a constant C\displaystyle C independent of n\displaystyle n, such that

0<δ2Cxn,0<\frac{\delta}{2}\leq Cx_{n},

which gives a contradiction for sufficiently large n\displaystyle n.

Next, we will study the relation between the initial point (0,kλ,L(0))\displaystyle(0,k_{\lambda,L}(0)) and λ\displaystyle\lambda for any L>H¯\displaystyle L>\bar{H}.

Lemma 4.12.

For any L>H¯\displaystyle L>\bar{H}, the initial point (0,kλ,L(0))\displaystyle(0,k_{\lambda,L}(0)) satisfies that
(1) there exists a constant C0>0\displaystyle C_{0}>0 (independent of L\displaystyle L), such that kλ,L(0)<a\displaystyle k_{\lambda,L}(0)<a for any λ>C0\displaystyle\lambda>C_{0}.
(2) there exists a c0λ0\displaystyle c_{0}\geq\lambda_{0} (independent of L\displaystyle L), such that kλ,L(0)a\displaystyle k_{\lambda,L}(0)\geq a for any λc0\displaystyle\lambda\leq c_{0}.

Proof.

(1) Suppose that there exist a free boundary point and a sufficiently large λ\displaystyle\lambda, such that kλ,L(0)a\displaystyle k_{\lambda,L}(0)\geq a. Then there exist a X0DL\displaystyle X_{0}\in D_{L} and a disc Br0(X0)\displaystyle B_{r_{0}}(X_{0}) with r0>0\displaystyle r_{0}>0 (independent of L\displaystyle L), such that Γλ,LBr02(X0)\displaystyle\Gamma_{\lambda,L}\cap B_{\frac{r_{0}}{2}}(X_{0})\neq\varnothing. It follows from the non-degeneracy Lemma 2.6 and Remark 2.3 that

Qr01r0Br0(X0)Qψλ,LdSc12λ,\frac{Q}{r_{0}}\geq\frac{1}{r_{0}}\fint_{\partial{B_{r_{0}}(X_{0})}}Q-\psi_{\lambda,L}dS\geq c_{\frac{1}{2}}^{*}\lambda,

which leads to a contradiction for sufficiently large λ\displaystyle\lambda.

Similarly, we can obtain a contradiction if there is no free boundary point, taking X0=(0,y0)\displaystyle X_{0}=(0,y_{0}) with y0(a,L)\displaystyle y_{0}\in(a,L) and using the non-degeneracy Lemma 2.5 for the boundary point (see Remark 2.3).

(2) Suppose not, then for any c0λ0\displaystyle c_{0}\geq\lambda_{0}, there exists a λ~[λ0,c0]\displaystyle\tilde{\lambda}\in[\lambda_{0},c_{0}], such that kλ~,L(0)<a\displaystyle k_{\tilde{\lambda},L}(0)<a. In view of Remark 4.4, we have

the asymptotic height h~=hλ~\displaystyle\tilde{h}=h_{\tilde{\lambda}} in downstream lies in [a,H]\displaystyle[a,H], (4.59)

provided that c0λ0\displaystyle c_{0}-\lambda_{0} is small. Let Ψ~ε(y)=Ψλ~(y+ε)\displaystyle\tilde{\Psi}_{\varepsilon}(y)=\Psi_{\tilde{\lambda}}(y+\varepsilon) for ε0\displaystyle\varepsilon\geq 0, choosing ε00\displaystyle\varepsilon_{0}\geq 0 to be the smallest one, such that

ψλ~,L(X)Ψ~ε0(y)in ΩL{Ψ~ε0<Q} and ψλ~,L(X0)=Ψ~ε0(y0)\psi_{\tilde{\lambda},L}(X)\leq\tilde{\Psi}_{\varepsilon_{0}}(y)\ \text{in $\displaystyle\Omega_{L}\cap\{\tilde{\Psi}_{\varepsilon_{0}}<Q\}$ and $\displaystyle\psi_{\tilde{\lambda},L}(X_{0})=\tilde{\Psi}_{\varepsilon_{0}}(y_{0})$}

for X0=(x0,y0)ΩL{Ψ~ε0<Q}¯\displaystyle X_{0}=(x_{0},y_{0})\in\overline{\Omega_{L}\cap\{\tilde{\Psi}_{\varepsilon_{0}}<Q\}}. It follows from (4.59) that ε0>0\displaystyle\varepsilon_{0}>0. We first show that

X0ΩL{Ψ~ε0<Q}.X_{0}\notin\Omega_{L}\cap\{\tilde{\Psi}_{\varepsilon_{0}}<Q\}. (4.60)

Suppose not, there exists a point X0ΩL{Ψ~ε0<Q}\displaystyle X_{0}\in\Omega_{L}\cap\{\tilde{\Psi}_{\varepsilon_{0}}<Q\}, such that 0<Ψ~ε0(X0)=ψλ~,L(X0)<Q\displaystyle 0<\tilde{\Psi}_{\varepsilon_{0}}(X_{0})=\psi_{\tilde{\lambda},L}(X_{0})<Q. Then there exists a small r>0\displaystyle r>0 such that

0<ψλ~,L<Qand 0<Ψ~ε0<QinBr(X0).0<\psi_{\tilde{\lambda},L}<Q\ \text{and}\ \ 0<\tilde{\Psi}_{\varepsilon_{0}}<Q\ \text{in}\ B_{r}(X_{0}).

On the other hand,

Δ(ψλ~,LΨ~ε0)+f0(ψλ~,L)f0(Ψ~ε0)=0in Br(X0).\Delta(\psi_{\tilde{\lambda},L}-\tilde{\Psi}_{\varepsilon_{0}})+f_{0}(\psi_{\tilde{\lambda},L})-f_{0}(\tilde{\Psi}_{\varepsilon_{0}})=0\ \ \ \text{in $\displaystyle B_{r}(X_{0})$}.

The strong maximum principle gives that ψλ~,L(X0)Ψ~ε0(X0)\displaystyle\psi_{\tilde{\lambda},L}(X_{0})\equiv\tilde{\Psi}_{\varepsilon_{0}}(X_{0}) in Br(X0)\displaystyle B_{r}(X_{0}), due to ψλ~,L(X0)=Ψ~ε0(X0)\displaystyle\psi_{\tilde{\lambda},L}(X_{0})=\tilde{\Psi}_{\varepsilon_{0}}(X_{0}). Applying the strong maximum principle again, then ψλ~,L(X0)Ψ~ε0(X0)\displaystyle\psi_{\tilde{\lambda},L}(X_{0})\equiv\tilde{\Psi}_{\varepsilon_{0}}(X_{0}) in ΩL\displaystyle\Omega_{L}, which leads a contradiction.

The boundary value of ψλ~,L\displaystyle\psi_{\tilde{\lambda},L} gives that X0ΩL\displaystyle X_{0}\notin\partial\Omega_{L}. Next, we claim that

X0(0,kλ~,L(0)).X_{0}\neq(0,k_{\tilde{\lambda},L}(0)). (4.61)

In fact, it follows from Proposition 4.10 that the free boundary Γλ~,L\displaystyle\Gamma_{\tilde{\lambda},L} is C1\displaystyle C^{1} at X0\displaystyle X_{0} and its tangent is in the direction of the positive y\displaystyle y-axis, the definition of ε0\displaystyle\varepsilon_{0} implies the claim (4.61).

Thus, let X0\displaystyle X_{0} be a free boundary point of ψλ~,L\displaystyle\psi_{\tilde{\lambda},L}. Since the free boundary Γλ~,L\displaystyle\Gamma_{\tilde{\lambda},L} is C3,α\displaystyle C^{3,\alpha} at X0\displaystyle X_{0}, it follows from Hopf’s lemma that

λ~=Ψ~ε0ν<ψλ~,Lν=λ~atX0,\tilde{\lambda}=\frac{\partial\tilde{\Psi}_{\varepsilon_{0}}}{\partial\nu}<\frac{\partial\psi_{\tilde{\lambda},L}}{\partial\nu}=\tilde{\lambda}\ \ \text{at}\ \ X_{0},

where ν=(0,1)\displaystyle\nu=(0,1) is the outer normal vector of Γλ~,L\displaystyle\Gamma_{\tilde{\lambda},L} at X0\displaystyle X_{0}, which leads a contradiction.

