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

Tangential contact between free and fixed boundaries for variational solutions to variable coefficient Bernoulli type Free boundary problems

Diego Moreira Departamento de Matemática, Universidade Federal da Ceara(Fortaleza, Brazil) [email protected]  and  Harish Shrivastava Tata Institute of Fundamental Researcher-Centre of of Applicable Mathematics [email protected],
Abstract.

In this paper, we show that given appropriate boundary data, the free boundary and the fixed boundary of minimizers of functionals of type (1.1) contact each other in a tangential fashion. We prove this result via classification of the global profiles, adapting the ideas from [14].

Keywords: Variational calculus, Bernoulli type free boundary problems, boundary behavior, Alt-Caffarelli-Friedman type minimizers.

2010 Mathematics Subjects Classification:49J05, 35B65, 35Q92, 35Q35

1. Introduction

Objective of the paper of to study the behavior of free boundary near the fixed boundary of domain, for minimizers of Bernoulli type functionals with Hölder continuous coefficients.

J(v;A,λ+,λ,Ω)=Ω(A(x)v,v+Λ(v))𝑑xJ(v;A,\lambda_{+},\lambda_{-},\Omega)=\int_{\Omega}\Big{(}\langle A(x)\nabla v,\nabla v\rangle+\Lambda(v)\Big{)}\,dx (1.1)

AA is an elliptic matrix with Hölder continuous entries, and Λ(v)=λ+χ{v>0}+λχ{v0}\Lambda(v)=\lambda_{+}\chi_{\{v>0\}}+\lambda_{-}\chi_{\{v\leq 0\}}. We prove that if the value of boundary data and its derivative at a point are equal to zero (i.e. it satisfies the (DPT) condition mentioned below), then the contact of free boundary and the fixed boundary is tangential.

Boundary interactions of free boundaries have gained significant attention in recent years. Whenever there are two medias involved, interactions of their respective diffusions can be modeled by free boundary problems. Often, free boundary of solution and fixed boundary of set come in contact. In applications, the Dam problem [3] and Jets, Wakes and Cavities [7] model phenomenas which involve understanding of free boundary and fixed boundary.

Very recently, works of Indrei [12], [13] study interactions of free boundaries and fixed boundaries for fully non-linear obstacle problems. We refer to [10] where authors shed more light into angle of contact between fixed boundary and free boundary for one phase Bernoulli problem.

As it is common by now, our strategy in this article is to classify blow up of minimizers by using the ideas from [14]. We prove that the blow up and also their positivity sets converge to a global solution in 𝒫\mathcal{P}_{\infty} as defined in [14]. In Section 2, we list the assumptions and set some notations and then in Section 3, we prove that blow ups of minimizers converge to that of global solutions (c.f. Definition 2.6). In the last section we prove our main result.

2. Setting up the problem

We consider the following class of function which we denote as 𝒫r(α,M,λ,𝒟,μ)\mathcal{P}_{r}(\alpha,M,\lambda,\mathcal{D},\mu). Before definition, we set the following notation

BR+:={xBR such that xN>0}BR:={xBR such that xN=0}.\begin{split}B_{R}^{+}:=\Big{\{}x\in B_{R}\mbox{ such that }x_{N}>0\Big{\}}\\ B_{R}^{\prime}:=\Big{\{}x\in B_{R}\mbox{ such that }x_{N}=0\Big{\}}.\end{split}

For xNx\in\mathbb{R}^{N} we denote xN1x^{\prime}\in\mathbb{R}^{N-1} as the projection of xx on the plane {xN=0}\{x_{N}=0\}, we denote the tangential gradient \nabla^{\prime} as follows

u:=(ux1,uxN1).\nabla^{\prime}u:=\Big{(}\frac{\partial u}{\partial x_{1}},...\,\frac{\partial u}{\partial x_{N-1}}\Big{)}.

We define the affine space set Hϕ1(BR+)H^{1}_{\phi}(B_{R}^{+}) as follows,

Hϕ1(BR+)={vH1(BR+):vϕH01(BR+)}.H^{1}_{\phi}(B_{R}^{+})=\left\{v\in H^{1}(B_{R}^{+})\,:\,v-\phi\in H_{0}^{1}(B_{R}^{+})\right\}. (2.1)

For a given function vH1(B2+)v\in H^{1}(B_{2}^{+}), we denote F(v)F(v) as F(v):={v>0}F(v):=\partial\{v>0\} and IdId is the notation for N×NN\times N identity matrix.

Definition 2.1.

A function uH1(B2/r+)u\in H^{1}(B_{2/r}^{+}) is said to belong to the class 𝒫r(α,M,λ±,𝒟,μ)\mathcal{P}_{r}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu) if there exists ACα(B2/r+)N×NA\in C^{\alpha}(B_{2/r}^{+})^{N\times N}, ϕC1,α(B2/r+)\phi\in C^{1,\alpha}(B_{2/r}^{+}), λ±>0\lambda_{\pm}>0, 0<μ<10<\mu<1 and 𝒟>0\mathcal{D}>0 such that

  1. ((P1))

    AL(B2/r+)M\|A\|_{L^{\infty}(B_{2/r}^{+})}\leq M, ϕL(B2/r+)M\|\nabla\phi\|_{L^{\infty}(B_{2/r}^{+})}\leq M, [A]Cα(B2/r+),[ϕ]C1,α(B2/r+)rαM[A]_{C^{\alpha}(B_{2/r}^{+})},[\nabla\phi]_{C^{1,\alpha}(B_{2/r}^{+})}\leq r^{\alpha}M
    and |ϕ(x)|Mr1+α|x|1+α|\phi(x^{\prime})|\leq Mr^{1+\alpha}|x^{\prime}|^{1+\alpha} ((xB2/rx^{\prime}\in B_{2/r}^{\prime})). ϕ\phi satisfies the following Degenerate Phase Transition condition (DPT) mentioned below.

    xB2/r such that ϕ(x)=0, then |ϕ(x)|=0.\mbox{$\forall x^{\prime}\in B_{2/r}^{\prime}$ such that $\phi(x^{\prime})=0$, then $|\nabla^{\prime}\phi(x^{\prime})|=0$}. (DPT)
  2. ((P2))

    A(0)=IdA(0)=Id, μ|ξ|2A(x)ξ,ξ1μ|ξ|2\mu|\xi|^{2}\leq\langle A(x)\xi,\xi\rangle\leq\frac{1}{\mu}|\xi|^{2} for all xB2/r+x\in B_{2/r}^{+} and ξN\xi\in\mathbb{R}^{N}.

  3. ((P3))

    0<λ<λ+0<\lambda_{-}<\lambda_{+}.

  4. ((P4))

    uu minimizes J(;A,λ+,λ,B2/r+)J(\cdot;A,\lambda_{+},\lambda_{-},B_{2/r}^{+}) ((c.f. (1.1))) that is for every uvH01(B2/r+)u-v\in H_{0}^{1}(B_{2/r}^{+})

    B2/r+(A(x)u,u+Λ(u))𝑑xB2/r+(A(x)v,v+Λ(v))𝑑x\int_{B_{2/r}^{+}}\Big{(}\langle A(x)\nabla u,\nabla u\rangle+\Lambda(u)\Big{)}\,dx\leq\int_{B_{2/r}^{+}}\Big{(}\langle A(x)\nabla v,\nabla v\rangle+\Lambda(v)\Big{)}\,dx\;\;

    (Λ(s)=λ+χ{s>0}+λχ{s0})(\Lambda(s)=\lambda_{+}\chi_{\{s>0\}}+\lambda_{-}\chi_{\{s\leq 0\}}) and 0F(u)B2/r+¯0\in F(u)\cap\overline{B_{2/r}^{+}}.

  5. ((P5))

    uHϕ1(B2/r+)u\in H_{\phi}^{1}(B_{2/r}^{+}).

  6. ((P6))

    There exists 0<r00<r_{0} such that for all 0<ρr00<\rho\leq r_{0} we have

    |Bρ+(0){u>0}||Bρ+(0)|>𝒟.\frac{|B_{\rho}^{+}(0)\cap\{u>0\}|}{|B_{\rho}^{+}(0)|}>\mathcal{D}. (2.2)
Remark 2.2.

In fact, the functions u𝒫r(α,M,λ±,𝒟,μ)u\in\mathcal{P}_{r}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu) carry more regularity than being only a Sobolev function. They are Hölder continuous in (B2/r+¯)(\overline{B_{2/r}^{+}}) (c.f. Lemma 3.2).

In the absence of ambiguity on values of α,M,λ±,𝒟,μ\alpha,M,\lambda_{\pm},\mathcal{D},\mu we use the notation 𝒫r\mathcal{P}_{r} in place of 𝒫r(α,M,λ±,𝒟,μ)\mathcal{P}_{r}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu). If ϕC1,α(B2+)\phi\in C^{1,\alpha}(B_{2}^{+}) satisfies (DPT), from [15, Lemma 10.1], we know that ϕ±|B2C1,α(B2)\phi^{\pm}\Big{|}_{B_{2}^{\prime}}\in C^{1,\alpha}(B_{2}^{\prime}) and also

ϕ±C1,α(B2)ϕC1,α(B2).\|\phi^{\pm}\|_{C^{1,\alpha}(B_{2}^{\prime})}\leq\|\phi\|_{C^{1,\alpha}(B_{2}^{\prime})}.

Given vH1(BR+)v\in H^{1}(B_{R}^{+}) and r>0r>0, we define the blow-up vrH1(BR/r+)v_{r}\in H^{1}(B_{R/r}^{+}) as follows

vr(x):=1rv(rx).v_{r}(x):=\frac{1}{r}v(rx). (2.3)

For the coefficient matrix AA, Ar(x)A^{r}(x) is defined as follows

Ar(x):=A(rx).A^{r}(x):=A(rx). (2.4)

One can check that if u𝒫1(α,M,λ±,𝒟,μ)u\in\mathcal{P}_{1}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu), then ur𝒫r(α,M,λ±,𝒟,μ)u_{r}\in\mathcal{P}_{r}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu). Indeed if u𝒫1u\in\mathcal{P}_{1} and uu minimizer the functional JJ (c.f. ((P4)))

J(v;A,λ+,λ,B2+):=B2+(A(x)v,v+Λ(v))𝑑x,(Λ(s)=λ+χ{s>0}+λχ{s0})J(v;A,\lambda_{+},\lambda_{-},B_{2}^{+}):=\int_{B_{2}^{+}}\Big{(}\langle A(x)\nabla v,\nabla v\rangle+\Lambda(v)\Big{)}\,dx,\qquad(\Lambda(s)=\lambda_{+}\chi_{\{s>0\}}+\lambda_{-}\chi_{\{s\leq 0\}})

with boundary data ϕC1,α(B2+)\phi\in C^{1,\alpha}(B_{2}^{+}) (i.e. uHϕ1(B2+)u\in H_{\phi}^{1}(B_{2}^{+})). Then by simple change of variables we can check that urHϕr1(B2/r+)u_{r}\in H^{1}_{\phi_{r}}(B_{2/r}^{+}) (this verifies ((P5))) and uru_{r} minimizes

J(v;Ar,λ+,λ,B2/r+):=B2/r+(Ar(x)v,v+Λ(v))𝑑x,(Λ(s)=λ+χ{s>0}+λχ{s0}).J(v;A^{r},\lambda_{+},\lambda_{-},B_{2/r}^{+}):=\int_{B_{2/r}^{+}}\Big{(}\langle A^{r}(x)\nabla v,\nabla v\rangle+\Lambda(v)\Big{)}\,dx,\qquad(\Lambda(s)=\lambda_{+}\chi_{\{s>0\}}+\lambda_{-}\chi_{\{s\leq 0\}}).

Moreover, if AA and ϕ\phi satisfy the conditions ((P1)), ((P2)) for r=1r=1, then ArA^{r} and ϕr\phi_{r} satisfy ((P1)), ((P2)) for rr. ((P3)) and ((P6)) remains invariant under the change variables. Therefore ur𝒫ru_{r}\in\mathcal{P}_{r}.

In order to study the blow-up limits (limr0ur\lim_{r\to 0}u_{r}) of functions u𝒫1(α,M,λ±,𝒟,μ)u\in\mathcal{P}_{1}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu), we define a class of global solutions 𝒫(C,λ±)\mathcal{P}_{\infty}(C,\lambda_{\pm}). Let us set the following notation before giving the definition

Π:={xN:xN=0}.\Pi:=\{x\in\mathbb{R}^{N}\,:\,x_{N}=0\}.
Definition 2.3 (Global solution).

We say that uH1(+N)u\in H^{1}(\mathbb{R}^{N}_{+}) belongs to the class 𝒫(C,λ±)\mathcal{P}_{\infty}(C,\lambda_{\pm}), that is, uu is a global solution if there exists C>0C>0 and 0<λ+<λ0<\lambda_{+}<\lambda_{-} such that

  1. (G1)

    |u(x)|C|x||u(x)|\leq C|x|, for all x+Nx\in\mathbb{R}^{N}_{+},

  2. (G2)

    uu is continuous up to the boundary Π\Pi,

  3. (G3)

    u=0u=0 on Π\Pi,

  4. (G4)

    and for every ball Br(x0)B_{r}(x_{0}), uu is a minimizer of J(;Id,λ+,λ,Br(x0)+N)J(\cdot;{Id,\lambda_{+},\lambda_{-},B_{r}(x_{0})\cap\mathbb{R}^{N}_{+}}) ((c.f. (1.1))), that is

    Br(x0)+N(|u|2+Λ(u))𝑑xBr(x0)+N(|v|2+Λ(v))𝑑x.\int_{B_{r}(x_{0})\cap\mathbb{R}^{N}_{+}}\Big{(}|\nabla u|^{2}+\Lambda(u)\Big{)}\,dx\leq\int_{B_{r}(x_{0})\cap\mathbb{R}^{N}_{+}}\Big{(}|\nabla v|^{2}+\Lambda(v)\Big{)}\,dx.

    Here (Λ(s)=λ+χ{s>0}+λχ{s0})(\Lambda(s)=\lambda_{+}\chi_{\{s>0\}}+\lambda_{-}\chi_{\{s\leq 0\}}) and for every vH1(Br(x0)+N)v\in H^{1}(B_{r}(x_{0})\cap\mathbb{R}^{N}_{+}) such that uvH01(Br(x0)+N)u-v\in H_{0}^{1}(B_{r}(x_{0})\cap\mathbb{R}^{N}_{+}).

Our main result intends to show that for a minimizer uu of J(;A,λ+,λ,B2+)J(\cdot;A,\lambda_{+},\lambda_{-},B_{2}^{+}) with A,λ±A,\lambda_{\pm} and uu satisfying the properties ((P1))-((P6)), the free boundary of every such minimizer touches the flat part of fixed boundary tangentially at the origin. For this, we prove that as we approach closer and closer to the origin, the free boundary points cannot lie outside any cone which is perpendicular to the flat boundary and has its tip at the origin. The main result in this paper is stated below.

Theorem 2.4.

There exists a constant ρ0\rho_{0} and a modulus of continuity σ\sigma such that if

u𝒫1(α,M,λ±,𝒟,μ)u\in\mathcal{P}_{1}(\alpha,M,\lambda_{\pm},\mathcal{D},\mu)

then

F(u)Bρ0+{x:xNσ(|x|)|x|}F(u)\cap B_{\rho_{0}}^{+}\subset\{x\,:\,x_{N}\leq\sigma(|x|)|x|\}

Here σ\sigma depends only on α,M,λ±,𝒟,μ\alpha,M,\lambda_{\pm},\mathcal{D},\mu.

3. Blow-up analysis

The following is a classical result (c.f. [1, Remark 4.2]) , we present the proof for the case of variable coefficients.

Lemma 3.1.

Given a strictly elliptic matrix and bounded A(x)A(x) and a non-negative continuous function ww such that div(A(x)w)=0\operatorname{div}(A(x)\nabla w)=0 in {w>0}B2+\{w>0\}\cap B_{2}^{+}, then wHloc1(B2+)w\in H^{1}_{loc}(B_{2}^{+}) and div(A(x)w)0\operatorname{div}(A(x)\nabla w)\geq 0 in weak sense in B2+B_{2}^{+}.

Proof.

Let DB2+D\Subset B_{2}^{+} and ηCc(B2+)\eta\in C_{c}^{\infty}(B_{2}^{+}) be cutoff function for DD. That is ηCc(B2+)\eta\in C_{c}^{\infty}(B_{2}^{+}) be such that

η(x)={1in D0on B2+.\eta(x)=\begin{cases}1\;\;\mbox{in $D$}\\ 0\;\;\mbox{on $\partial B_{2}^{+}$}.\end{cases}

Since div(A(x)w)=0\operatorname{div}(A(x)\nabla w)=0 in {w>0}\{w>0\}, we have

0=B2+A(x)w,((wε)+η2)𝑑x=B2+{w>ε}η2A(x)w,w𝑑x+B2+{w>ε}wA(x)w,η2𝑑x+εB2+{w>ε}A(x)w,η2𝑑x\begin{split}0&=\int_{B_{2}^{+}}\langle A(x)\nabla w,\nabla\big{(}(w-{\varepsilon})^{+}\eta^{2}\big{)}\rangle\,dx\\ &=\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\eta^{2}\langle A(x)\nabla w,\nabla w\rangle\,dx+\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}w\langle A(x)\nabla w,\nabla\eta^{2}\rangle\,dx+{\varepsilon}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\langle A(x)\nabla w,\nabla\eta^{2}\rangle\,dx\end{split}

which implies

B2+{w>ε}A(x)w,wη2𝑑xB2+{w>ε}|wA(x)w,η2|dx+εB2+{w>ε}|A(x)w,η2|𝑑x.\begin{split}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\langle A(x)\nabla w,\nabla w\rangle\eta^{2}\,dx\leq\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}&\Big{|}w\langle A(x)\nabla w,\nabla\eta^{2}\rangle\Big{|}\,dx\\ &+{\varepsilon}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\Big{|}\langle A(x)\nabla w,\nabla\eta^{2}\rangle\Big{|}\,dx.\end{split}

By the choice of η\eta and ellipticity of the matrix AA, we obtain using Young’s inequality

μB2+{w>ε}|w|2η2𝑑xμ|B2+{w>ε}ηw|w||η|𝑑x|+εμ|B2+{w>ε}η|w||η|𝑑x|C1(μ)[1δB2+{w>ε}w2|η|2dx+δB2+{w>ε}η2|w|2dx+δεB2+{w>ε}η2|w|2dx+εδB2+{w>ε}|η|2dx]\begin{split}\mu\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}|\nabla w|^{2}\eta^{2}\,dx&\leq\mu\Big{|}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\eta w|\nabla w||\nabla\eta|\,dx\Big{|}+{\varepsilon}\mu\Big{|}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\eta|\nabla w||\nabla\eta|\,dx\Big{|}\\ &\leq C_{1}(\mu)\Bigg{[}\frac{1}{\delta}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}w^{2}|\nabla\eta|^{2}\,dx+\delta\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\eta^{2}|\nabla w|^{2}\,dx\\ &\qquad\qquad\qquad\qquad+\delta{\varepsilon}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}\eta^{2}|\nabla w|^{2}\,dx+\frac{{\varepsilon}}{\delta}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}|\nabla\eta|^{2}\,dx\Bigg{]}\end{split}

after choosing of δ>0\delta>0 very small and rearranging the terms in the equation above, since η=1\eta=1 in DD we finally get

D{w>ε}|w|2𝑑xB2+{w>ε}|w|2η2𝑑xC(μ)[B2+{w>ε}w2|n|2𝑑x+B2+{w>ε}|η|2𝑑x]\begin{split}\int_{D\cap\{w>{\varepsilon}\}}|\nabla w|^{2}\,dx&\leq\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}|\nabla w|^{2}\eta^{2}\,dx\\ &\leq C(\mu)\Bigg{[}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}w^{2}|\nabla n|^{2}\,dx+\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}|\nabla\eta|^{2}\,dx\Bigg{]}\end{split}

As ε0{\varepsilon}\to 0, we obtain

D|w|2𝑑x={w>0}D|w|2𝑑x=limε0{w>ε}D|w|2C(μ)limε0[B2+{w>ε}w2|n|2𝑑x+B2+{w>ε}|η|2𝑑x]C(μ,D)[B2+supp(η)w2𝑑x+1].\begin{split}\int_{D}|\nabla w|^{2}\,dx=\int_{\{w>0\}\cap D}|\nabla w|^{2}\,dx&=\lim_{{\varepsilon}\to 0}\int_{\{w>{\varepsilon}\}\cap D}|\nabla w|^{2}\,\\ &\leq C(\mu)\lim_{{\varepsilon}\to 0}\Bigg{[}\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}w^{2}|\nabla n|^{2}\,dx+\int_{B_{2}^{+}\cap\{w>{\varepsilon}\}}|\nabla\eta|^{2}\,dx\Bigg{]}\\ &\leq C(\mu,D)\Bigg{[}\int_{B_{2}^{+}\cap\operatorname{supp}(\eta)}w^{2}\,dx+1\Bigg{]}.\end{split}

Since wC(B2+)w\in C(B_{2}^{+}) therefore, ww is uniformly bounded in supp(η)\operatorname{supp}(\eta) and therefore

D(|w|2+|w2|)𝑑xC(μ,D)[B2+supp(η)w2𝑑x+1]C(μ,D,wL(supp(η))).\int_{D}\Big{(}|\nabla w|^{2}+|w^{2}|\Big{)}\,dx\leq C(\mu,D)\Bigg{[}\int_{B_{2}^{+}\cap\operatorname{supp}(\eta)}w^{2}\,dx+1\Bigg{]}\leq C(\mu,D,\|w\|_{L^{\infty}(\operatorname{supp}(\eta))}).

Now, for 0φCc(B2+)0\leq\varphi\in C_{c}^{\infty}(B_{2}^{+}) consider the test function

v=φ(1(min(2wε,1))+).v=\varphi\Big{(}1-\big{(}\min(2-\frac{w}{{\varepsilon}},1)\big{)}^{+}\Big{)}.
B2+A(x)w,φdx=B2+A(x)w,(φ((2wε)1)+))dx\int_{B_{2}^{+}}\langle A(x)\nabla w,\nabla\varphi\rangle\,dx=\int_{B_{2}^{+}}\langle A(x)\nabla w,\nabla\big{(}\varphi((2-\frac{w}{{\varepsilon}})\wedge 1)^{+})\big{)}\rangle\,dx

We can easily check that v0v\geq 0 in B2+B_{2}^{+} and vH01(B2+)v\in H_{0}^{1}(B_{2}^{+}), in particular

φ((2wε)1)+={φx{wε},φ(2wε)x{ε<w2ε},0x{w>2ε}.\varphi\Big{(}(2-\frac{w}{{\varepsilon}})\wedge 1\Big{)}^{+}=\begin{cases}\varphi\;\;\;\qquad\qquad\;\;\qquad\qquad x\in\{w\leq{\varepsilon}\},\\ \varphi\cdot\Big{(}2-\frac{w}{{\varepsilon}}\Big{)}\;\qquad\;\;\;\qquad x\in\{{\varepsilon}<w\leq 2{\varepsilon}\},\\ 0\;\;\;\;\qquad\qquad\qquad\qquad\;\;x\in\{w>2{\varepsilon}\}.\end{cases}

Therefore, we have

B2+A(x)w,φ𝑑x=B2+A(x)w,(φ((2wε)1)+))dx=B2+{wε}A(x)w,φ𝑑x+B2+{ε<w2ε}A(x)w,(φ(2wε))𝑑x=B2+{wε}A(x)w,φ𝑑x+2B2+{ε<w2ε}A(x)w,φ𝑑x2εB2+{ε<w2ε}A(x)w,(wφ)𝑑xC(μ)B2+{ε<u2ε}|w||φ|𝑑x+2εB2+{ε<w2ε}wA(x)w,φ𝑑x2εB2+{ε<w2ε}φA(x)w,w𝑑xC(μ)B2+{ε<w2ε}|w||φ|𝑑x\begin{split}\int_{B_{2}^{+}}\langle A(x)\nabla w,\nabla\varphi\rangle\,dx&=\int_{B_{2}^{+}}\langle A(x)\nabla w,\nabla\big{(}\varphi((2-\frac{w}{{\varepsilon}})\wedge 1)^{+})\big{)}\rangle\,dx\\ &=\int_{B_{2}^{+}\cap\{w\leq{\varepsilon}\}}\langle A(x)\nabla w,\nabla\varphi\rangle\,dx+\int_{B_{2}^{+}\cap\{{\varepsilon}<w\leq 2{\varepsilon}\}}\langle A(x)\nabla w,\nabla\Big{(}\varphi\big{(}2-\frac{w}{{\varepsilon}}\big{)}\Big{)}\rangle\,dx\\ &=\int_{B_{2}^{+}\cap\{w\leq{\varepsilon}\}}\langle A(x)\nabla w,\nabla\varphi\rangle\,dx+2\int_{B_{2}^{+}\cap\{{\varepsilon}<w\leq 2{\varepsilon}\}}\langle A(x)\nabla w,\nabla\varphi\rangle\,dx\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\frac{2}{{\varepsilon}}\int_{B_{2}^{+}\cap\{{\varepsilon}<w\leq 2{\varepsilon}\}}\langle A(x)\nabla w,\nabla(w\varphi)\rangle\,dx\\ &\leq C(\mu)\int_{B_{2}^{+}\cap\{{\varepsilon}<u\leq 2{\varepsilon}\}}|\nabla w||\nabla\varphi|\,dx+\frac{2}{{\varepsilon}}\int_{B_{2}^{+}\cap\{{\varepsilon}<w\leq 2{\varepsilon}\}}w\langle A(x)\nabla w,\nabla\varphi\rangle\,dx\\ &\qquad\qquad\qquad\qquad\qquad\qquad-\frac{2}{{\varepsilon}}\int_{B_{2}^{+}\cap\{{\varepsilon}<w\leq 2{\varepsilon}\}}\varphi\langle A(x)\nabla w,\nabla w\rangle\,dx\\ \leq&C(\mu)\int_{B_{2}^{+}\cap\{{\varepsilon}<w\leq 2{\varepsilon}\}}|\nabla w||\nabla\varphi|\,dx\end{split}

The last term goes to zero as ε0{\varepsilon}\to 0. Therefore, we can say that

B2+A(x)w,φ𝑑x0.\int_{B_{2}^{+}}\langle A(x)\nabla w,\nabla\varphi\rangle\,dx\leq 0.

This concludes the proof. ∎

Lemma 3.2 (Hölder continuity).

If u𝒫1u\in\mathcal{P}_{1} then uC0,α0(B2+¯)u\in C^{0,\alpha_{0}}(\overline{B_{2}^{+}}) for some 0<α0<10<\alpha_{0}<1. In fact,

uCα0(B2+)C(μ,λ±)uL(B2+).\|u\|_{C^{\alpha_{0}}(B_{2}^{+})}\leq C(\mu,\lambda_{\pm})\|u\|_{L^{\infty}(B_{2}^{+})}.
Proof.

The functional J(;A,λ+,λ,B2+)J(\cdot;A,\lambda_{+},\lambda_{-},B_{2}^{+}) satisfy the hypothesis of [9, Theorem 7.3] and ϕC1,α(B2+¯)\phi\in C^{1,\alpha}(\overline{B_{2}^{+}}), therefore Lemma 3.2 follows from the arguments in [9, Section 7.8]. ∎

Remark 3.3.