With the aid of Lemma 4.12, we can define

ΣL={λλ0kλ,L(0)<a}.\Sigma_{L}=\{\lambda\geq\lambda_{0}\mid k_{\lambda,L}(0)<a\}.

and the set ΣL\displaystyle\Sigma_{L} is non-empty for sufficiently large λ\displaystyle\lambda.

Moreover, set

λL=infλΣLλ.\lambda_{L}=\inf_{\lambda\in\Sigma_{L}}\lambda. (4.62)

Furthermore, it follows from Lemma 4.12 that

λ0λLC0<,\lambda_{0}\leq\lambda_{L}\leq C_{0}<\infty, (4.63)

where C0\displaystyle C_{0} is a constant independent of L\displaystyle L.

By virtue of Lemma 4.11 and Lemma 4.12, we have

Proposition 4.13.

For any L>H¯=maxx0g(x)\displaystyle L>\bar{H}=\max_{x\leq 0}g(x), there exists a λLλ0\displaystyle\lambda_{L}\geq\lambda_{0} as in (4.62), such that the free boundary ΓλL,L\displaystyle\Gamma_{\lambda_{L},L} satisfies the continuous fit condition and the smooth fit condition

kλL,L(0)=g(0)=aandkλL,L(0)=g(0).k_{\lambda_{L},L}(0)=g(0)=a\ \ \ \text{and}\ \ \ k^{\prime}_{\lambda_{L},L}(0)=g^{\prime}(0). (4.64)

Moreover,

kλL,L(x)H¯for any x[0,L],k_{\lambda_{L},L}(x)\leq\bar{H}\ \ \ \text{for any $\displaystyle x\in[0,L]$}, (4.65)

and

ψλL,L(x,y)min{ΨλL(y),Q}in ΩL.\psi_{\lambda_{L},L}(x,y)\leq\min\{\Psi_{\lambda_{L}}(y),Q\}\ \ \text{in $\displaystyle\Omega_{L}$}. (4.66)
Proof.

Step 1. The definition of λL\displaystyle\lambda_{L} and the continuous dependence of kλ,L(0)\displaystyle k_{\lambda,L}(0) with respect to λ\displaystyle\lambda imply that

kλL,L(0)=g(0)=a,k_{\lambda_{L},L}(0)=g(0)=a,

which together with Proposition 4.10 gives that kλL,L(0)=g(0)\displaystyle k^{\prime}_{\lambda_{L},L}(0)=g^{\prime}(0). Thus, the continuous fit condition and the smooth fit condition (4.64) hold.

Step 2. Suppose that the assertion (4.65) is not true, then there exists an x~(0,L]\displaystyle\tilde{x}\in(0,L], such that

kλL,L(x~)>H¯=maxx0g(x).k_{\lambda_{L},L}(\tilde{x})>\bar{H}=\max_{x\leq 0}g(x). (4.67)

On the other hand, the fact λLλ0\displaystyle\lambda_{L}\geq\lambda_{0} implies that the asymptotic height hλLHH¯\displaystyle h_{\lambda_{L}}\leq H\leq\bar{H}. Thus we have

kλL,L(x~)>H¯HhλL.k_{\lambda_{L},L}(\tilde{x})>\bar{H}\geq H\geq h_{\lambda_{L}}. (4.68)

For μ<λ0\displaystyle\mu<\lambda_{0} and λ0μ\displaystyle\lambda_{0}-\mu is small, define

ωμ(y)=0yu1,μ(t)𝑑t,u1,μ=u02(χ1(t;pdiff))+2pdiffwithpdiff=μ2λ022,\omega_{\mu}(y)=\int_{0}^{y}u_{1,\mu}(t)dt,\ \ u_{1,\mu}=\sqrt{u_{0}^{2}(\chi^{-1}(t;p_{diff}))+2p_{diff}}\ \ \text{with}\ \ p_{diff}=\frac{\mu^{2}-\lambda^{2}_{0}}{2},

where χ1(t;pdiff)\displaystyle\chi^{-1}(t;p_{diff}) is defined as in Subsection 4.3. Denote ωμε(y)=min{ωμ(yε),Q}\displaystyle\omega_{\mu}^{\varepsilon}(y)=\min\{\omega_{\mu}(y-\varepsilon),Q\} for any ε0\displaystyle\varepsilon\geq 0. Let ε00\displaystyle\varepsilon_{0}\geq 0 be the smallest one, such that

ωμε0(y)ψλL,L(X)in ΩL, and ωμε0(X0)=ψλL,L(X0)\omega_{\mu}^{\varepsilon_{0}}(y)\leq\psi_{\lambda_{L},L}(X)\ \ \text{in $\displaystyle\Omega_{L}$, and $\displaystyle\omega_{\mu}^{\varepsilon_{0}}(X_{0})=\psi_{\lambda_{L},L}(X_{0})$}

for some X0=(x0,y0)ΩL{ψλL,L<Q}¯\displaystyle X_{0}=(x_{0},y_{0})\in\overline{\Omega_{L}\cap\{\psi_{\lambda_{L},L}<Q\}}. It follows from (4.68) that ε0>0\displaystyle\varepsilon_{0}>0. The strong maximum principle implies that X0ΩL{ψλL,L<Q}\displaystyle X_{0}\notin\Omega_{L}\cap\{\psi_{\lambda_{L},L}<Q\}, and thus ψλL,L(X0)=Q\displaystyle\psi_{\lambda_{L},L}(X_{0})=Q. Similar to the proof of Lemma 4.12, we can show that x0L\displaystyle x_{0}\neq L. We next consider the following three cases.

Case 1. X0ΓλL,L\displaystyle X_{0}\in\Gamma_{\lambda_{L},L}. Since the free boundary ΓλL,L\displaystyle\Gamma_{\lambda_{L},L} is C3,α\displaystyle C^{3,\alpha} at X0\displaystyle X_{0}, the Hopf’s lemma gives that

μ=ωμε0ν>ψλL,Lν=λLatX0,\mu=\frac{\partial\omega_{\mu}^{\varepsilon_{0}}}{\partial\nu}>\frac{\partial\psi_{\lambda_{L},L}}{\partial\nu}=\lambda_{L}\ \ \text{at}\ \ X_{0},

where ν=(0,1)\displaystyle\nu=(0,1) is the outer normal vector of ΓλL,L\displaystyle\Gamma_{\lambda_{L},L} at X0\displaystyle X_{0}, which contradicts to the assumption μ<λ0λL\displaystyle\mu<\lambda_{0}\leq\lambda_{L}.

Case 2. X0DL\displaystyle X_{0}\in\partial D_{L} and ΓλL,LBr(X0)=\displaystyle\Gamma_{\lambda_{L},L}\cap B_{r}(X_{0})=\varnothing for small r>0\displaystyle r>0. In view of that the boundary DL\displaystyle\partial D_{L} is C2\displaystyle C^{2} at X0\displaystyle X_{0}, it follows from Hopf’s lemma that

λLλ0>μ=ωμε0ν>ψλL,LνatX0.\lambda_{L}\geq\lambda_{0}>\mu=\frac{\partial\omega_{\mu}^{\varepsilon_{0}}}{\partial\nu}>\frac{\partial\psi_{\lambda_{L},L}}{\partial\nu}\ \ \text{at}\ \ X_{0}. (4.69)

On the other hand, the statement (4) in Proposition 4.4 gives that

ψλL,LνλLatX0,\frac{\partial\psi_{\lambda_{L},L}}{\partial\nu}\geq\lambda_{L}\ \ \text{at}\ \ X_{0},

which contradicts to (4.69).