Since every function u𝒫1u\in\mathcal{P}_{1} is continuous. Therefore, the positivity set {u>0}\{u>0\} is an open set.

Corollary 3.4.

If u𝒫1u\in\mathcal{P}_{1}, then u±u^{\pm} are AA-subharmonic.

Proof.

The claim follows directly from Lemma 3.2 and Lemma 3.1. ∎

Lemma 3.5.

If u𝒫1u\in\mathcal{P}_{1} then

|u(x)|C(μ)M|x| in B1+.|u(x)|\leq C(\mu)M\,|x|\mbox{ in $B_{1}^{+}$.} (3.1)

.

Proof.

Let ww be such that

{div(A(x)w)=0,in B2+w=ϕ+,in B2+.\begin{cases}\operatorname{div}(A(x)w)=0,\qquad\mbox{in $B_{2}^{+}$}\\ w=\phi^{+},\qquad\qquad\qquad\mbox{in $\partial B_{2}^{+}$}.\end{cases}

Since uu is AA-subharmonic in B2+B_{2}^{+} (c.f. Corollary 3.4), by maximum principle, if xB1+x\in B_{1}^{+} we have

u+(x)w(x)w(x)w(x)+w(x)wL(B1+)|xx|+|ϕ+(x)|=wL(B1+)xN+|ϕ+(x)|(wL(B1+)+M)|x|xB1+\begin{split}u^{+}(x)\leq w(x)&\leq w(x)-w(x^{\prime})+w(x^{\prime})\\ &\leq\|\nabla w\|_{L^{\infty}(B_{1}^{+})}|x-x^{\prime}|+|\phi^{+}(x^{\prime})|\\ &=\|\nabla w\|_{L^{\infty}(B_{1}^{+})}x_{N}+|\phi^{+}(x^{\prime})|\\ &\leq(\|\nabla w\|_{L^{\infty}(B_{1}^{+})}+M)|x|\;\forall x\in B_{1}^{+}\end{split} (3.2)

in the last inequality, we have used ((P1)). Now, we prove that the term wL(B1+)\|\nabla w\|_{L^{\infty}(B_{1}^{+})} is uniformly bounded.

From [4, Theorem 2], we have

wL(B1+)C(μ)[wL(B2+)+ϕ+C1,α(B2)].\|\nabla w\|_{L^{\infty}(B_{1}^{+})}\leq{C(\mu)}\Big{[}\|w\|_{L^{\infty}(B_{2}^{+})}+\|\phi^{+}\|_{C^{1,\alpha}(B_{2}^{\prime})}\Big{]}. (3.3)

By comparison principle, wL(B2+)=wL(B2+)=ϕ+L(B2+)M\|w\|_{L^{\infty}(B_{2}^{+})}=\|w\|_{L^{\infty}(\partial B_{2}^{+})}=\|\phi^{+}\|_{L^{\infty}(\partial B_{2}^{+})}\leq M. Plugging this information in (3.3)

wL(B1+)C(μ)M\|\nabla w\|_{L^{\infty}(B_{1}^{+})}\leq C(\mu)M

and then using (3.2), we obtain

u+(x)C(μ)M|x|xB1+.u^{+}(x)\leq C(\mu)M|x|\;\forall\;x\in B_{1}^{+}. (3.4)

And analogously,

u(x)C(μ)M|x|xB1+.u^{-}(x)\leq C(\mu)M|x|\;\forall\;x\in B_{1}^{+}. (3.5)

By adding (3.4) and (3.5), we prove (3.1). ∎

Remark 3.6.

We can check that for every u𝒫1u\in\mathcal{P}_{1}, ur𝒫ru_{r}\in\mathcal{P}_{r} and uru_{r} is ArA^{r}-subharmonic and satisfies (3.1) in B1/r+B_{1/r}^{+}. That is

|ur(x)|C(μ)M|x|,xB1/r+.|u_{r}(x)|\leq C(\mu)M\,|x|,\,\,x\in B_{1/r}^{+}.
Lemma 3.7 (Uniform bounds in H1(B1+)H^{1}(B_{1}^{+}) norm).

Let u𝒫1u\in\mathcal{P}_{1}. Then for R>0R>0 such that 2R2r2R\leq\frac{2}{r}, we have

BR+|ur|2𝑑xC(N,λ,μ,R,M).\int_{B_{R}^{+}}|\nabla u_{r}|^{2}\,dx\leq C(N,\lambda,\mu,R,M).
Proof.

Since ur𝒫ru_{r}\in\mathcal{P}_{r}, from ((P4)) we can say that uru_{r} is a minimizer of J(;Ar,λ±,B2R+)J(\cdot;A^{r},\lambda_{\pm},B_{2R}^{+}) with boundary data ϕr\phi_{r}. Here ArA^{r} and ϕr\phi_{r} satisfy the conditions ((P1)) and ((P2)). Precisely speaking, uru_{r} is minimizer of the following functional

J(v;Ar,λ±,B2R+):=B2R+(Ar(x)v,v+Λ(v))𝑑xJ(v;A^{r},\lambda_{\pm},B_{2R}^{+}):=\int_{B_{2R}^{+}}\Big{(}\langle A^{r}(x)\nabla v,\nabla v\rangle+\Lambda(v)\Big{)}\,dx

here (Λ(v)=λ+χ{v>0}+λχ{v0})\big{(}\Lambda(v)=\lambda_{+}\chi_{\{v>0\}}+\lambda_{-}\chi_{\{v\leq 0\}}\big{)}. Consider hH1(B2R+))h\in H^{1}(B_{2R}^{+})) be a harmonic replacement

{div(Ar(x)h)=0in B2R+hurH01(B2R+).\begin{cases}\operatorname{div}(A^{r}(x)\nabla h)=0\qquad\mbox{in $B_{2R}^{+}$}\\ h-u_{r}\in H_{0}^{1}(B_{2R}^{+}).\end{cases}

in other words, hh is the minimizer of B2R+Ar(x)h,h𝑑x\int_{B_{2R}^{+}}\langle A^{r}(x)\nabla h,\nabla h\rangle\,dx, in the set Hur1(B2R+)H_{u_{r}}^{1}(B_{2R}^{+}).

From minimality of uru_{r} and the choice of hh, we have

B2R+Ar(x)(urh),(urh)dx=B2R+Ar(x)(urh),(ur+h2h)dx=B2R+Ar(x)(urh),(ur+h)𝑑x2B2R+Ar(x)(urh),h𝑑x=B2R+(Ar(x)ur,urAr(x)h,h)𝑑x(since h is Ar-harmonic in BR+)B2R+(Λ(h)Λ(ur))𝑑xC(N,λ,R).\begin{split}\int_{B_{2R}^{+}}\langle A^{r}(x)\nabla(u_{r}-h),\nabla&(u_{r}-h)\rangle\,dx=\int_{B_{2R}^{+}}\langle A^{r}(x)\nabla(u_{r}-h),\nabla(u_{r}+h-2h)\rangle\,dx\\ &=\int_{B_{2R}^{+}}\langle A^{r}(x)\nabla(u_{r}-h),\nabla(u_{r}+h)\rangle\,dx-2\int_{B_{2R}^{+}}\langle A^{r}(x)\nabla(u_{r}-h),\nabla h\rangle\,dx\\ &=\int_{B_{2R}^{+}}\Big{(}\langle A^{r}(x)\nabla u_{r},\nabla u_{r}\rangle-\langle A^{r}(x)\nabla h,\nabla h\rangle\Big{)}\,dx\;\;\;\mbox{(since $h$ is $A^{r}$-harmonic in $B_{R}^{+}$)}\\ &\leq\int_{B_{2R}^{+}}\Big{(}\Lambda(h)-\Lambda(u_{r})\Big{)}\,dx\leq C(N,\lambda,R).\end{split}

We use ellipticity of AA, we get

BR+|(urh)|2𝑑xB2R+|(urh)|2𝑑xC(N,λ,μ,R)\int_{B_{R}^{+}}|\nabla(u_{r}-h)|^{2}\,dx\leq\int_{B_{2R}^{+}}|\nabla(u_{r}-h)|^{2}\,dx\leq C(N,\lambda,\mu,R)

expanding the left hand side, we get

BR+|ur|2𝑑xBR+|ur|2𝑑x+BR+|h|2𝑑xC(N,λ,μ,R)+2BR+urhdxC(N,λ,μ,R)+εBR+|ur|2𝑑x+1εBR+|h|2𝑑x\begin{split}\int_{B_{R}^{+}}|\nabla u_{r}|^{2}\,dx\leq\int_{B_{R}^{+}}|\nabla u_{r}|^{2}\,dx+\int_{B_{R}^{+}}|\nabla h|^{2}\,dx&\leq C(N,\lambda,\mu,R)+2\int_{B_{R}^{+}}\nabla u_{r}\cdot\nabla h\,dx\\ &\leq C(N,\lambda,\mu,R)+{\varepsilon}\int_{B_{R}^{+}}|\nabla u_{r}|^{2}\,dx+\frac{1}{{\varepsilon}}\int_{B_{R}^{+}}|\nabla h|^{2}\,dx\end{split}

by choosing ε=18{\varepsilon}=\frac{1}{8} we are left with the following,

BR+|ur|2𝑑xC(N,λ,μ,R)(1+BR+|h|2𝑑x).\int_{B_{R}^{+}}|\nabla u_{r}|^{2}\,dx\leq C(N,\lambda,\mu,R)\Big{(}1+\int_{B_{R}^{+}}|\nabla h|^{2}\,dx\Big{)}.

From [4, Theorem 2], hL(BR+)C(μ,M)\|\nabla h\|_{L^{\infty}(B_{R}^{+})}\leq C(\mu,M). Thus we obtain a uniform bound on BR+|h|2𝑑x\int_{B_{R}^{+}}|\nabla h|^{2}\,dx. ∎

Lemma 3.8 (Compactness).

Let rj0+r_{j}\to 0^{+}, and a sequence {vj}𝒫1\{v_{j}\}\in\mathcal{P}_{1}. Then the blow-ups uj:=(vj)rju_{j}:=(v_{j})_{r_{j}} ((as defined in (2.3))) converges ((up to subsequece)) uniformly in BR+B_{R}^{+} and weakly in H1(BR+)H^{1}(B_{R}^{+}) to some limit for any R>0R>0. Moreover, if u0u_{0} is such a limit of uju_{j} in the above mentioned topologies, then u0u_{0} belongs to 𝒫\mathcal{P}_{\infty}.

Proof.

We fix R>0R>0, since vj𝒫1v_{j}\in\mathcal{P}_{1}, therefore uj𝒫rju_{j}\in\mathcal{P}_{r_{j}} and as argued in the proof of previous Lemma, the functions uju_{j} are minimizers of the functional J(;Aj,λ±,BR+)J(\cdot;A_{j},\lambda_{\pm},B_{R}^{+}) for jj sufficiently large that R<1rjR<\frac{1}{r_{j}}. We set the notation for the functional JjJ_{j} as

Jj(v):=J(v;Aj,λ±,BR+)=BR+(Aj(x)v,v+Λ(v))𝑑x,vHuj1(BR+).J_{j}(v):=J(v;A_{j},\lambda_{\pm},B_{R}^{+})=\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla v,\nabla v\rangle+\Lambda(v)\Big{)}\,dx,\qquad v\in H_{u_{j}}^{1}(B_{R}^{+}). (3.6)

We also denote the boundary values for uj𝒫rju_{j}\in\mathcal{P}_{r_{j}} as ϕj\phi_{j}. Here the sequences AjCα(B2/rj+)N×NA_{j}\in{C^{\alpha}(B_{2/r_{j}}^{+})}^{N\times N} and ϕjC1,α(B2/rj)\phi_{j}\in C^{1,\alpha}(B_{2/r_{j}}) satisfy the condition ((P1)), ((P2)) with r=rjr=r_{j}, Λ(v):=λ+χ{v>0}+λχ{v0}\Lambda(v):=\lambda_{+}\chi_{\{v>0\}}+\lambda_{-}\chi_{\{v\leq 0\}}. We set the following notation for the functional J0J_{0}

J0(v;BR+):=BR+|v|2+Λ(v)dx.J_{0}(v;B_{R}^{+}):=\int_{B_{R}^{+}}|\nabla v|^{2}+\Lambda(v)\,dx. (3.7)

From Lemma 3.2, we know that ujCα0(B2/rj+)¯u_{j}\in C^{\alpha_{0}}(\overline{B_{2/r_{j}}^{+})} which implies Cα0(BR+)¯C^{\alpha_{0}}(\overline{B_{R}^{+})}. In particular ujC0,α0(BR+)C(μ,λ±)\|u_{j}\|_{C^{0,\alpha_{0}}(B_{R}^{+})}\leq C(\mu,\lambda_{\pm}). Hence, uju_{j} is a uniformly bounded and equicontinuous sequence in BR+¯\overline{B_{R}^{+}}, we can apply Arzela Ascoli theorem to show that uju_{j} uniformly converges to a function u0C0,α0(BR+¯)u_{0}\in C^{0,\alpha_{0}}(\overline{B_{R}^{+}}).

Since uj=ϕju_{j}=\phi_{j} on BRB_{R}^{\prime}, from ((P1)) we have |ϕj(x)|Mrj1+α|x|1+α|\phi_{j}(x)|\leq Mr_{j}^{1+\alpha}|x|^{1+\alpha} for xBRx\in B_{R}^{\prime}, therefore
|ϕj(x)|C(M,α)rj1+αR1+α|\phi_{j}(x)|\leq C(M,\alpha)r_{j}^{1+\alpha}R^{1+\alpha}. Hence ϕj0\phi_{j}\to 0 uniformly on BRB_{R}^{\prime}. We have

u0=limjuj=limjϕj=0 on BR.u_{0}=\lim_{j\to\infty}u_{j}=\lim_{j\to\infty}\phi_{j}=0\mbox{ on $B_{R}^{\prime}$.}

Thus u0u_{0} satisfies (G2) and (G3) inside the domain BR+¯\overline{B_{R}^{+}}. Also, from Lemma 3.7 we have

BR+|uj|2𝑑xC(N,λ±,μ,R,M,α)\int_{B_{R}^{+}}|\nabla u_{j}|^{2}\,dx\leq C(N,\lambda_{\pm},\mu,R,M,\alpha) (3.8)

then, by the linear growth condition (c.f. Remark 3.6), uju_{j} also satisfies

|uj(x)|C(μ,α)M|x|xBR+.|u_{j}(x)|\leq C(\mu,\alpha)M|x|\;\;x\in B_{R}^{+}. (3.9)

Hence, passing to the limit, we have |u0(x)|C(μ,α)M|x|,xBR+|u_{0}(x)|\leq C(\mu,\alpha)M|x|,\;\forall x\in B_{R}^{+}. In other words u0u_{0}, satisfies (G1) in BR+¯\overline{B_{R}^{+}}. Moreover, we have

BR+|uj|2𝑑xC(μ,α,M)BR+|x|2𝑑xC(μ,α,M,N,R).\int_{B_{R}^{+}}|u_{j}|^{2}\,dx\leq C(\mu,\alpha,M)\int_{B_{R}^{+}}|x|^{2}\,dx\leq C(\mu,\alpha,M,N,R). (3.10)

Thus (3.8) and (3.10) imply that uju_{j} is a bounded sequence in H1(BR+)H^{1}(B_{R}^{+}). Hence, up to a subsequence, uju0u_{j}\rightharpoonup u_{0} weakly in H1(BR+)H^{1}(B_{R}^{+}). We rename the subsequence again as uju_{j}.

We have found a blow-up limit up to a subsequence u0u_{0} and have shown that u0u_{0} satisfies (G1), (G2) and (G3) in BR+¯\overline{B_{R}^{+}}. In order to show that u0𝒫u_{0}\in\mathcal{P}_{\infty}, it only remains to verify that u0u_{0} satisfies (G4), i.e. u0u_{0} is a local minimizer of J0(;BR+)J_{0}(\cdot;B_{R}^{+}) for all R>0R>0 (c.f. (3.7)). For that, we first claim that

BR+(|u0|2+Λ(u0))𝑑xlim infjBR+(Aj(x)uj,uj+Λ(uj))𝑑x.\int_{B_{R}^{+}}\Big{(}|\nabla u_{0}|^{2}+\Lambda(u_{0})\Big{)}\,dx\leq\liminf_{j\to\infty}\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla u_{j},\nabla u_{j}\rangle+\Lambda(u_{j})\Big{)}\,dx. (3.11)

Indeed, let us look separately at the term Jj(uj)J_{j}(u_{j}) on the right hand side of the above equation

Jj(uj)=BR+(Aj(x)uj,uj+λ+χ{uj>0}+λχ{uj0})𝑑x.J_{j}(u_{j})=\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla u_{j},\nabla u_{j}\rangle+\lambda_{+}\chi_{\{u_{j}>0\}}+\lambda_{-}\chi_{\{u_{j}\leq 0\}}\Big{)}\,dx.

We rewrite the first term as follows

BR+Aj(x)uj,uj𝑑x=BR+(Aj(x)Id)uj,uj𝑑x+BR+|uj|2𝑑x.\int_{B_{R}^{+}}\langle A_{j}(x)\nabla u_{j},\nabla u_{j}\rangle\,dx=\int_{B_{R}^{+}}\langle(A_{j}(x)-Id)\nabla u_{j},\nabla u_{j}\rangle\,dx+\int_{B_{R}^{+}}|\nabla u_{j}|^{2}\,dx. (3.12)

From ((P1)) and ((P2)) we have for all xBR+x\in B_{R}^{+}

Aj(x)IdL(BR+)Mrjα|x|αC(M,R,α)rjα0 as j.\|A_{j}(x)-Id\|_{L^{\infty}(B_{R}^{+})}\leq Mr_{j}^{\alpha}|x|^{\alpha}\leq C(M,R,\alpha)r_{j}^{\alpha}\to 0\mbox{ as $j\to\infty$}.

Therefore AjIdA_{j}\to Id uniformly and ujL2(B2+)\|\nabla u_{j}\|_{L^{2}(B_{2}^{+})} is bounded (c.f. (3.8)). Hence, the first term on the right hand side of (3.12) tends to zero as jj\to\infty. Thus, from (3.12) and by weak lower semi-continuity of H1H^{1} norm, we have

BR+|u0|2𝑑xlim infjBR+|uj|2𝑑x=lim infjBR+Aj(x)uj,uj𝑑x.\int_{B_{R}^{+}}|\nabla u_{0}|^{2}\,dx\leq\liminf_{j\to\infty}\int_{B_{R}^{+}}|\nabla u_{j}|^{2}\,dx=\liminf_{j\to\infty}\int_{B_{R}^{+}}\langle A_{j}(x)\nabla u_{j},\nabla u_{j}\rangle\,dx. (3.13)

For the second term, we claim that

BR+λ+χ{u0>0}+λχ{u00}dxlim infjBR+λ+χ{uj>0}+λχ{uj0}dx.\begin{split}\int_{B_{R}^{+}}\lambda_{+}\chi_{\{u_{0}>0\}}+\lambda_{-}\chi_{\{u_{0}\leq 0\}}\,dx&\leq\liminf_{j\to\infty}\int_{B_{R}^{+}}\lambda_{+}\chi_{\{u_{j}>0\}}+\lambda_{-}\chi_{\{u_{j}\leq 0\}}\,dx.\\ \end{split} (3.14)

To see this, we first show that for almost every xBR+x\in B_{R}^{+}, we have

λ+χ{u0>0}(x)+λχ{u00}(x)lim infj(λ+χ{uj>0}(x)+λχ{uj0}(x)).\lambda_{+}\chi_{\{u_{0}>0\}}(x)+\lambda_{-}\chi_{\{u_{0}\leq 0\}}(x)\leq\liminf_{j\to\infty}\left(\lambda_{+}\chi_{\{u_{j}>0\}}(x)+\lambda_{-}\chi_{\{u_{j}\leq 0\}}(x)\right). (3.15)

Indeed, let x0BR+({u0>0}{u0<0})x_{0}\in B_{R}^{+}\cap\big{(}\{u_{0}>0\}\cup\{u_{0}<0\}\big{)}. Then by the uniform convergence of uju_{j} to u0u_{0}, we can easily see that uj(x0)u_{j}(x_{0}) attains the sign of u0(x0)u_{0}(x_{0}) for sufficiently large value of jj. Hence, (3.15) holds in {u0>0}{u0<0}\{u_{0}>0\}\cup\{u_{0}<0\}.

Now, assume x0BR+{u0=0}x_{0}\in B_{R}^{+}\cap\{u_{0}=0\}. Then left hand side of (3.15) is equal to

λ+χ{u0>0}(x0)+λχ{u00}(x0)=λ.\lambda_{+}\chi_{\{u_{0}>0\}}(x_{0})+\lambda_{-}\chi_{\{u_{0}\leq 0\}}(x_{0})=\lambda_{-}.

Regarding RHS of (3.15), we see that

λ+χ{uj>0}(x0)+λχ{uj0}(x0)={λ+,if uj(x0)>0λ,if uj(x0)0.\lambda_{+}\chi_{\{u_{j}>0\}}(x_{0})+\lambda_{-}\chi_{\{u_{j}\leq 0\}}(x_{0})=\begin{cases}\lambda_{+},\qquad\mbox{if $u_{j}(x_{0})>0$}\\ \lambda_{-},\qquad\mbox{if $u_{j}(x_{0})\leq 0$}.\end{cases}

Since λ<λ+\lambda_{-}<\lambda_{+} (c.f. ((P3))), the right hand side in the equation above is always greater than or equal to λ\lambda_{-}. Then

λ+χ{u0>0}(x0)+λχ{u00}(x0)=λlim infj(λ+χ{uj>0}(x0)+λχ{uj0}(x0)).\begin{split}\lambda_{+}\chi_{\{u_{0}>0\}}(x_{0})+\lambda_{-}\chi_{\{u_{0}\leq 0\}}(x_{0})=\lambda_{-}\leq\liminf_{j\to\infty}\left(\lambda_{+}\chi_{\{u_{j}>0\}}(x_{0})+\lambda_{-}\chi_{\{u_{j}\leq 0\}}(x_{0})\right).\end{split}

Thus, (3.15) is proven for all xBR+x\in B_{R}^{+} and hence (3.14) holds by Fotou’s lemma.

By adding (3.13) and (3.14) and [6, Theorem 3.127] we obtain (3.11). Now we will use (3.11) prove the minimality of u0u_{0} for the functional J0(;BR+)J_{0}(\cdot;B_{R}^{+}) (c.f. 3.7).

Pick any wH1(BR+)w\in H^{1}(B_{R}^{+}) such that, u0wH01(BR+)u_{0}-w\in H_{0}^{1}(B_{R}^{+}). We construct an admissible competitor wjδw_{j}^{\delta} to compare the minimality of uju_{j} for the functional Jj(;BR+)J_{j}(\cdot;B_{R}^{+}). Then we intend to use (3.11).

In this direction, we define two cutoff functions ηδ:N\eta_{\delta}:\mathbb{R}^{N}\to\mathbb{R} and θ:\theta:\mathbb{R}\to\mathbb{R} as follows,

ηδ(x):={1,xBRδ0,xNBR,θ(t):={1,|t|1/20,|t|1.\eta_{\delta}(x):=\begin{cases}1,\;x\in B_{R-\delta}\\ 0,\;x\in\mathbb{R}^{N}\setminus B_{R}\end{cases},\theta(t):=\begin{cases}1,\;|t|\leq 1/2\\ 0,\;|t|\geq 1.\end{cases}

we can take |ηδ|C(N)δ|\nabla\eta_{\delta}|\leq\frac{C(N)}{\delta}. We define θj(x)=θ(xNdj)\theta_{j}(x)=\theta(\frac{x_{N}}{d_{j}}), for a sequence dj0d_{j}\to 0, which we be suitably chosen in later steps of the proof. Let wjδw_{j}^{\delta} be a test function defined as

wjδ:=w+(1ηδ)(uju0)+ηδθjϕj.w_{j}^{\delta}:=w+(1-\eta_{\delta})(u_{j}-u_{0})+\eta_{\delta}\theta_{j}\phi_{j}. (3.16)

Since, the function wjδw=(1ηδ)(uju0)+ηδθjϕjw_{j}^{\delta}-w=(1-\eta_{\delta})(u_{j}-u_{0})+\eta_{\delta}\theta_{j}\phi_{j} is continuous in BR+¯\overline{B_{R}^{+}} and is pointwise equal to zero on BR+\partial B_{R}^{+}, which is a Lipschitz surface in N\mathbb{R}^{N}. Therefore, ujwjδH01(BR+)u_{j}-w_{j}^{\delta}\in H_{0}^{1}(B_{R}^{+}). For further steps, the reader can refer to the Figure 1.

Refer to caption
Figure 1. (curvy line represents the free boundary of ww)

Let Ωδ,j=BR+{θj=0}{ηδ=1}\Omega_{\delta,j}=B_{R}^{+}\cap\{\theta_{j}=0\}\cap\{\eta_{\delta}=1\}, and δ,j=BR+Ωδ,j\mathcal{R}_{\delta,j}=B_{R}^{+}\setminus\Omega_{\delta,j} by observing wjδ=ww_{j}^{\delta}=w on Ωδ,j\Omega_{\delta,j} we see that

|{wjδ>0}BR+|=|{wjδ>0}Ωδ,j|+|{wjδ>0}δ,j|=|{w>0}Ωδ,j|+|{wjδ>0}δ,j|=|{w>0}(BR+δ,j)|+|{wjδ>0}δ,j|=|{w>0}BR+||{w>0}δ,j|+|{wjδ>0}δ,j|.\begin{split}|\{w_{j}^{\delta}>0\}\cap B_{R}^{+}|&=|\{w_{j}^{\delta}>0\}\cap\Omega_{\delta,j}|+|\{w_{j}^{\delta}>0\}\cap\mathcal{R}_{\delta,j}|\\ &=|\{w>0\}\cap\Omega_{\delta,j}|+|\{w_{j}^{\delta}>0\}\cap\mathcal{R}_{\delta,j}|\\ &=|\{w>0\}\cap(B_{R}^{+}\setminus\mathcal{R}_{\delta,j})|+|\{w_{j}^{\delta}>0\}\cap\mathcal{R}_{\delta,j}|\\ &=|\{w>0\}\cap B_{R}^{+}|-|\{w>0\}\cap\mathcal{R}_{\delta,j}|+|\{w_{j}^{\delta}>0\}\cap\mathcal{R}_{\delta,j}|.\end{split}

From the above discussions, we have

|{w>0}BR+||δ,j||{wjδ>0}BR+||{w>0}BR+|+|δ,j|.\begin{split}|\{w>0\}\cap B_{R}^{+}|-|\mathcal{R}_{\delta,j}|\leq|\{w_{j}^{\delta}>0\}\cap B_{R}^{+}|\leq|\{w>0\}\cap B_{R}^{+}|+|\mathcal{R}_{\delta,j}|.\end{split}

Since we know that limδ0(limj0|δ,j|)=0\lim_{\delta\to 0}\left(\lim_{j\to 0}|\mathcal{R}_{\delta,j}|\right)=0, we have

limδ0(limj|{wjδ>0}BR+|)=|{w>0}BR+|\lim_{\delta\to 0}\left(\lim_{j\to\infty}|\{w_{j}^{\delta}>0\}\cap B_{R}^{+}|\right)=|\{w>0\}\cap B_{R}^{+}| (3.17)

and similarly

limδ0(limj|{wjδ0}BR+|)=|{w0}BR+|.\lim_{\delta\to 0}\left(\lim_{j\to\infty}|\{w_{j}^{\delta}\leq 0\}\cap B_{R}^{+}|\right)=|\{w\leq 0\}\cap B_{R}^{+}|. (3.18)