Case 3. X0DL\displaystyle X_{0}\in\partial D_{L} and ΓλL,LBr(X0)\displaystyle\Gamma_{\lambda_{L},L}\cap B_{r}(X_{0})\neq\varnothing for any r>0\displaystyle r>0. In view of Proposition 4.10 and ωμε0(y)ψλL,L(X)\displaystyle\omega_{\mu}^{\varepsilon_{0}}(y)\leq\psi_{\lambda_{L},L}(X) in ΩL\displaystyle\Omega_{L}, one has

λ0>μ=ωμε0νψλL,Lν=λ0atX0,\lambda_{0}>\mu=\frac{\partial\omega_{\mu}^{\varepsilon_{0}}}{\partial\nu}\geq\frac{\partial\psi_{\lambda_{L},L}}{\partial\nu}=\lambda_{0}\ \ \text{at}\ \ X_{0},

which leads a contradiction.

Step 3. In this step, we will show the assertion (4.66). Denote ωε(y)=min{ΨλL(y+ε),Q}\displaystyle\omega_{\varepsilon}(y)=\min\{\Psi_{\lambda_{L}}(y+\varepsilon),Q\} for any ε0\displaystyle\varepsilon\geq 0, and choosing the smallest ε00\displaystyle\varepsilon_{0}\geq 0, such that

ωε0(y)ψλL,L(X)in ΩL{ψλL,L<Q}, and ωε0(y0)=ψλL,L(X0)\omega_{\varepsilon_{0}}(y)\geq\psi_{\lambda_{L},L}(X)\ \ \text{in $\displaystyle\Omega_{L}\cap\{\psi_{\lambda_{L},L}<Q\}$, and $\displaystyle\omega_{\varepsilon_{0}}(y_{0})=\psi_{\lambda_{L},L}(X_{0})$}

for some X0=(x0,y0)ΩL{ωε0<Q}¯\displaystyle X_{0}=(x_{0},y_{0})\in\overline{\Omega_{L}\cap\{\omega_{\varepsilon_{0}}<Q\}}. It suffices to show that

ε0=0.\varepsilon_{0}=0.

If not, suppose that ε0>0\displaystyle\varepsilon_{0}>0. By virtue of the strong maximum principle and the boundary value of ψλL,L\displaystyle\psi_{\lambda_{L},L}, we can choose X0\displaystyle X_{0} to be the free boundary point of ψλL,L\displaystyle\psi_{\lambda_{L},L}. Thus, we can derive a contradiction by using Hopf’s lemma.

4.7. The existence of the incompressible jet flow

In previous subsections, we show that there exist a λL\displaystyle\lambda_{L} and a minimizer ψλL,L\displaystyle\psi_{\lambda_{L},L} for any L>H¯\displaystyle L>\bar{H}, such that the free boundary ΓλL,L\displaystyle\Gamma_{\lambda_{L},L} satisfies the continuous fit condition (4.64). Taking L+\displaystyle L\rightarrow+\infty, we will obtain the existence of the solution to the jet flow problem in this subsection.

First, we consider the following variational problem.

The variational problem (Pλ)\displaystyle(P_{\lambda}): For any L0>H¯=maxx0g(x)\displaystyle L_{0}>\bar{H}=\max_{x\leq 0}g(x), find a ψλK\displaystyle\psi_{\lambda}\in K such that

Jλ,L0(ψλ)Jλ,L0(ψ),J_{\lambda,L_{0}}(\psi_{\lambda})\leq J_{\lambda,L_{0}}(\psi),

for any ψK\displaystyle\psi\in K with ψ=ψλ\displaystyle\psi=\psi_{\lambda} on ΩL0\displaystyle\partial\Omega_{L_{0}}, the admissible set

K={ψHloc1(2)ψ=0lies belowT,ψ=Qlies aboveN}.K=\{\psi\in H^{1}_{loc}(\mathbb{R}^{2})\mid\psi=0~~\text{lies below}~T,\psi=Q~~\text{lies above}~N\}.

By using the similar arguments in Lemma 4.11, taking a sequence {Ln}\displaystyle\{L_{n}\} with Ln\displaystyle L_{n}\rightarrow\infty, such that

λLnλandψλLn,Lnψλweakly   inHloc1(2)and uniformly in any compact subset of 2,\lambda_{L_{n}}\rightarrow\lambda~~\text{and}~~\psi_{\lambda_{L_{n}},L_{n}}\rightarrow\psi_{\lambda}\ \ \text{weakly~~ in}~~H_{loc}^{1}(\mathbb{R}^{2})~\text{and uniformly in any compact subset of $\displaystyle\mathbb{R}^{2}$},

as n\displaystyle n\rightarrow\infty, and ψλ\displaystyle\psi_{\lambda} is a minimizer to the variational problem (Pλ)\displaystyle(P_{\lambda}). It follows from (4.63) that

λ0λ<+.\lambda_{0}\leq\lambda<+\infty.

Lemma 4.6 gives that ψλ(x,y)\displaystyle\psi_{\lambda}(x,y) is monotone increasing with respect to y\displaystyle y, and the free boundary Γλ\displaystyle\Gamma_{\lambda} of ψλ\displaystyle\psi_{\lambda} is x\displaystyle x-graph. Moreover, Γλ\displaystyle\Gamma_{\lambda} can be described by a continuous function y=kλ(x)\displaystyle y=k_{\lambda}(x) for x(0,+)\displaystyle x\in(0,+\infty), and

Γλ=Ω{ψλ<Q}={(x,y)Ωx>0,y=kλ(x)}.\Gamma_{\lambda}=\Omega\cap\partial\{\psi_{\lambda}<Q\}=\{(x,y)\in\Omega\mid x>0,y=k_{\lambda}(x)\}.

By virtue of Proposition 4.13, one has

kλ(0)=g(0)=a,kλ(0)=g(0)and kλ(x)H¯ for any x(0,),k_{\lambda}(0)=g(0)=a,\ \ k^{\prime}_{\lambda}(0)=g^{\prime}(0)\ \ \text{and $\displaystyle k_{\lambda}(x)\leq\bar{H}$ for any $\displaystyle x\in(0,\infty)$}, (4.70)

and

ψλ(x,y)min{Ψλ(y),Q}in Ω.\psi_{\lambda}(x,y)\leq\min\{\Psi_{\lambda}(y),Q\}\ \ \text{in $\displaystyle\Omega$}. (4.71)

Theorem 3.15 and Proposition 2.2 give that the free boundary Γλ\displaystyle\Gamma_{\lambda} is C3,α\displaystyle C^{3,\alpha} and |ψλ|=ψλν=λ\displaystyle|\nabla\psi_{\lambda}|=\frac{\partial\psi_{\lambda}}{\partial\nu}=\lambda on Γλ\displaystyle\Gamma_{\lambda}, where ν\displaystyle\nu is the outer normal vector.

Next, we will obtain the positivity of horizontal velocity.

Lemma 4.14.

u=ψλy>0\displaystyle u=\frac{\partial\psi_{\lambda}}{\partial y}>0 in Ω{ψλ<Q}¯\displaystyle\overline{\Omega\cap\{\psi_{\lambda}<Q\}}.

Proof.

Denote u=ψλy\displaystyle u=\frac{\partial\psi_{\lambda}}{\partial y}, one has

Δu+b(X)u=0inΩ{ψλ<Q},\Delta u+b(X)u=0\ \ \text{in}~\Omega\cap\{\psi_{\lambda}<Q\},

where b(X)=f0(ψ(X))0\displaystyle b(X)=f_{0}^{\prime}(\psi(X))\leq 0. The monotonicity of ψλ(x,y)\displaystyle\psi_{\lambda}(x,y) with respect to y\displaystyle y implies that u0\displaystyle u\geq 0 in Ω{ψλ<Q}\displaystyle\Omega\cap\{\psi_{\lambda}<Q\}, the strong maximum principle gives that

u>0 in Ω{ψλ<Q}.\text{$\displaystyle u>0$ in $\displaystyle\Omega\cap\{\psi_{\lambda}<Q\}$}.