Given uj𝒫rju_{j}\in\mathcal{P}_{r_{j}} and wjδujH01(BR+)w_{j}^{\delta}-u_{j}\in H_{0}^{1}(B_{R}^{+}), from the minimility of uju_{j} for the functional JjJ_{j} we have

BR+(Aj(x)uj,uj+λ+χ{uj>0}+λχ{uj0})𝑑xBR+(Aj(x)wjδ,wjδ+λ+χ{wjδ>0}+λχ{wjδ0})𝑑x\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla u_{j},\nabla u_{j}\rangle+\lambda_{+}\chi_{\{u_{j}>0\}}+\lambda_{-}\chi_{\{u_{j}\leq 0\}}\Big{)}\,dx\leq\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla w_{j}^{\delta},\nabla w_{j}^{\delta}\rangle+\lambda_{+}\chi_{\{w_{j}^{\delta}>0\}}+\lambda_{-}\chi_{\{w_{j}^{\delta}\leq 0\}}\Big{)}\,dx

and from (3.11) we obtain

BR+(|u0|2+λ+χ{u0>0}+λχ{u00})𝑑xlim infjBR+(Aj(x)wjδ,wjδ+λ+χ{wjδ>0}+λχ{wjδ0})𝑑xlim supjBR+(Aj(x)wjδ,wjδ+λ+χ{wjδ>0}+λχ{wjδ0})𝑑x\begin{split}\int_{B_{R}^{+}}\Big{(}|\nabla u_{0}|^{2}+\lambda_{+}\chi_{\{u_{0}>0\}}+\lambda_{-}\chi_{\{u_{0}\leq 0\}}\Big{)}\,dx&\leq\liminf_{j\to\infty}\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla w_{j}^{\delta},\nabla w_{j}^{\delta}\rangle+\lambda_{+}\chi_{\{w_{j}^{\delta}>0\}}+\lambda_{-}\chi_{\{w_{j}^{\delta}\leq 0\}}\Big{)}\,dx\\ &\leq\limsup_{j\to\infty}\int_{B_{R}^{+}}\Big{(}\langle A_{j}(x)\nabla w_{j}^{\delta},\nabla w_{j}^{\delta}\rangle+\lambda_{+}\chi_{\{w_{j}^{\delta}>0\}}+\lambda_{-}\chi_{\{w_{j}^{\delta}\leq 0\}}\Big{)}\,dx\end{split} (3.19)

from the same reasoning as for the justification of (3.13), we have

lim supjBR+Aj(x)wjδ,wjδ𝑑x=lim supjBR+|wjδ|2𝑑x\limsup_{j\to\infty}\int_{B_{R}^{+}}\langle A_{j}(x)\nabla w_{j}^{\delta},\nabla w_{j}^{\delta}\rangle\,dx=\limsup_{j\to\infty}\int_{B_{R}^{+}}|\nabla w_{j}^{\delta}|^{2}\,dx (3.20)

therefore rewriting (3.19)

BR+(|u0|2+λ+χ{u0>0}+λχ{u00})𝑑xlim supjBR+(|wjδ|2dx+λ+χ{wjδ>0}+λχ{wjδ0})𝑑x.\int_{B_{R}^{+}}\Big{(}|\nabla u_{0}|^{2}+\lambda_{+}\chi_{\{u_{0}>0\}}+\lambda_{-}\chi_{\{u_{0}\leq 0\}}\Big{)}\,dx\leq\limsup_{j\to\infty}\int_{B_{R}^{+}}\Big{(}|\nabla w_{j}^{\delta}|^{2}\,dx+\lambda_{+}\chi_{\{w_{j}^{\delta}>0\}}+\lambda_{-}\chi_{\{w_{j}^{\delta}\leq 0\}}\Big{)}\,dx. (3.21)

We claim that

limδ0(lim supjBR+|wjδ|2𝑑x)=BR+|w|2𝑑x\lim_{\delta\to 0}\left(\limsup_{j\to\infty}\int_{B_{R}^{+}}|\nabla w_{j}^{\delta}|^{2}\,dx\right)=\int_{B_{R}^{+}}|\nabla w|^{2}\,dx (3.22)

To obtain the claim above, we prove that

limδ0(lim supjBR+|(wjδw)|2𝑑x)=0.\lim_{\delta\to 0}\left(\limsup_{j\to\infty}\int_{B_{R}^{+}}|\nabla(w_{j}^{\delta}-w)|^{2}\,dx\right)=0.

From the definition of wjδw_{j}^{\delta}, we know that

wjδw=(1ηδ)(uju0)+ηδθjϕjw^{\delta}_{j}-w=(1-\eta_{\delta})(u_{j}-u_{0})+\eta_{\delta}\theta_{j}\phi_{j}

therefore we have

BR+|(wjδw)|2𝑑xC(BR+|((1ηδ)(uju0))|2𝑑x+BR+|(θjηδϕj)|2𝑑x)C(N)(BR+(1ηδ)2|(uju0)|2dx+1δ2BR+|uju0|2dx+BR+|(θjηδϕj)|2dx).\begin{split}\int_{B_{R}^{+}}|\nabla(w_{j}^{\delta}-w)|^{2}\,dx&\leq C\Bigg{(}\int_{B_{R}^{+}}|\nabla\big{(}(1-\eta_{\delta})(u_{j}-u_{0})\big{)}|^{2}\,dx+\int_{B_{R}^{+}}|\nabla(\theta_{j}\eta_{\delta}\phi_{j})|^{2}\,dx\Bigg{)}\\ &\leq C(N)\Bigg{(}\int_{B_{R}^{+}}(1-\eta_{\delta})^{2}|\nabla(u_{j}-u_{0})|^{2}\,dx+\frac{1}{\delta^{2}}\int_{B_{R}^{+}}|u_{j}-u_{0}|^{2}\,dx\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad+\int_{B_{R}^{+}}|\nabla(\theta_{j}\eta_{\delta}\phi_{j})|^{2}\,dx\Bigg{)}.\end{split} (3.23)

Let us consider the first term on the right hand side. We know that BR+|(uju0)|2𝑑x\int_{B_{R}^{+}}|\nabla(u_{j}-u_{0})|^{2}\,dx is uniformly bounded in jj\in\mathbb{N} (c.f. (3.8)). Therefore

limδ0(lim supjBR+(1ηδ)2|(uju0)|2𝑑x)=(limδ01ηδL(BR+)2)(lim supjBR+|(uju0)|2𝑑x)C(N,λ,μ,R,M,α)limδ01ηδL(BR+)=0\begin{split}\lim_{\delta\to 0}\Big{(}\limsup_{j\to\infty}\int_{B_{R}^{+}}(1-\eta_{\delta})^{2}|\nabla(u_{j}-u_{0})|^{2}\,dx\Big{)}&=\Big{(}\lim_{\delta\to 0}\|1-\eta_{\delta}\|_{L^{\infty}(B_{R}^{+})}^{2}\Big{)}\Big{(}\limsup_{j\to\infty}\int_{B_{R}^{+}}|\nabla(u_{j}-u_{0})|^{2}\,dx\Big{)}\\ &\leq C(N,\lambda,\mu,R,M,\alpha)\lim_{\delta\to 0}\|1-\eta_{\delta}\|_{L^{\infty}(B_{R}^{+})}=0\end{split} (3.24)

Regarding the second term, since |uju0||u_{j}-u_{0}| tends to zero in L2(BR+)L^{2}(B_{R}^{+}) as jj\to\infty, therefore the second term also tends to zero as jj\to\infty. We write

limδ0(limj1δ2BR+|uju0|2𝑑x)=0.\lim_{\delta\to 0}\Big{(}\lim_{j\to\infty}\frac{1}{\delta^{2}}\int_{B_{R}^{+}}|u_{j}-u_{0}|^{2}\,dx\Big{)}=0. (3.25)

Lastly, we claim that

limjBR+|(θjηδϕj)|2𝑑x=0.\lim_{j\to\infty}\int_{B_{R}^{+}}|\nabla(\theta_{j}\eta_{\delta}\phi_{j})|^{2}\,dx=0. (3.26)

Indeed, we have

BR+|(θjηδϕj)|2𝑑xC(BR+|ηδ|2(θjϕj)2𝑑x+BR+|ϕj|2(ηδθj)2𝑑x+BR+|θj|2(ηδϕj)2𝑑x).\int_{B_{R}^{+}}|\nabla(\theta_{j}\eta_{\delta}\phi_{j})|^{2}\,dx\leq C\left(\int_{B_{R}^{+}}|\nabla\eta_{\delta}|^{2}(\theta_{j}\phi_{j})^{2}\,dx+\int_{B_{R}^{+}}|\nabla\phi_{j}|^{2}(\eta_{\delta}\theta_{j})^{2}\,dx+\int_{B_{R}^{+}}|\nabla\theta_{j}|^{2}(\eta_{\delta}\phi_{j})^{2}\,dx\right). (3.27)

Since ηδ,θj1\eta_{\delta},\theta_{j}\leq 1, |ηδ|C(N)δ|\nabla\eta_{\delta}|\leq\frac{C(N)}{\delta} and ϕjL(BR+)M\|\nabla\phi_{j}\|_{L^{\infty}(B_{R}^{+})}\leq M (c.f. ((P1))), we obtain

BR+|ηδ|2(θjϕj)2𝑑x+BR+|ϕj|2(ηδθj)2𝑑xC(N)|{θj0}BR+|(1δ2+M2)\begin{split}\int_{B_{R}^{+}}|\nabla\eta_{\delta}|^{2}(\theta_{j}\phi_{j})^{2}\,dx+\int_{B_{R}^{+}}|\nabla\phi_{j}|^{2}(\eta_{\delta}\theta_{j})^{2}\,dx&\leq C(N)|\{\theta_{j}\neq 0\}\cap B_{R}^{+}|\Big{(}\frac{1}{\delta^{2}}+M^{2}\Big{)}\\ \end{split} (3.28)

We know that |{θj0}BR+|0|\{\theta_{j}\neq 0\}\cap B_{R}^{+}|\to 0 as jj\to\infty, hence from (3.28), the first and second term in (3.27) tend to zero as jj\to\infty. The last term in (3.23) also tends to zero as jj\to\infty, indeed from ((P1)) we have [ϕj]Cα(BR+)rjαM[\nabla\phi_{j}]_{C^{\alpha}(B_{R}^{+})}\leq r_{j}^{\alpha}M. Since ϕj(0)=0\phi_{j}(0)=0, therefore we have |ϕj|Rα[ϕj]BR+RαrjαM|\phi_{j}|\leq R^{\alpha}[\nabla\phi_{j}]_{B_{R}^{+}}\leq R^{\alpha}r_{j}^{\alpha}M in BR+B_{R}^{+}. Also, observing that |θj|1dj|\nabla\theta_{j}|\leq\frac{1}{d_{j}}, ηδ1\eta_{\delta}\leq 1 in BR+B_{R}^{+} we have

BR+|θj|2(ηδϕj)2𝑑xMR2αrj2αdj2|BR+|\int_{B_{R}^{+}}|\nabla\theta_{j}|^{2}(\eta_{\delta}\phi_{j})^{2}\,dx\leq MR^{2\alpha}\frac{r_{j}^{2\alpha}}{d_{j}^{2}}|B_{R}^{+}|

if we choose a sequence dj0+d_{j}\to 0^{+} such that we also have rjαdj0\frac{r_{j}^{\alpha}}{d_{j}}\to 0, the third term in (3.27) tends to zero as jj\to\infty. Plugging in the estimates above (3.24), (3.25), (3.26) in (3.23), we obtain the claim (3.22).

From the equations (3.17), (3.18), and (3.22) we obtain that the right hand side of (3.21) is equal to J0(w;BR+)J_{0}(w;B_{R}^{+}), therefore u0u_{0} is a minimizer of J0(;BR+)J_{0}(\cdot;B_{R}^{+}). That is

J0(u0;BR+)J0(w;BR+)J_{0}(u_{0};B_{R}^{+})\leq J_{0}(w;B_{R}^{+})\;

for every wH1(BR+)w\in H^{1}(B_{R}^{+}) such that u0wH01(BR+)u_{0}-w\in H_{0}^{1}(B_{R}^{+}). Since the inequality above (which corresponds to (G4)) and other verified properties of u0u_{0} (i.e. (G1), (G2) and (G3) in BR+B_{R}^{+}) hold for every R>0R>0, therefore u0Hloc1(+N)u_{0}\in H_{loc}^{1}(\mathbb{R}^{N}_{+}) satisfies all the properties in the Definition 2.3. Hence u0𝒫u_{0}\in\mathcal{P}_{\infty}.

After proving that the (subsequential) limits of blow-up are global solutions, we proceed to show that the positivity sets (and hence the free boundaries) of blow-ups converge in certain sense to that of blow-up limit. For this we will need to establish that the minimizers u𝒫ru\in\mathcal{P}_{r} are non-degenerate near the free boundary. In the proof below, we adapt the ideas from [2].

Proposition 3.9 (Non-degeneracy near the free boundary).

Let u𝒫r0u\in\mathcal{P}_{r_{0}} for some r0>0r_{0}>0 and x0B2/r0+x_{0}\in B_{2/r_{0}}^{+}. Then, for every 0<κ<10<\kappa<1 there exists a constant c(μ,N,κ,λ±)>0c(\mu,N,\kappa,\lambda_{\pm})>0 such that for all Br(x0)B2r0+B_{r}(x_{0})\subset B_{\frac{2}{r_{0}}}^{+}, we have

1rBr(x0)u+𝑑N1(x)<c(μ,N,κ,λ±)u+=0 in Bκr(x0).\frac{1}{r}\fint_{\partial B_{r}(x_{0})}u^{+}\,d\mathcal{H}^{N-1}(x)\,<\,c(\mu,N,\kappa,\lambda_{\pm})\implies\mbox{$u^{+}=0$ in $B_{\kappa r}(x_{0})$}. (3.29)
Proof.