Since ψλ<Q\displaystyle\psi_{\lambda}<Q in Ω{x<0}\displaystyle\Omega\cap\{x<0\} and ψλ=Q\displaystyle\psi_{\lambda}=Q on N\displaystyle N, it follows from f0(Q)0\displaystyle f^{\prime}_{0}(Q)\leq 0 and Hopf’s lemma that

u=11+(g(x))2ψλν>0on NA.u=\frac{1}{\sqrt{1+(g^{\prime}(x))^{2}}}\frac{\partial\psi_{\lambda}}{\partial\nu}>0\ \ \text{on $\displaystyle N\setminus A$}.

Similarly, we have

u>0on T.u>0\ \ \text{on $\displaystyle T$.}

Next, we will show that u>0\displaystyle u>0 on Γλ\displaystyle\Gamma_{\lambda}. Suppose that there exists x0>0\displaystyle x_{0}>0, such that u=yψλ=0\displaystyle u=\partial_{y}\psi_{\lambda}=0 on (x0,kλ(x0))\displaystyle(x_{0},k_{\lambda}(x_{0})). Since the free boundary Γλ\displaystyle\Gamma_{\lambda} is C3,α\displaystyle C^{3,\alpha}-smooth at (x0,kλ(x0))\displaystyle(x_{0},k_{\lambda}(x_{0})), the normal outer vector is (1,0)\displaystyle(1,0) at (x0,kλ(x0))\displaystyle(x_{0},k_{\lambda}(x_{0})) and |xψλ|=λ\displaystyle|\partial_{x}\psi_{\lambda}|=\lambda on Γλ\displaystyle\Gamma_{\lambda}. Thanks to Hopf’s lemma, we have

|ux|=|uν|>0at(x0,kλ(x0)).\left|u_{x}\right|=\left|\frac{\partial u}{\partial\nu}\right|>0\ \ \text{at}\ \ (x_{0},k_{\lambda}(x_{0})). (4.72)

Since u2+v2=λ2\displaystyle u^{2}+v^{2}=\lambda^{2} on Γλ\displaystyle\Gamma_{\lambda}, we have

0=(u2+v2)s=2uuy+2vvy=2vvy=2λuxat (x0,kλ(x0)),\begin{array}[]{rl}0=\frac{\partial(u^{2}+v^{2})}{\partial s}=2uu_{y}+2vv_{y}=2vv_{y}=-2\lambda u_{x}\ \ \text{at $\displaystyle(x_{0},k_{\lambda}(x_{0}))$,}\end{array}

which contradicts to (4.72).

By virtue of the boundedness of g(0)\displaystyle g^{\prime}(0), we have u>0\displaystyle u>0 at A\displaystyle A.

The positivity of horizontal velocity u\displaystyle u implies that the function y=kλ(x)\displaystyle y=k_{\lambda}(x) is C1\displaystyle C^{1} for any x0\displaystyle x\geq 0. ∎

Next, we will obtain the asymptotic height of the free boundary Γλ\displaystyle\Gamma_{\lambda}.

Lemma 4.15.

limxkλ(x)=hλ\displaystyle\lim_{x\rightarrow\infty}k_{\lambda}(x)=h_{\lambda}, where hλ\displaystyle h_{\lambda} is the asymptotic height of the free boundary Γλ\displaystyle\Gamma_{\lambda} and hλa\displaystyle h_{\lambda}\leq a.

Proof.

Suppose that the limit limxkλ(x)\displaystyle\lim_{x\rightarrow\infty}k_{\lambda}(x) does not exist. It follows from the non-degeneracy 2.6 and (4.70) that

0<c0kλ(x)H¯for any x[0,),0<c_{0}\leq k_{\lambda}(x)\leq\bar{H}\ \ \ \text{for any $\displaystyle x\in[0,\infty)$},

which implies that there exist two sequence {xn}\displaystyle\{x_{n}\} and {x~n}\displaystyle\{\tilde{x}_{n}\}, such that

limxnkλ(xn)=lim supxkλ(x)=hλ,1andlimx~nkλ(x~n)=lim infxkλ(x)=hλ,2\lim_{x_{n}\rightarrow\infty}k_{\lambda}(x_{n})=\limsup_{x\rightarrow\infty}k_{\lambda}(x)=h_{\lambda,1}\ \ \text{and}\ \ \lim_{\tilde{x}_{n}\rightarrow\infty}k_{\lambda}(\tilde{x}_{n})=\liminf_{x\rightarrow\infty}k_{\lambda}(x)=h_{\lambda,2} (4.73)

with hλ,1>hλ,2\displaystyle h_{\lambda,1}>h_{\lambda,2}.

By virtue of Remark 4.4, there exist λ1,λ2\displaystyle\lambda_{1},\lambda_{2} with λ0λ1<λ2<\displaystyle\lambda_{0}\leq\lambda_{1}<\lambda_{2}<\infty, such that

0<c0hλ,2=hλ2<hλ1=hλ,1H¯.0<c_{0}\leq h_{\lambda,2}=h_{\lambda_{2}}<h_{\lambda_{1}}=h_{\lambda,1}\leq\bar{H}. (4.74)

Set ψn(t,s)=ψλ(xn+t,s)\displaystyle\psi_{n}(t,s)=\psi_{\lambda}(x_{n}+t,s) and the free boundary Γn:s=kλ(xn+t)\displaystyle\Gamma_{n}:s=k_{\lambda}(x_{n}+t). Define a curve

γδ:s=H¯δη(txn)\gamma_{\delta}:s=\bar{H}-\delta\eta\left(\frac{t}{\sqrt{x_{n}}}\right)

where the function η(x)\displaystyle\eta(x) is defined as in (4.55). Let δn\displaystyle\delta_{n} be the largest one, such that the curve γδn\displaystyle\gamma_{\delta_{n}} touches the free boundary Γn\displaystyle\Gamma_{n} at some points (tn,sn)\displaystyle(t_{n},s_{n}), and thus

sn=kλ(xn+tn)\displaystyle s_{n}=k_{\lambda}(x_{n}+t_{n}) and δnH¯kλ(xn)\displaystyle\delta_{n}\leq\bar{H}-k_{\lambda}(x_{n}). (4.75)

Let ωn\displaystyle\omega_{n} be the solution to the Dirichlet problem

{Δω+f0(ω)=0inEn,ω=Qonγδn,ω=ψnonEnγδn,\left\{\begin{array}[]{ll}&\Delta\omega+f_{0}(\omega)=0\ \text{in}\ \ \ E_{n},\\ &\omega=Q\ \text{on}\ \gamma_{\delta_{n}},\ \ \omega=\psi_{n}\ \ \text{on}\ \ \partial E_{n}\setminus\gamma_{\delta_{n}},\end{array}\right. (4.76)

where domain En\displaystyle E_{n} is bounded by t=xn2\displaystyle t=-\frac{x_{n}}{2}, t=xn2\displaystyle t=\frac{x_{n}}{2}, s=0\displaystyle s=0 and γδn\displaystyle\gamma_{\delta_{n}}.

Since Δψn+f0(ψn)0\displaystyle\Delta\psi_{n}+f_{0}(\psi_{n})\leq 0 in En\displaystyle E_{n} and ψn(tn,sn)=ωn(tn,sn)=Q\displaystyle\psi_{n}(t_{n},s_{n})=\omega_{n}(t_{n},s_{n})=Q, the maximum principle implies that ψnωn\displaystyle\psi_{n}\geq\omega_{n} in En\displaystyle E_{n} and

λ=ψnνnωnνnat(tn,sn),\lambda=\frac{\partial\psi_{n}}{\partial\nu_{n}}\leq\frac{\partial\omega_{n}}{\partial\nu_{n}}\ \ \text{at}\ \ (t_{n},s_{n}), (4.77)

where νn\displaystyle\nu_{n} is the outer normal vector of Γn\displaystyle\Gamma_{n}.

By virtue of (4.75), we can choose a subsequence {xn}\displaystyle\{x_{n}\}, such that

tnxnt0[1,1]andH¯δnη(txn)h¯λ1.\frac{t_{n}}{\sqrt{x_{n}}}\rightarrow t_{0}\in[-1,1]\ \ \text{and}\ \ \bar{H}-\delta_{n}\eta\left(\frac{t}{\sqrt{x_{n}}}\right)\rightarrow\bar{h}_{\lambda_{1}}. (4.78)

After a translation in the t\displaystyle t direction, such that the free boundary point (tn,sn)\displaystyle(t_{n},s_{n}) lies on the s\displaystyle s-axis, then we have

EnE0={(t,s)<t<,0<s<hλ1}and(tn,sn)(0,hλ1),E_{n}\rightarrow E_{0}=\{(t,s)\mid-\infty<t<\infty,0<s<h_{\lambda_{1}}\}\ \ \text{and}\ \ (t_{n},s_{n})\rightarrow(0,h_{\lambda_{1}}), (4.79)

and ωnω0\displaystyle\omega_{n}\rightarrow\omega_{0} uniformly in any compact subset of E0\displaystyle E_{0}. Moreover, ω0\displaystyle\omega_{0} satisfies that 0ω0(t,s)Q\displaystyle 0\leq\omega_{0}(t,s)\leq Q and

{Δω0+f0(ω0)=0inE0,ω0(t,0)=0,ω0(t,hλ1)=Q.\left\{\begin{array}[]{ll}&\Delta\omega_{0}+f_{0}(\omega_{0})=0\ \text{in}\ \ \ E_{0},\\ &\omega_{0}(t,0)=0,\ \ \omega_{0}(t,h_{\lambda_{1}})=Q.\end{array}\right. (4.80)

By using the similar arguments in [35], we can show that the Dirichlet problem (4.80) has a unique solution

ω0(s)=0su1(x)𝑑x,\omega_{0}(s)=\int_{0}^{s}u_{1}(x)dx,

where u1(t)=u02(χ1(t;pdiff))+2pdiff\displaystyle u_{1}(t)=\sqrt{u_{0}^{2}(\chi^{-1}(t;p_{diff}))+2p_{diff}} with pdiff=λ12λ022\displaystyle p_{diff}=\frac{\lambda_{1}^{2}-\lambda^{2}_{0}}{2}, and χ1(t;pdiff)\displaystyle\chi^{-1}(t;p_{diff}) is defined as in Subsection 4.3.

It follows from (4.77) and (4.79) that

λlimnωn(tn,sn)νn=ω0(0,hλ1)s=λ1.\lambda\leq\lim_{n\rightarrow\infty}\frac{\partial\omega_{n}(t_{n},s_{n})}{\partial\nu_{n}}=\frac{\omega_{0}(0,h_{\lambda_{1}})}{\partial s}=\lambda_{1}. (4.81)

Next, we will show that

λλ2.\lambda\geq\lambda_{2}. (4.82)

.

Define ψ~n(t,s)=ψλ(x~n+t,s)\displaystyle\tilde{\psi}_{n}(t,s)=\psi_{\lambda}(\tilde{x}_{n}+t,s), Γ~n:s=kλ(x~n+t)\displaystyle\tilde{\Gamma}_{n}:s=k_{\lambda}(\tilde{x}_{n}+t) as the free boundary of ψ~n\displaystyle\tilde{\psi}_{n} and a curve

γ~δ:s=h22+δη(tx~n).\tilde{\gamma}_{\delta}:s=\frac{h_{2}}{2}+\delta\eta\left(\frac{t}{\sqrt{\tilde{x}_{n}}}\right).

Let δn\displaystyle\delta_{n} be the largest one, such that the curve γ~δn\displaystyle\tilde{\gamma}_{\delta_{n}} touches the free boundary Γ~n\displaystyle\tilde{\Gamma}_{n} at a point (tn,sn)\displaystyle(t_{n},s_{n}), and thus

sn=kλ(x~n+tn)\displaystyle s_{n}=k_{\lambda}(\tilde{x}_{n}+t_{n}) and δnkλ(x~n)hλ,22\displaystyle\delta_{n}\leq k_{\lambda}(\tilde{x}_{n})-\frac{h_{\lambda,2}}{2}.

Denote a domain E~n\displaystyle\tilde{E}_{n}, which is bounded by t=x~n2\displaystyle t=-\frac{\tilde{x}_{n}}{2}, t=x~n2\displaystyle t=\frac{\tilde{x}_{n}}{2}, s=0\displaystyle s=0 and γ~δn\displaystyle\tilde{\gamma}_{\delta_{n}}. Let ω~n\displaystyle\tilde{\omega}_{n} be the solutions to the Dirichlet problem (4.76), replacing En\displaystyle E_{n} and γδn\displaystyle\gamma_{\delta_{n}} by E~n\displaystyle\tilde{E}_{n} and γ~δn\displaystyle\tilde{\gamma}_{\delta_{n}}. Obviously, Δψn+f0(ψn)=0\displaystyle\Delta\psi_{n}+f_{0}(\psi_{n})=0 in E~n\displaystyle\tilde{E}_{n}, it follows from the maximum principle that ω~nψ~n\displaystyle\tilde{\omega}_{n}\geq\tilde{\psi}_{n} in E~n\displaystyle\tilde{E}_{n} and

λ=ψ~nνnω~nνnat(tn,sn),\lambda=\frac{\partial\tilde{\psi}_{n}}{\partial\nu_{n}}\geq\frac{\partial\tilde{\omega}_{n}}{\partial\nu_{n}}\ \ \text{at}\ \ (t_{n},s_{n}), (4.83)

where νn\displaystyle\nu_{n} is the outer normal vector of Γn\displaystyle\Gamma_{n}. By using the previous arguments, we can show that there exists a subsequence {x~n}\displaystyle\{\tilde{x}_{n}\}, such that

E~nE~0={(t,s)<t<,0<s<hλ2}and(tn,sn)(0,hλ2),\tilde{E}_{n}\rightarrow\tilde{E}_{0}=\{(t,s)\mid-\infty<t<\infty,0<s<h_{\lambda_{2}}\}\ \ \text{and}\ \ (t_{n},s_{n})\rightarrow(0,h_{\lambda_{2}}),

and ω~nω~0\displaystyle\tilde{\omega}_{n}\rightarrow\tilde{\omega}_{0} uniformly in any compact subset of E0\displaystyle E_{0} and

limnω~n(tn,sn)νn=ω~0(0,hλ2)s=λ2,\lim_{n\rightarrow\infty}\frac{\partial\tilde{\omega}_{n}(t_{n},s_{n})}{\partial\nu_{n}}=\frac{\tilde{\omega}_{0}(0,h_{\lambda_{2}})}{\partial s}=\lambda_{2},

which together with (4.83) give that (4.82) holds.

Hence, it follows from (4.81) and (4.82) that

λ2λλ1,\lambda_{2}\leq\lambda\leq\lambda_{1},

which contradicts to the fact λ1<λ2\displaystyle\lambda_{1}<\lambda_{2}.

Thus, there exists an asymptotic height h~\displaystyle\tilde{h} of the free boundary Γλ\displaystyle\Gamma_{\lambda}, such that

kλ(x)h~.k_{\lambda}(x)\rightarrow\tilde{h}.

Next, we claim that

h~=hλ.\text{$\displaystyle\tilde{h}=h_{\lambda}$}.

If not, without loss of generality, we assume that h~>hλ\displaystyle\tilde{h}>h_{\lambda}, and there exists a λ~<λ\displaystyle\tilde{\lambda}<\lambda, such that h~=hλ~\displaystyle\tilde{h}=h_{\tilde{\lambda}}. Similar to the proof of (4.81) and (4.82), we can show that

λλ~,\lambda\leq\tilde{\lambda},

which contradicts to the assumption λ~<λ\displaystyle\tilde{\lambda}<\lambda.