We fix x0{u>0}B2+x_{0}\in\{u>0\}\cap B_{2}^{+} and r>0r>0 such that Br(x0)B2+B_{r}(x_{0})\subset B_{2}^{+}. We denote
γ:=1rBr(x0)u+𝑑x\gamma:=\frac{1}{r}\fint_{B_{r}(x_{0})}u^{+}\,dx. We know from Lemma 3.2 that the set {u>0}\{u>0\} is open. Also, since u𝒫r0u\in\mathcal{P}_{r_{0}}, there exists ACα(B2/r0+)N×NA\in C^{\alpha}(B_{2/r_{0}}^{+})^{N\times N}, φC1,α(B2/r0+)\varphi\in C^{1,\alpha}(B_{2/r_{0}}^{+}), λ±\lambda_{\pm} satisfying ((P1))-((P6)). Therefore uu solves the PDE div(A(x)u)=0\operatorname{div}(A(x)\nabla u)=0 in {u>0}B2/r0+\{u>0\}\cap B_{2/r_{0}}^{+}. By elliptic regularity theory, uu is locally Cloc1,α({u>0}B2/r0+)C_{loc}^{1,\alpha}(\{u>0\}\cap B_{2/r_{0}}^{+}). Then, for almost every ϵ>0\epsilon>0, Br{u>ε}B_{r}\cap\partial\{u>{\varepsilon}\} is a C1,αC^{1,\alpha} surface. Pick one such small ε>0{\varepsilon}>0 and we consider the test function vεv_{{\varepsilon}} given by

{div(A(x)vε)=0in (Br(x0)Bκr(x0)){u>ε}vε=uin Br(x0){uε}vε=εin Bκr(x0){u>ε}vε=uon Br(x0).\begin{cases}\operatorname{div}(A(x)\nabla v_{{\varepsilon}})=0\qquad\mbox{in $(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}))\cap\{u>{\varepsilon}\}$}\\ v_{{\varepsilon}}=u\qquad\qquad\;\;\qquad\qquad\mbox{in $B_{r}(x_{0})\cap\{u\leq{\varepsilon}\}$}\\ v_{{\varepsilon}}={\varepsilon}\qquad\;\;\qquad\qquad\qquad\mbox{in $B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}$}\\ v_{{\varepsilon}}=u\qquad\;\;\qquad\qquad\qquad\mbox{on $\partial B_{r}(x_{0})$.}\end{cases}

The function vεv_{{\varepsilon}} defined above belongs to H1(Br(x0))H^{1}(B_{r}(x_{0})), thanks to [5, Theorem 3.44] (\big{(}[5, Theorem 3.44] is proven for C1C^{1} domains, but the proof can also be adapted for Lipschitz domains [8, Theorem 4.6])\big{)}.

Refer to caption
Figure 2. Graph of vεv_{{\varepsilon}}.

We intend to show that vεv_{{\varepsilon}} is bounded in H1(Br(x0))H^{1}(B_{r}(x_{0})). This ensures the existence of limit limε0vε\lim_{{\varepsilon}\to 0}v_{{\varepsilon}} exists in weak sense in H1(Br(x0))H^{1}(B_{r}(x_{0})) and strong sense in L2(Br(x0))L^{2}(B_{r}(x_{0})). Let GG be the Green function for L(v)=div(A(x)v)L(v)=\operatorname{div}(A(x)\nabla v) in the ring Br(x0)Bκr(x0)B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}). Then if there is a function ww such that

{div(A(x)w)=0in Br(x0)Bκr(x0)w=uon Br(x0){u>ε}w=εelsewhere on (Br(x0)Bκr(x0)).\begin{cases}\operatorname{div}(A(x)\nabla w)=0\;&\mbox{in $B_{r}(x_{0})\setminus B_{\kappa r}(x_{0})$}\\ w=u\;&\mbox{on $\partial B_{r}(x_{0})\cap\{u>{\varepsilon}\}$}\\ w={\varepsilon}\;&\mbox{elsewhere on $\partial(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0})).$}\end{cases} (3.30)

We can also write that wε=(uε)+w-{\varepsilon}=(u-{\varepsilon})^{+} on Br(x0)\partial B_{r}(x_{0}) and wε=0w-{\varepsilon}=0 on Bκr(x0)\partial B_{\kappa r}(x_{0}). Consider any sequence {xk}Br(x0)Bκr(x0)\{x_{k}\}\subset B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}) such that xkx¯Bκr(x0)x_{k}\to\bar{x}\in\partial B_{\kappa r}(x_{0}). By Green representation formulae for wεw-{\varepsilon} in (3.30), we have w(xk)w(x¯)w(x_{k})\to w(\bar{x}), indeed since

0=(wε)(x¯)=limk(wε)(xk)=limk(Br(x0)Bκr(x0))(uε)+(y)(A(y)yG(xk,y))νy𝑑σ(y)=limkBr(x0){u>ε}(uε)+(y)(A(y)yG(xk,y))νy𝑑σ(y),xBκr(x0)\begin{split}0=(w-{\varepsilon})(\bar{x})&=\lim_{k\to\infty}(w-{\varepsilon})(x_{k})\\ &=\lim_{k\to\infty}\int_{\partial(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}))}\left(u-{\varepsilon})^{+}(y)(A(y)\nabla_{y}G(x_{k},y)\right)\cdot\nu_{y}\,d\sigma(y)\\ &=\lim_{k\to\infty}\int_{\partial B_{r}(x_{0})\cap\{u>{\varepsilon}\}}\left(u-{\varepsilon})^{+}(y)(A(y)\nabla_{y}G(x_{k},y)\right)\cdot\nu_{y}\,d\sigma(y),\qquad\forall x\in\partial B_{\kappa r}(x_{0})\end{split}

where νy\nu_{y} is the unit outer normal vector at a point yy on the boundary. We apply same arguments as above to w(x¯)\nabla w(\bar{x}) and from [11, Theorem 3.3 (vi)] on G(x,y)G(x,y) and therefore for x¯Bκr(x0)\bar{x}\in\partial B_{\kappa r}(x_{0})

|w(x¯)|C(μ)limkBr(x0){u>ε}|x(νyG(xk,y))(uε)+|dxC(μ)limkBr(x0)1|xky|N(uε)+𝑑N1(y)C(μ,N)(1κ)N1rBr(uε)+𝑑N1(y)C(μ,N,κ)γon Bκr(x0).\begin{split}\big{|}\nabla w(\bar{x})\big{|}&\leq C(\mu)\lim_{k\to\infty}\int_{\partial B_{r}(x_{0})\cap\{u>{\varepsilon}\}}\Big{|}\nabla_{x}\left(\frac{\partial}{\partial_{\nu_{y}}}G(x_{k},y)\right)(u-{\varepsilon})^{+}\Big{|}\,dx\\ &\leq C(\mu)\lim_{k\to\infty}\int_{\partial B_{r}(x_{0})}\frac{1}{|x_{k}-y|^{N}}(u-{\varepsilon})^{+}\,d\mathcal{H}^{N-1}(y)\\ &\leq\frac{C(\mu,N)}{(1-\kappa)^{N}}\frac{1}{r}\fint_{\partial B_{r}}(u-{\varepsilon})^{+}\,d\mathcal{H}^{N-1}(y)\leq C(\mu,N,\kappa)\gamma\;\mbox{on $\partial B_{\kappa r}(x_{0})$}.\end{split} (3.31)

We can easily check by respective definitions that wvεw\geq v_{{\varepsilon}} on (Br(x0)Bκr(x0))\partial(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0})), moreover, by maximum principle, since div(A(x)w)=0\operatorname{div}(A(x)\nabla w)=0 in Br(x0)Bκr(x0)B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}) and wεw\geq{\varepsilon} on Br(x0)Bκr(x0)\partial B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}), we have w>εw>{\varepsilon} in Br(x0)Bκr(x0)B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}). In particular wvεw\geq v_{{\varepsilon}} on Dε\partial D_{{\varepsilon}} where

Dε:=(Br(x0)Bκr(x0)){u>ε}.D_{{\varepsilon}}:=(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}))\cap\{u>{\varepsilon}\}.

By comparison principle, we know wvεw\geq v_{{\varepsilon}} in DεD_{{\varepsilon}} and since w=vε=εw=v_{{\varepsilon}}={\varepsilon} on Bκr(x0){u>ε}\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}, hence from (3.31)

|vε||w|C(μ,N,κ)γon Bκr(x0){u>ε}.|\nabla v_{{\varepsilon}}|\leq|\nabla w|\leq C(\mu,N,\kappa)\gamma\;\;\mbox{on $\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}$}. (3.32)

Given that div(A(x)vε)=0\operatorname{div}(A(x)\nabla v_{{\varepsilon}})=0 in DεD_{{\varepsilon}}, we have by divergence theorem and (3.32)

Dε(A(x)vε)(vεu)dx=Bκr(x0){u>ε}(uvε)(A(x)vε)ν(y)𝑑N1(y)C(μ)Bκr(x0){u>ε}|uε||vε|𝑑N1(y)C(μ,N,κ)γBκr(x0){u>ε}|uε|dN1(y)=:M0(u)\begin{split}\int_{D_{{\varepsilon}}}(A(x)\nabla v_{{\varepsilon}})\cdot\nabla(v_{\varepsilon}-u)\,dx&=\int_{\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}}(u-v_{{\varepsilon}})(A(x)\nabla v_{{\varepsilon}})\cdot\nu(y)\,d\mathcal{H}^{N-1}(y)\\ &\leq{C(\mu)}\int_{\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}}|u-{\varepsilon}||\nabla v_{{\varepsilon}}|\,d\mathcal{H}^{N-1}(y)\\ &\leq{C(\mu,N,\kappa)\gamma}\int_{\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}}|u-{\varepsilon}|\,d\mathcal{H}^{N-1}(y)=:M_{0}(u)\end{split}

justification of use of divergence theorem in DεD_{{\varepsilon}} can be found in [2, equation (3.4)]. From the calculations above, we can write

Dε(A(x)vε)(vεu)dxM0\displaystyle\int_{D_{{\varepsilon}}}(A(x)\nabla v_{{\varepsilon}})\cdot\nabla(v_{\varepsilon}-u)\,dx\leq M_{0}
\displaystyle\Rightarrow Dε(A(x)vε)vεdxM0+Dε(A(x)vε)udx\displaystyle\int_{D_{{\varepsilon}}}(A(x)\nabla v_{{\varepsilon}})\cdot\nabla v_{\varepsilon}\,dx\leq M_{0}+\int_{D_{{\varepsilon}}}(A(x)\nabla v_{{\varepsilon}})\cdot\nabla u\,dx
\displaystyle\Rightarrow μDε|vε|2𝑑xM0+1μDε|vε||u|𝑑x\displaystyle\mu\int_{D_{{\varepsilon}}}|\nabla v_{{\varepsilon}}|^{2}\,dx\leq M_{0}+\frac{1}{\mu}\int_{D_{{\varepsilon}}}|\nabla v_{{\varepsilon}}||\nabla u|\,dx
\displaystyle\Rightarrow μDε|vε|2𝑑xM0+ε02μDε|vε|2𝑑x+12ε0μDε|u|2𝑑x\displaystyle\mu\int_{D_{{\varepsilon}}}|\nabla v_{{\varepsilon}}|^{2}\,dx\leq M_{0}+\frac{{\varepsilon}_{0}}{2\mu}\int_{D_{{\varepsilon}}}|\nabla v_{{\varepsilon}}|^{2}\,dx+\frac{1}{2{\varepsilon}_{0}\mu}\int_{D_{\varepsilon}}|\nabla u|^{2}\,dx

putting very small ε0>0{\varepsilon}_{0}>0 in the last inequality, we have

Dε|vε|2dxM0+C(μ)Dε|u|2dx=:M1(u).\int_{D_{{\varepsilon}}}|\nabla v_{{\varepsilon}}|^{2}\,dx\leq M_{0}+C(\mu)\int_{D_{\varepsilon}}|\nabla u|^{2}\,dx=:M_{1}(u).