Finally, it follows from (4.71) that

hλa.h_{\lambda}\leq a.

Finally, the asymptotic behavior of the jet flow will be obtained in the following.

Proposition 4.16.

The rotational jet flow satisfies the following asymptotic behavior in the far fields,

(u,v,p)(u0(y),0,pin),u(0,u0(y)),v0,p0,(u,v,p)\rightarrow(u_{0}(y),0,p_{in}),\ \nabla u\rightarrow(0,u_{0}^{\prime}(y)),~~\nabla v\rightarrow 0,~~\nabla p\rightarrow 0,

uniformly in any compact subset of (0,H)\displaystyle(0,H), as x\displaystyle x\rightarrow-\infty, and

(u,v,p)(u1(y),0,patm),u(0,u1(y)),v0,p0,(u,v,p)\rightarrow(u_{1}(y),0,p_{atm}),\ \nabla u\rightarrow(0,u_{1}^{\prime}(y)),~~\nabla v\rightarrow 0,~~\nabla p\rightarrow 0,

uniformly in any compact subset of (0,hλ)\displaystyle(0,h_{\lambda}), as x+,\displaystyle x\rightarrow+\infty, where u1(y)\displaystyle u_{1}(y) and hλ\displaystyle h_{\lambda} are uniquely determined by u0(y)\displaystyle u_{0}(y), pin\displaystyle p_{in} and patm\displaystyle p_{atm} as in Remark 4.4.

Proof.

For any sequence {ψλ(xn,y)}\displaystyle\{\psi_{\lambda}(x-n,y)\}, by using the uniform elliptic estimate, there exists a subsequence {ψλ(xn,y)}\displaystyle\{\psi_{\lambda}(x-n,y)\}, such that

ψλ(xn,y)ψ0(x,y)in C2,α(S),\psi_{\lambda}(x-n,y)\rightarrow\psi_{0}(x,y)\ \ \text{in $\displaystyle C^{2,\alpha}(S)$},

where S\displaystyle S is any compact subsect of (0,H)\displaystyle(0,H), 0ψ0Q\displaystyle 0\leq\psi_{0}\leq Q and

{Δψ0+f0(ψ0)=0in{<x<+}×{0<y<H},ψ0(x,H)=Qandψ0(x,0)=0for <x<+.\left\{\begin{array}[]{ll}&\Delta\psi_{0}+f_{0}(\psi_{0})=0\ \text{in}~~\{-\infty<x<+\infty\}\times\{0<y<H\},\\ &\psi_{0}(x,H)=Q\ \text{and}\ \psi_{0}(x,0)=0\ \ \text{for $\displaystyle-\infty<x<+\infty$}.\end{array}\right. (4.84)

Then boundary value problem (4.84) possesses a unique solution

ψ0(y)=0yu0(s)𝑑s,\psi_{0}(y)=\int_{0}^{y}u_{0}(s)ds,

which implies that

(u,v)(x,y)(u0(y),0),u(x,y)(0,u0(y)))andv(x,y)(0,0)(u,v)(x,y)\rightarrow\left(u_{0}(y),0\right),\ \nabla u(x,y)\rightarrow\left(0,u_{0}^{\prime}(y))\right)\ \text{and}\ \nabla v(x,y)\rightarrow(0,0)

uniformly in any compact subset of (0,H)\displaystyle(0,H), as x\displaystyle x\rightarrow-\infty. The Bernoulli’s law gives that

p(x,y)pin=λ22u02(H)2+patmandp(x,y)0,p(x,y)\rightarrow p_{in}=\frac{\lambda^{2}}{2}-\frac{u_{0}^{2}(H)}{2}+p_{atm}\ \text{and}\ \nabla p(x,y)\rightarrow 0,

uniformly in any compact subset of (0,H)\displaystyle(0,H), as x\displaystyle x\rightarrow-\infty.

For any sequence {ψn}\displaystyle\{\psi_{n}\} with ψn(x,y)=ψλ(x+n,y)\displaystyle\psi_{n}(x,y)=\psi_{\lambda}(x+n,y) and R>0\displaystyle R>0, by virtue of Lemma 4.15, one has

limnkλ(x+n)=hλ.\lim_{n\rightarrow\infty}k_{\lambda}(x+n)=h_{\lambda}. (4.85)

For any large R\displaystyle R and small ε>0\displaystyle\varepsilon>0, it follows from (4.85) that there exists a N=N(R,ε)\displaystyle N=N(R,\varepsilon), such that the free boundary of ψn(X)\displaystyle\psi_{n}(X) lies in a ε\displaystyle\varepsilon-neighborhood of the line {y=hλ}\displaystyle\{y=h_{\lambda}\} in BR(0)\displaystyle B_{R}(0), and thus the free boundary of ψn\displaystyle\psi_{n} satisfies the flatness condition in BR(0)\displaystyle B_{R}(0), provided that n\displaystyle n is sufficiently large. It follows from Theorem 3.15 that the free boundary of ψn\displaystyle\psi_{n} convergence to the line {y=hλ}\displaystyle\{y=h_{\lambda}\} in C1\displaystyle C^{1} norm, namely,

kλ(x+n)0asn,for any |x|<R.k^{\prime}_{\lambda}(x+n)\rightarrow 0\ \ \text{as}\ n\rightarrow\infty,\ \ \text{for any $\displaystyle|x|<R$}.

Applying Theorem 3.15 again, one has

|kλ(j)(x)|Cfor sufficiently large x>0,j=2,3.\left|k^{(j)}_{\lambda}(x)\right|\leq C\ \ \text{for sufficiently large $\displaystyle x>0$},\ \ j=2,3.

By using the uniform elliptic estimate, there exists a subsequence {ψn}\displaystyle\{\psi_{n}\}, such that

ψn(x,y)ψ1(x,y),\psi_{n}(x,y)\rightarrow\psi_{1}(x,y),

and

{Δψ1+f0(ψ1)=0inE,ψ1(x,hλ)=Qandψ1(x,0)=0for <x<+,\left\{\begin{array}[]{ll}&\Delta\psi_{1}+f_{0}(\psi_{1})=0\ \text{in}~~E,\\ &\psi_{1}(x,h_{\lambda})=Q\ \ \text{and}\ \ \psi_{1}(x,0)=0\ \ \text{for $\displaystyle-\infty<x<+\infty$},\end{array}\right. (4.86)

where E={<x<+}×{0<y<hλ}\displaystyle E=\{-\infty<x<+\infty\}\times\{0<y<h_{\lambda}\}.

Recalling Remark 4.4,

ψ1(y)=0yu1(y)𝑑sinE,\psi_{1}(y)=\int_{0}^{y}u_{1}(y)ds\ \ \ \text{in}\ \ E, (4.87)

solves uniquely Dirichlet problem (4.86) in E\displaystyle E, where u1(t)=u02(χ1(y;pdiff))+2pdiff\displaystyle u_{1}(t)=\sqrt{u_{0}^{2}(\chi^{-1}(y;p_{diff}))+2p_{diff}} with pdiff=λ2u02(H)2\displaystyle p_{diff}=\frac{\lambda^{2}-u_{0}^{2}(H)}{2}.

In view of (4.87), we conclude that

(u,v,p)(x,y)(u1(y),0,patm),u(x,y)(0,u1(y)),v(x,y)(0,0)(u,v,p)(x,y)\rightarrow\left(u_{1}(y),0,p_{atm}\right),\ \nabla u(x,y)\rightarrow\left(0,u_{1}^{\prime}(y)\right),\ \nabla v(x,y)\rightarrow(0,0)

and p(x,y)(0,0)\displaystyle\nabla p(x,y)\rightarrow(0,0) uniformly in any compact subset of (0,hλ)\displaystyle(0,h_{\lambda}), as x+\displaystyle x\rightarrow+\infty.