Since vε=ε in Bκr(x0){u>ε}v_{{\varepsilon}}={\varepsilon}\mbox{ in $B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}$} which implies vε=0\nabla v_{{\varepsilon}}=0 in Bκr(x0){u>ε}B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\} and vε=uv_{\varepsilon}=u in Br(x0)DεB_{r}(x_{0})\setminus D_{{\varepsilon}},

Br(x0)|vε|2dx=Dε|vε|2dx+Br(x0)Dε|u|2dx=:M2(u).\int_{B_{r}(x_{0})}|\nabla v_{\varepsilon}|^{2}\,dx=\int_{D_{{\varepsilon}}}|\nabla v_{{\varepsilon}}|^{2}\,dx+\int_{B_{r}(x_{0})\setminus D_{{\varepsilon}}}|\nabla u|^{2}\,dx=:M_{2}(u). (3.33)

By the definition of vεv_{{\varepsilon}}, 0<vεu0<v_{{\varepsilon}}\leq u on Dε\partial D_{{\varepsilon}} and div(A(x)vε)=div(A(x)u)=0\operatorname{div}(A(x)\nabla v_{{\varepsilon}})=\operatorname{div}(A(x)\nabla u)=0 in DεD_{{\varepsilon}}, therefore by comparison principle 0<vε<u0<v_{{\varepsilon}}<u in DεD_{{\varepsilon}}. In the set Bκr(x0){u>ε}B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\} we have vε=ε<uv_{{\varepsilon}}={\varepsilon}<u and vε=uv_{{\varepsilon}}=u in Br(x0){uε}B_{r}(x_{0})\cap\{u\leq{\varepsilon}\}. Overall we have 0<vεu0<v_{{\varepsilon}}\leq u in Br(x0){u>ε}B_{r}(x_{0})\cap\{u>{\varepsilon}\} Therefore

Br(x0)|vε|2𝑑xBr(x0){uε}|u|2𝑑x+Br(x0){u>ε}|u|2𝑑x=Br(x0)|u|2𝑑x.\int_{B_{r}(x_{0})}|v_{\varepsilon}|^{2}\,dx\leq\int_{B_{r}(x_{0})\cap\{u\leq{\varepsilon}\}}|u|^{2}\,dx+\int_{B_{r}(x_{0})\cap\{u>{\varepsilon}\}}|u|^{2}\,dx=\int_{B_{r}(x_{0})}|u|^{2}\,dx. (3.34)

Hence, from (3.33) and (3.34), vεv_{{\varepsilon}} is bounded in H1(Br(x0))H^{1}(B_{r}(x_{0})). Therefore, up to a subsequence, there exists a limit v=limε0vεv=\lim_{{\varepsilon}\to 0}v_{{\varepsilon}} in weak H1H^{1} sense, such that vv satisfies the following

{div(A(x)v)=0in (Br(x0)Bκr(x0)){u>0}v=uin Br(x0){u0}v=0in Bκr(x0){u>0}v=uon Br(x0).\begin{cases}\operatorname{div}(A(x)\nabla v)=0\qquad\mbox{in $(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}))\cap\{u>0\}$}\\ v=u\qquad\qquad\;\;\qquad\qquad\mbox{in $B_{r}(x_{0})\cap\{u\leq 0\}$}\\ v=0\qquad\;\;\qquad\qquad\qquad\mbox{in $B_{\kappa r}(x_{0})\cap\{u>0\}$}\\ v=u\qquad\;\;\qquad\qquad\qquad\mbox{on $\partial B_{r}(x_{0})$}.\end{cases} (3.35)

We verify the above properties (3.35) of vv at the end of this proof.

Let us use the function vv as a test function with respect to minimality condition on uu in Br(x0)B_{r}(x_{0}), we have

Br(x0)(A(x)u,u+λ(u))𝑑xBr(x0)(A(x)v,v+λ(v))𝑑x\begin{split}\int_{B_{r}(x_{0})}\Big{(}\langle A(x)\nabla u,\nabla u\rangle+\lambda(u)\Big{)}\,dx&\leq\int_{B_{r}(x_{0})}\Big{(}\langle A(x)\nabla v,\nabla v\rangle+\lambda(v)\Big{)}\,dx\\ \end{split}

since, v=uv=u in {u0}\{u\leq 0\} and {v>0}{u>0}\{v>0\}\subset\{u>0\}, the integration in the set {u0}\{u\leq 0\} gets cancelled from both sides and we are left with the terms mentioned below.
Set D0:=(Br(x0)Bκr(x0)){u>0}D_{0}:=(B_{r}(x_{0})\setminus B_{\kappa r}(x_{0}))\cap\{u>0\}, we have

Br(x0){u>0}(A(x)u,uA(x)v,v)𝑑xBr(x0){u>0}(Λ(v)Λ(u))𝑑x=Bκr(x0){u>0}(Λ(v)Λ(u))𝑑x=λ0|Bκr(x0){u>0}|.(λ0:=(λ+λ)).\begin{split}\int_{B_{r}(x_{0})\cap\{u>0\}}\Big{(}\langle A(x)\nabla u,\nabla u\rangle-\langle A(x)\nabla v,\nabla v\rangle\Big{)}\,dx&\leq\int_{B_{r}(x_{0})\cap\{u>0\}}(\Lambda(v)-\Lambda(u))\,dx\\ &=\int_{B_{\kappa r(x_{0})}\cap\{u>0\}}(\Lambda(v)-\Lambda(u))\,dx\\ &=\lambda_{0}|B_{\kappa r}(x_{0})\cap\{u>0\}|.\qquad(\lambda_{0}:=-(\lambda_{+}-\lambda_{-})).\\ \end{split}

We have second equality above because χ{v>0}=χ{u>0}\chi_{\{v>0\}}=\chi_{\{u>0\}} in D0D_{0}. Since v=0v=0 in D0D_{0}, we have

Bκr(x0){u>0}A(x)u,u𝑑x+D0(A(x)u,uA(x)v,v)𝑑xλ0|Bκr(x0){u>0}|.\begin{split}\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\langle A(x)\nabla u,\nabla u\rangle\,dx+\int_{D_{0}}\Big{(}\langle A(x)\nabla u,\nabla u\rangle-\langle A(x)\nabla v,\nabla v\rangle\Big{)}\,dx\leq\lambda_{0}|B_{\kappa r}(x_{0})\cap\{u>0\}|.\end{split}

Using the ellipticity of AA and shuffling the terms in the above equation, we obtain

Bκr(x0){u>0}(μ|u|2λ0)𝑑xD0(A(x)v,vA(x)u,u)𝑑x=D0(A(x)(vu),(v+u))𝑑x=D0(A(x)(vu),(uv+2v))𝑑x2D0A(x)v,(vu)𝑑xlim infε02D0A(x)vε,(vεu)=lim infε02DεA(x)vε,(vεu)𝑑x(since vε=u in DεD0)=lim infε02Bκr(x0){u>ε}(uε)(A(x)vε)ν𝑑xlim infε02μBκr(x0){u>ε}(uε)|νvε|𝑑x:=M.\begin{split}\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\Big{(}\mu|\nabla u|^{2}-\lambda_{0}\Big{)}\,dx&\leq\int_{D_{0}}\Big{(}\langle A(x)\nabla v,\nabla v\rangle-\langle A(x)\nabla u,\nabla u\rangle\Big{)}\,dx\\ &=\int_{D_{0}}\Big{(}\langle A(x)\nabla(v-u),\nabla(v+u)\rangle\Big{)}\,dx\\ &=\int_{D_{0}}\Big{(}\langle A(x)\nabla(v-u),\nabla(u-v+2v)\rangle\Big{)}\,dx\\ &\leq 2\int_{D_{0}}\langle A(x)\nabla v,\nabla(v-u)\rangle\,dx\\ &\leq\liminf_{{\varepsilon}\to 0}2\int_{D_{0}}\langle A(x)\nabla v_{{\varepsilon}},\nabla(v_{{\varepsilon}}-u)\rangle\,\\ &=\liminf_{{\varepsilon}\to 0}2\int_{D_{{\varepsilon}}}\langle A(x)\nabla v_{{\varepsilon}},\nabla(v_{{\varepsilon}}-u)\rangle\,dx\qquad\mbox{(since $v_{{\varepsilon}}=u$ in $D_{{\varepsilon}}\setminus D_{0}$)}\\ &=\liminf_{{\varepsilon}\to 0}2\int_{\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}}(u-{\varepsilon})(A(x)\nabla v_{{\varepsilon}})\cdot\nu\,dx\\ &\leq\liminf_{{\varepsilon}\to 0}\frac{2}{\mu}\int_{\partial B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\}}(u-{\varepsilon})\big{|}\nu\cdot\nabla v_{{\varepsilon}}\big{|}\,dx\,:=\,M.\end{split} (3.36)

The second to last equality in above calculation is obtained from integration by parts, its justification can be found in [2, equation (3.4)]. From (3.36) and (3.32), and using the trace inequality in H1(Bκr)H^{1}(B_{\kappa r}) we have (for some different constant C(κ)C(\kappa)),

MC(μ,N,κ)γBκr(x0)u+𝑑N1(x)C(μ,N,κ)γBκr(x0)(|u+|+1ru+)𝑑xC(μ,N,κ)γ[|Bκr(x0){u>0}|1/2(Bκr(x0)|u+|2𝑑x)1/2+1rsupBκr(x0)(u+)|{Bκr(x0){u>0}}|]C(μ,N,κ)γ[12λ0Bκr(x0){u>0}|u+|2dx+2λ0|Bκr(x0){u>0}|+1rsupBκr(x0)(u+)Bκr(x0){u>0}1dx]=C(μ,N,κ)γ2λ0(Bκr(x0){u>0}|u+|2λ0dx)+C(μ,N,κ)γλ0rsupBκr(x0)(u+)Bκr(x0){u>0}λ0𝑑x\begin{split}M&\leq C(\mu,N,\kappa)\gamma\int_{\partial B_{\kappa r(x_{0})}}u^{+}\,d\mathcal{H}^{N-1}(x)\\ &\leq C(\mu,N,\kappa)\gamma\int_{B_{\kappa r(x_{0})}}\Big{(}|\nabla u^{+}|+\frac{1}{r}u^{+}\Big{)}\,dx\\ &\leq C(\mu,N,\kappa)\gamma\Bigg{[}|B_{\kappa r(x_{0})}\cap\{u>0\}|^{1/2}\left(\int_{B_{\kappa r(x_{0})}}|\nabla u^{+}|^{2}\,dx\right)^{1/2}+\frac{1}{r}\sup_{B_{\kappa r}(x_{0})}(u^{+})\big{|}\{B_{\kappa r}(x_{0})\cap\{u>0\}\}\big{|}\Bigg{]}\\ &\leq C(\mu,N,\kappa)\gamma\Bigg{[}\frac{1}{2\sqrt{-\lambda_{0}}}\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}|\nabla u^{+}|^{2}\,dx+2\sqrt{-\lambda_{0}}|B_{\kappa r(x_{0})}\cap\{u>0\}|\\ &\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad+\frac{1}{r}\sup_{B_{\kappa r}(x_{0})}(u^{+})\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}1\,dx\Bigg{]}\\ &=\frac{C(\mu,N,\kappa)\gamma}{2\sqrt{-\lambda_{0}}}\left(\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}|\nabla u^{+}|^{2}-\lambda_{0}\,dx\right)+\frac{C(\mu,N,\kappa)\gamma}{\lambda_{0}r}\sup_{B_{\kappa r}(x_{0})}(u^{+})\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\lambda_{0}\,dx\end{split} (3.37)

we have used Hölder’s inequality and then Young’s inequality above. From Lemma 3.1, u+u^{+} is AA-subharmonic in Br(x0)B_{r}(x_{0}). If GG^{\prime} is the Green’s function for L(v)=div(A(x)v)L^{\prime}(v)=\operatorname{div}(A(x)\nabla v) in Br(x0)B_{r}(x_{0}), then by comparison principle and Green’s representation

u+(x)Br(x0)u+(y)(A(y)yG(x,y))νy𝑑N1(y)xBκr(x0).u^{+}(x)\leq\int_{\partial B_{r}(x_{0})}u^{+}(y)\left(A(y)\nabla_{y}G^{\prime}(x,y)\right)\cdot\nu_{y}\,d\mathcal{H}^{N-1}(y)\;\;\forall x\in B_{\kappa r}(x_{0}).

Since for all yBr(x0)y\in\partial B_{r}(x_{0}) and xBκr(x0)x\in B_{\kappa r}(x_{0}), we have 1|xy|N1C(κ)rN1\frac{1}{|x-y|^{N-1}}\leq\frac{C(\kappa)}{r^{N-1}}, then using the Green’s function estimates c.f. [11, Theorem 3.3 (v)] we get

supBκr(x0)u+C(μ)Br(x0)u+(y)|xy|N1𝑑N1(y)C(μ,κ,N)Br(x0)u+𝑑N1(y)=C(μ,κ,N)γr.\begin{split}\sup_{B_{\kappa r}(x_{0})}u^{+}&\leq C(\mu)\int_{\partial B_{r}(x_{0})}\frac{u^{+}(y)}{|x-y|^{N-1}}\,d\mathcal{H}^{N-1}(y)\\ &\leq C(\mu,\kappa,N)\fint_{\partial B_{r}(x_{0})}u^{+}\,d\mathcal{H}^{N-1}(y)=C(\mu,\kappa,N)\gamma r.\end{split} (3.38)

Use (3.36) and (3.38) in (3.37) and we have

μBκr(x0){u>0}(|u|2λ0)𝑑xC(μ,κ,N)γ2λ0Bκr(x0){u>0}(|u+|2λ0)𝑑x+C(μ,κ,N)γλ0rsupBκr(x0)(u+)Bκr(x0){u>0}λ0𝑑xC(μ,κ,N)γμλ0(1+C(κ)γλ0)μBκr(x0){u>0}(|u|2λ0)𝑑x.\begin{split}\mu\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\Big{(}|\nabla u|^{2}-\lambda_{0}\Big{)}\,dx&\leq\frac{C(\mu,\kappa,N)\gamma}{2\sqrt{-\lambda_{0}}}\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\Big{(}|\nabla u^{+}|^{2}-\lambda_{0}\Big{)}\,dx\\ &\qquad+\frac{C(\mu,\kappa,N)\gamma}{\lambda_{0}r}\sup_{B_{\kappa r}(x_{0})}(u^{+})\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\lambda_{0}\,dx\\ &\leq\frac{C(\mu,\kappa,N)\gamma}{\mu\sqrt{-\lambda_{0}}}\left(1+\frac{C(\kappa)\gamma}{\sqrt{-\lambda_{0}}}\right)\mu\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\Big{(}|\nabla u|^{2}-\lambda_{0}\Big{)}\,dx.\end{split}

If γ\gamma is small enough, then

Bκr(x0){u>0}(|u|2λ0)𝑑x=0\int_{B_{\kappa r}(x_{0})\cap\{u>0\}}\Big{(}|\nabla u|^{2}-\lambda_{0}\Big{)}\,dx=0

in particular |{u>0}Bκr(x0)|=0|\{u>0\}\cap B_{\kappa r(x_{0})}|=0, that is u+=0u^{+}=0 almost everywhere in Bκr(x0)B_{\kappa r}(x_{0}).