4.8. The uniqueness of the incompressible jet flow

In this subsection, we will obtain the uniqueness of λ\displaystyle\lambda and the solution to the jet flow problem.

Proposition 4.17.

λ\displaystyle\lambda and the solution (u,v,p,Γ)\displaystyle(u,v,p,\Gamma) to the jet flow problem are unique.

Proof.

Suppose that (ψλ1,λ1,Γλ1)\displaystyle(\psi_{\lambda_{1}},\lambda_{1},\Gamma_{\lambda_{1}}) and (ψ~λ2,λ2,Γ~λ2)\displaystyle(\tilde{\psi}_{\lambda_{2}},\lambda_{2},\tilde{\Gamma}_{\lambda_{2}}) are two solutions to the jet flow problem, which satisfy the conditions in Definition 4.1.

Step 1. In this step, we will show that λ1=λ2\displaystyle\lambda_{1}=\lambda_{2}. Suppose not, without loss of generality, we assume that λ1<λ2\displaystyle\lambda_{1}<\lambda_{2}. In view of the asymptotic behaviors of ψλ1\displaystyle\psi_{\lambda_{1}} and ψ~λ2\displaystyle\tilde{\psi}_{\lambda_{2}} in Proposition 4.16, one has

kλ1(x)>k~λ2(x)for sufficiently large x>0.k_{\lambda_{1}}(x)>\tilde{k}_{\lambda_{2}}(x)\ \ \text{for sufficiently large $\displaystyle x>0$.} (4.88)

Consider a function ψλ1ε(x,y)=ψλ1(x,yε)\displaystyle\psi_{\lambda_{1}}^{\varepsilon}(x,y)=\psi_{\lambda_{1}}(x,y-\varepsilon) for ε0\displaystyle\varepsilon\geq 0 and Γλ1ε\displaystyle\Gamma_{\lambda_{1}}^{\varepsilon} is the free boundary of ψλ1ε\displaystyle\psi_{\lambda_{1}}^{\varepsilon}, choosing the smallest ε00\displaystyle\varepsilon_{0}\geq 0 such that

ψλ1ε0(X)ψ~λ2(X)in Ω, andψλ1ε0(X0)=ψ~λ2(X0)for some X0Ω{ψ~λ2<Q}¯.\psi_{\lambda_{1}}^{\varepsilon_{0}}(X)\leq\tilde{\psi}_{\lambda_{2}}(X)\ \text{in $\displaystyle\Omega$, and}\ \psi_{\lambda_{1}}^{\varepsilon_{0}}(X_{0})=\tilde{\psi}_{\lambda_{2}}(X_{0})\ \text{for some $\displaystyle X_{0}\in\overline{\Omega\cap\{\tilde{\psi}_{\lambda_{2}}<Q\}}$}. (4.89)

Similar to the proof of (4.60), the strong maximum principle implies that X0Ω{ψ~λ2<Q}\displaystyle X_{0}\notin\Omega\cap\{\tilde{\psi}_{\lambda_{2}}<Q\}. We next consider the following two cases.

Case 1. ε0=0\displaystyle\varepsilon_{0}=0, we can choose X0=A\displaystyle X_{0}=A. Then one has

ψλ1ν=|ψλ1|=λ1andψ~λ2ν=|ψ~λ2|=λ2at A,\frac{\partial\psi_{\lambda_{1}}}{\partial\nu}=|\nabla\psi_{\lambda_{1}}|=\lambda_{1}\ \ \text{and}\ \ \frac{\partial\tilde{\psi}_{\lambda_{2}}}{\partial\nu}=|\nabla\tilde{\psi}_{\lambda_{2}}|=\lambda_{2}\ \ \ \text{at $\displaystyle A$,}

where ν\displaystyle\nu is the outer normal vector. Since ψλ1ψ~λ2\displaystyle\psi_{\lambda_{1}}\leq\tilde{\psi}_{\lambda_{2}} in Ω\displaystyle\Omega, which yields that

λ1=ψλ1νψ~λ2ν=λ2at A,\lambda_{1}=\frac{\partial\psi_{\lambda_{1}}}{\partial\nu}\geq\frac{\partial\tilde{\psi}_{\lambda_{2}}}{\partial\nu}=\lambda_{2}\ \ \ \text{at $\displaystyle A$,}

which leads a contradiction to the assumption λ1<λ2\displaystyle\lambda_{1}<\lambda_{2}.

Case 2. ε0>0\displaystyle\varepsilon_{0}>0, it follows from (4.88) that |X0|<+\displaystyle|X_{0}|<+\infty. Then we can take X0\displaystyle X_{0} be the free boundary point of ψλ1ε0\displaystyle\psi^{\varepsilon_{0}}_{\lambda_{1}} and ψ~λ2\displaystyle\tilde{\psi}_{\lambda_{2}}, and the strong maximum principle gives that

ψλ1ε0(X)<ψ~λ2(X)in Ω{ψ~λ~2<Q}.\psi_{\lambda_{1}}^{\varepsilon_{0}}(X)<\tilde{\psi}_{\lambda_{2}}(X)\ \ \text{in $\displaystyle\Omega\cap\{\tilde{\psi}_{\tilde{\lambda}_{2}}<Q\}$}.

Since the free boundaries Γλ1ε0\displaystyle\Gamma^{\varepsilon_{0}}_{\lambda_{1}} and Γ~λ2\displaystyle\tilde{\Gamma}_{\lambda_{2}} are C3,α\displaystyle C^{3,\alpha} at X0\displaystyle X_{0}, it follows from Hopf’s lemma that

λ1=ψλ1ε0ν>ψ~λ2ν=λ2at X0,\lambda_{1}=\frac{\partial\psi_{\lambda_{1}}^{\varepsilon_{0}}}{\partial\nu}>\frac{\partial\tilde{\psi}_{\lambda_{2}}}{\partial\nu}=\lambda_{2}\ \ \ \text{at $\displaystyle X_{0}$,}

where ν\displaystyle\nu is the outer normal vector to Γλ1ε0Γ~λ2\displaystyle\Gamma_{\lambda_{1}}^{\varepsilon_{0}}\cap\tilde{\Gamma}_{\lambda_{2}} at X0\displaystyle X_{0}. This leads a contradiction to the assumption λ1<λ2\displaystyle\lambda_{1}<\lambda_{2}.

Hence, we obtain the uniqueness of λ\displaystyle\lambda, and denote λ=λ1=λ2\displaystyle\lambda=\lambda_{1}=\lambda_{2}.

Step 2. In this step, we will show that ψλ=ψ~λ\displaystyle\psi_{\lambda}=\tilde{\psi}_{\lambda}. By virtue of the asymptotic behaviors of ψλ1\displaystyle\psi_{\lambda_{1}} and ψ~λ2\displaystyle\tilde{\psi}_{\lambda_{2}} in Proposition 4.16, we have

limx+kλ(x)=limx+k~λ(x)=hλ.\lim_{x\rightarrow+\infty}k_{\lambda}(x)=\lim_{x\rightarrow+\infty}\tilde{k}_{\lambda}(x)=h_{\lambda}.

Without loss of generality, we assume that there exists some x0(0,+)\displaystyle x_{0}\in(0,+\infty), such that

kλ(x0)<k~λ(x0).k_{\lambda}(x_{0})<\tilde{k}_{\lambda}(x_{0}). (4.90)

Denote ψλε(x,y)=ψλ(x,yε)\displaystyle\psi_{\lambda}^{\varepsilon}(x,y)=\psi_{\lambda}(x,y-\varepsilon) for ε0\displaystyle\varepsilon\geq 0, and let ε00\displaystyle\varepsilon_{0}\geq 0 to be the smallest one, such that

ψλε0(X)ψ~λ(X)in Ω, andψλε0(X0)=ψ~λ(X0)for some X0Ω{ψ~λ<Q}¯.\psi_{\lambda}^{\varepsilon_{0}}(X)\leq\tilde{\psi}_{\lambda}(X)\ \text{in $\displaystyle\Omega$, and}\ \psi_{\lambda}^{\varepsilon_{0}}(X_{0})=\tilde{\psi}_{\lambda}(X_{0})\ \text{for some $\displaystyle X_{0}\in\overline{\Omega\cap\{\tilde{\psi}_{\lambda}<Q\}}$}. (4.91)

It follows from assumption (4.90) that ε0>0\displaystyle\varepsilon_{0}>0 and |X0|<+\displaystyle|X_{0}|<+\infty.

Similar to the proof of Case 1 in Step 1, we can take X0\displaystyle X_{0} be the free boundary point of ψλε0\displaystyle\psi_{\lambda}^{\varepsilon_{0}} and ψ~λ\displaystyle\tilde{\psi}_{\lambda}, applying Hopf’s lemma yields that

λ=ψλε0ν>ψ~λν=λat X0,\lambda=\frac{\partial\psi_{\lambda}^{\varepsilon_{0}}}{\partial\nu}>\frac{\partial\tilde{\psi}_{\lambda}}{\partial\nu}=\lambda\ \ \ \text{at $\displaystyle X_{0}$,}

where ν\displaystyle\nu is the outer normal vector to Γλε0Γ~λ\displaystyle\Gamma_{\lambda}^{\varepsilon_{0}}\cap\tilde{\Gamma}_{\lambda} at X0\displaystyle X_{0}. This leads a contradiction. ∎

References

  • [1] S. Agmon, A. Douglis, L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I, Comm. Pure Appl. Math., 12, 623-727, (1959).
  • [2] H. W. Alt, L. A. Caffarelli, Existence and regularity for a minimum problem with free boundary, J. Reine Angew. Math., 325, 105-144, (1981).
  • [3] H. W. Alt, L. A. Caffarelli, A. Friedman, Asymmetric jet flows, Comm. Pure Appl. Math., 35, 29-68, (1982).
  • [4] H. W. Alt, L. A. Caffarelli, A. Friedman, Jet flows with gravity, J. Reine Angew. Math., 35, 58-103, (1982).
  • [5] H. W. Alt, L. A. Caffarelli, A. Friedman, Axially symmetric jet flows, Arch. Ration. Mech. Anal., 81, 97-149, (1983).
  • [6] H. W. Alt, L. A. Caffarelli, A. Friedman, A free boundary problem for quasilinear elliptic equations, Ann. Scuola Norm. Sup. Pisa Cl. Sci., 11, 1-44, (1984).
  • [7] H. W. Alt, L. A. Caffarelli, A. Friedman, Compressible flows of jets and cavities, J. Differential Equations, 56, 82-141, (1985).
  • [8] H. W. Alt, D. Phillips, A free boundary problem for semilinear elliptic equations, J. Reine Angew. Math., 368, 63-107, (1986).
  • [9] L. A. Caffarelli, A Harnack inequality approach to the regularity of free boundaries, I. Lipschitz free boundaries are C1,α\displaystyle C^{1,\alpha}, Rev. Mat. Iberoamericana, 3, 139-162, (1987).
  • [10] L. A. Caffarelli, A Harnack inequality approach to the regularity of free boundaries. II. Flat free boundaries are Lipschitz, Comm. Pure Appl. Math., 42, 55-78, (1989).
  • [11] M. C. Cerutti, F. Ferrari, S. Salsa, Two-phase problems for linear elliptic operators with variable coefficients: Lipschitz free boundaries are C1,γ\displaystyle C^{1,\gamma}, Arch. Ration. Mech. Anal., 171, 329-348, (2004).
  • [12] J. F. Cheng, L. L. Du, Hydrodynamic jet incident on an uneven wall, Math. Models Methods Appl. Sci., 28, 771-827, (2018).
  • [13] J. F. Cheng, L. L. Du, Y, F, Wang, Two-dimensional impinging jets in hydrodynamic rotational flows, Ann. Inst. H. Poincaré Anal. Non Linéaire, 34, 1355-1386, (2017).
  • [14] J. F. Cheng, L. L. Du, Y, F, Wang, On incompressible oblique impinging jet flows, J. Differential Equations, 265, 4687-4748, (2018).
  • [15] J. F. Cheng, L. L. Du, W, Xiang, Incompressible jet flows in a de Laval nozzle with smooth detachment, Arch. Ration. Mech. Anal., 232, 1031-1072, (2019).
  • [16] J. F. Cheng, L. L. Du, Q. Zhang, Two-dimensional cavity flow in an infinitely long channel with non-zero vorticity, J. Differential Equations, 263, 4126-4155, (2017).
  • [17] L. L. Du, C. J. Xie, On subsonic Euler flows with stagnation points in two dimensional nozzles, Indiana Univ. Math. J., 63, 1499-1523, (2014).
  • [18] L. L. Du, C. J. Xie, Z. P. Xin, Steady subsonic ideal flows through an infinitely long nozzle with large vorticity, Comm. Math. Phys., 328, 327-354, (2014).
  • [19] L. C. Evans, Measure theory and fine properties of functions, Studies in Advanced Mathematics, American Mathematical Society, Providence, 1992.
  • [20] H. Federer, Geometric measure theory, Berlin-Heidelberg, New York, 1969.
  • [21] F. Ferrari, S. Salsa, Regularity of the free boundary in two-phase problems for linear elliptic operators, Adv. Math., 214, 288-322, (2007).
  • [22] F. Ferrari, S. Salsa, Subsolutions of elliptic operators in divergence form and application to two-phase free boundary problems, Bound. Value Probl., Article 57049, 21 pp., (2007).
  • [23] A. Friedman, Variational principles and free-boundary problems, Pure and Applied Mathematics, John Wiley Sons, Inc., New York, 1982.
  • [24] A. Friedman, Axially symmetric cavities in rotational flows, Comm. Partial Differential Equations, 8, 949-997, (1983).
  • [25] D. Gilbarg, The Phragme`\displaystyle\grave{\text{e}}n-Lindelöf theorem for elliptic partical differential equations, J. Rational Mech. Anal., 1, 411-417, (1952).
  • [26] D. Gilbarg, N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics in Mathematics, Springer-Verlag, Berlin, 2001.
  • [27] D. Kinderlehrer, L. Nirenberg, The smoothness of the free boundary in the one phase Stefan problem, Comm. Pure Appl. Math., 31, 257-282, (1978).
  • [28] C. B. Morrey Jr, Multiple integrals in the calculus of variations, Springer-Verlag, Berlin, 1966.
  • [29] S. Omata, A free boundary problem for nonlinear elliptic equations, Proc. Japan Acad. Ser. A Math. Sci., 66, 281-286, (1990).
  • [30] D. De Silva, Free boundary regularity for a problem with right hand side, Interfaces Free Bound., 13, 223-238, (2011).
  • [31] N. S. Trudinger, On Harnack type inequalities and their application to quasilinear elliptic equations, Comm. Pure Appl. Math., 20, 721-747, (1967).
  • [32] P. Y. Wang, Regularity of free boundaries of two-phase problems for fully nonlinear elliptic equations of second order. I. Lipschitz free boundaries are C1,α\displaystyle C^{1,\alpha}, Comm. Pure Appl. Math., 53, 799-810, (2000).
  • [33] P. Y. Wang, Regularity of free boundaries of two-phase problems for fully nonlinear elliptic equations of second order. II. Flat free boundaries are Lipschitz, Comm. Partial Differential Equations, 27, 1497-1514, (2002).
  • [34] G. Weiss, G. H. Zhang, A free boundary approach to two-dimensional steady capillary gravity water waves, Arch. Ration. Mech. Anal., 203, 747-768, (2012).
  • [35] C. J. Xie, Z. P. Xin, Existence of global steady subsonic Euler flows through infinitely long nozzle, SIAM J. Math. Anal., 42, 751-784, (2010).