It remains to verify the properties of vv in (3.35). Before looking at the proof, we observe that for a given φCc(D0)\varphi\in C_{c}^{\infty}(D_{0}), then there exists ε0>0{\varepsilon}_{0}>0 such that φCc(Dε)\varphi\in C_{c}^{\infty}(D_{{\varepsilon}}) for all ε<ε0{\varepsilon}<{\varepsilon}_{0}. Indeed, since supp(φ)\operatorname{supp}(\varphi) is a compact set, and ε>0Dε\bigcup_{{\varepsilon}>0}D_{{\varepsilon}} is a cover of supp(φ)\operatorname{supp}(\varphi), then for a finite set {ε1,,εn}\{{\varepsilon}_{1},...,{\varepsilon}_{n}\} we have supp(φ)i=1nDεiDεmax\operatorname{supp}(\varphi)\subset\bigcup_{i=1}^{n}D_{{\varepsilon}_{i}}\subset D_{{\varepsilon}_{max}} where εmax=max(ε1,,εn){\varepsilon}_{max}=\max({\varepsilon}_{1},...,{\varepsilon}_{n}). Therefore, φCc(Dε)\varphi\in C_{c}^{\infty}(D_{{\varepsilon}}) for all ε<εmax{\varepsilon}<{\varepsilon}_{max}.

Let us first verify that div(A(x)v)=0\operatorname{div}(A(x)\nabla v)=0 in D0D_{0}. For this let φCc(D0)\varphi\in C_{c}^{\infty}(D_{0}), then from continuity of uu, there exists a ε0>0{\varepsilon}_{0}>0 such that supp(φ)Dε\operatorname{supp}(\varphi)\subset D_{\varepsilon} for all ε<ε0{\varepsilon}<{\varepsilon}_{0}, also we have

D0Av,φ𝑑x=supp(φ)Av,φ𝑑x\int_{D_{0}}\langle A\nabla v,\nabla\varphi\rangle\,dx=\int_{\operatorname{supp}(\varphi)}\langle A\nabla v,\nabla\varphi\rangle\,dx (3.39)

since supp(φ)Dε\operatorname{supp}(\varphi)\subset D_{\varepsilon}, from the definition of vεv_{{\varepsilon}} we have

supp(φ)Avε,φ𝑑x=0\int_{\operatorname{supp}(\varphi)}\langle A\nabla v_{{\varepsilon}},\nabla\varphi\rangle\,dx=0

and we know that vv is a weak limit of vεv_{{\varepsilon}} in H1(Br(x0))H^{1}({B_{r}(x_{0})}), therefore from (3.39) we have

D0Av,φ𝑑x=supp(φ)Av,φ𝑑x=limε0supp(φ)vε,φ𝑑x=0.\int_{D_{0}}\langle A\nabla v,\nabla\varphi\rangle\,dx=\int_{\operatorname{supp}(\varphi)}\langle A\nabla v,\nabla\varphi\rangle\,dx=\lim_{{\varepsilon}\to 0}\int_{\operatorname{supp}(\varphi)}\langle\nabla v_{{\varepsilon}},\nabla\varphi\rangle\,dx=0.

Hence we show that div(A(x)v)=0\operatorname{div}(A(x)\nabla v)=0 in D0D_{0}. To show that v=0v=0 in Bκr(x0){u>0}B_{\kappa r}(x_{0})\cap\{u>0\}, we now take the function φCc(Bκr(x0){u>0})\varphi\in C_{c}^{\infty}(B_{\kappa r}(x_{0})\cap\{u>0\}). From the same reasoning as above we know that there exists an ε0>0{\varepsilon}_{0}>0 such that supp(φ)Bκr(x0){u>ε}\operatorname{supp}(\varphi)\subset B_{\kappa r}(x_{0})\cap\{u>{\varepsilon}\} for all ε<ε0{\varepsilon}<{\varepsilon}_{0}. From the definition of vεv_{{\varepsilon}}, we have

supp(φ)vεφ𝑑x={u>ε}Bκr(x0)vεφ𝑑x=ε{u>ε}Bκr(x0)φ𝑑x\int_{\operatorname{supp}(\varphi)}v_{{\varepsilon}}\varphi\,dx=\int_{\{u>{\varepsilon}\}\cap B_{\kappa r}(x_{0})}v_{{\varepsilon}}\varphi\,dx={\varepsilon}\int_{\{u>{\varepsilon}\}\cap B_{\kappa r}(x_{0})}\varphi\,dx

and in limit ε0{\varepsilon}\to 0, from the above equation we have

supp(φ)vφ𝑑x=limε0supp(φ)vεφ𝑑x=0\int_{\operatorname{supp}(\varphi)}v\varphi\,dx=\lim_{{\varepsilon}\to 0}\int_{\operatorname{supp}(\varphi)}v_{{\varepsilon}}\varphi\,dx=0

and therefore v=0v=0 a.e. in Bκr(x0){u>0}B_{\kappa r}(x_{0})\cap\{u>0\}. To prove that v=uv=u in {u0}\{u\leq 0\}, we observe that {u0}{uε}\{u\leq 0\}\subset\{u\leq{\varepsilon}\}, hence from the definition of vεv_{{\varepsilon}}, have

vε=u in {u0}.v_{{\varepsilon}}=u\mbox{ in $\{u\leq 0\}$}.

since the weak limits maintain the equality (c.f. [16, Lemma 3.14]) the claim follows in the limit ε0{\varepsilon}\to 0. Apart from that, since vε=uv_{{\varepsilon}}=u on Br(x0)\partial B_{r}(x_{0}) therefore from conservation of traces in weak convergence, it follows that v=uv=u on Br(x0)\partial B_{r}(x_{0}). This completes the proof of Proposition 3.9. ∎

Remark 3.10.

In the proposition above the constant is local in nature, this means, the value of the constant depends on the choice of compact set KB2+K\subset\subset B_{2}^{+} where x0Kx_{0}\in K.

Lemma 3.11.

Let u0u_{0} and uku_{k} be as in Theorem 3.8. Then, for a subsequence of uku_{k}, for any R>0R>0 we have

χ{uk>0}BR+χ{u0>0}BR+ a.e. in BR+.\chi_{\{u_{k}>0\}\cap B_{R}^{+}}\to\chi_{\{u_{0}>0\}\cap B_{R}^{+}}\qquad\mbox{ a.e. in $B_{R}^{+}$}. (3.40)

This in turn implies

χ{uk>0}BR+χ{u0>0}BR+ in L1(BR+).\chi_{\{u_{k}>0\}\cap B_{R}^{+}}\to\chi_{\{u_{0}>0\}\cap B_{R}^{+}}\qquad\mbox{ in $L^{1}(B_{R}^{+})$}. (3.41)
Proof.

From Lemma 3.8, we can consider a subsequence of uku_{k} such that uku0u_{k}\to u_{0} in L(BR+)L^{\infty}(B_{R}^{+}). Let xBR+x\in B_{R}^{+}. If x{u0>0}BR+x\in\{u_{0}>0\}\cap B_{R}^{+} (or χ{u0>0}BR+(x)=1\chi_{\{u_{0}>0\}\cap B_{R}^{+}}(x)=1), then uk(x)>u(x)2>0u_{k}(x)>\frac{u(x)}{2}>0 (or χ{uk>0}BR+(x)=1\chi_{\{u_{k}>0\}\cap B_{R}^{+}}(x)=1) for kk sufficiently large. Thus we conclude that

 χ{uk>0}BR+(x)χ{u0>0}BR+(x) as k for all x{u0>0}BR+.\mbox{ $\chi_{\{u_{k}>0\}\cap B_{R}^{+}}(x)\to\chi_{\{u_{0}>0\}\cap B_{R}^{+}}(x)$ as $k\to\infty$ for all $x\in\{u_{0}>0\}\cap B_{R}^{+}$}.

If x{u00}oBR+x\in\{u_{0}\leq 0\}^{o}\cap B_{R}^{+} (or χ{u0>0}BR+(x)=0\chi_{\{u_{0}>0\}\cap B_{R}^{+}}(x)=0), then there exists δ>0\delta>0 such that Bδ(x){u00}BR+B_{\delta}(x)\subset\{u_{0}\leq 0\}\cap B_{R}^{+}. Thus we have 1δBδ(x)u0+𝑑N1=0\frac{1}{\delta}\fint_{\partial B_{\delta}(x)}u_{0}^{+}\,d\mathcal{H}^{N-1}=0. Again, by the uniform covergence of uku_{k} to u0u_{0} in BR+B_{R}^{+} (c.f. Lemma 3.8) we obtain

1δBδ(x)uk+𝑑N112c(μ,N,λ±)for k sufficiently large.\frac{1}{\delta}\fint_{\partial B_{\delta}(x)}u_{k}^{+}\,d\mathcal{H}^{N-1}\leq\frac{1}{2}c(\mu,N,\lambda_{\pm})\qquad\mbox{for $k$ sufficiently large.} (3.42)

Here c(μ,N,λ±)c(\mu,N,\lambda_{\pm}) is as in Proposition 3.9. This implies uk0u_{k}\leq 0 in Bδ2(x)B_{\frac{\delta}{2}}(x) (c.f. Proposition 3.9). In particular, χ{uk(x)0}(x)=0\chi_{\{u_{k}(x)\leq 0\}}(x)=0 for kk sufficiently large. This way, we obtain

 χ{uk>0}BR+(x)χ{u0>0}BR+(x) as k for all x{u00}BR+.\mbox{ $\chi_{\{u_{k}>0\}\cap B_{R}^{+}}(x)\to\chi_{\{u_{0}>0\}\cap B_{R}^{+}}(x)$ as $k\to\infty$ for all $x\in\{u_{0}\leq 0\}\cap B_{R}^{+}$}. (3.43)

From the representation theorem [2, Theorem 7.3], we know that |{u0>0}BR+|=0|\partial\{u_{0}>0\}\cap B_{R}^{+}|=0. From (3.42), (3.43) and the fact that |{u0>0}BR+|=0|\partial\{u_{0}>0\}\cap B_{R}^{+}|=0, we obtain the claim (3.40). Since |χ{uk>0}BR+|1|\chi_{\{u_{k}>0\}\cap B_{R}^{+}}|\leq 1, the claim (3.41) follows from Lebesgue’s dominated convergence theorem. ∎

4. The main result

We rephrase the notion of the tangential touch of the free boundary to the fixed boundary, which is equivalent to the tangential touch condition mentioned in statement of Theorem 2.4.

In the proof of our main result, we will show that given u𝒫1u\in\mathcal{P}_{1}, for every ε>0{\varepsilon}>0 there exists ρε>0\rho_{{\varepsilon}}>0 such that

{u>0}Bρ+Bρ+Kε, 0<ρρε\partial\{u>0\}\cap B_{\rho}^{+}\subset B_{\rho}^{+}\setminus K_{{\varepsilon}},\qquad\forall\;0<\rho\leq\rho_{{\varepsilon}} (4.1)

where

Kε:={x+N:xNεx12++xN12}.K_{{\varepsilon}}:=\big{\{}x\in\mathbb{R}^{N}_{+}\,:\,x_{N}\geq{\varepsilon}\sqrt{x_{1}^{2}+...+x_{N-1}^{2}}\big{\}}.
Proof of Theorem 2.4.

We assume, by contradiction that the free boundaries of functions in 𝒫1\mathcal{P}_{1} do not touch the origin in a tangential fashion to the plane. Then, there exists ε>0{\varepsilon}>0 and sequences vj𝒫1v_{j}\in\mathcal{P}_{1} and xjF(vj)Kεx_{j}\in F(v_{j})\cap K_{{\varepsilon}} such that |xj|0|x_{j}|\to 0 as jj\to\infty. Let rj=|xj|r_{j}=|x_{j}| and we consider the blowups uj:=(vj)rju_{j}:=(v_{j})_{r_{j}}.

Let u0:=limjuju_{0}:=\lim_{j\to\infty}u_{j} as in Lemma 3.8. Also, let x0B1+Kεx_{0}\in\partial B_{1}^{+}\cap K_{{\varepsilon}} be a limit up to a subsequence (still called xjx_{j}) such that x0=limjxj|xj|x_{0}=\lim_{j\to\infty}\frac{x_{j}}{|x_{j}|}. Since xjF(vj)x_{j}\in F(v_{j}), we have vj(xj)=0v_{j}(x_{j})=0. Therefore on rescaling, uj(xjrj)=1rjvj(xj)=0u_{j}(\frac{x_{j}}{r_{j}})=\frac{1}{r_{j}}v_{j}(x_{j})=0. In the limit as jj\to\infty we have

u0(x0)=limjuj(xjrj)=0.u_{0}(x_{0})=\lim_{j\to\infty}u_{j}\left(\frac{x_{j}}{r_{j}}\right)=0.

From the density assumption that uju_{j} satisfy condition (2.2) and Lemma 3.41 we have for any given R>0R>0

|{u0>0}BR+||BR+|=BR+χ{u0>0}𝑑x=limjBR+χ{uj>0}𝑑x=limj1|BRrj+|BRrj+χ{vj>0}𝑑x=limj|{vj>0}BRrj+||BRrj+|>𝒟.\begin{split}\frac{|\{u_{0}>0\}\cap B_{R}^{+}|}{|B_{R}^{+}|}=\fint_{B_{R}^{+}}\chi_{\{u_{0}>0\}}\,dx&=\lim_{j\to\infty}\fint_{B_{R}^{+}}\chi_{\{u_{j}>0\}}\,dx\\ &=\lim_{j\to\infty}\frac{1}{|B_{Rr_{j}}^{+}|}\int_{B_{Rr_{j}}^{+}}\chi_{\{v_{j}>0\}}\,dx\\ &=\lim_{j\to\infty}\frac{|\{v_{j}>0\}\cap B_{Rr_{j}}^{+}|}{|B_{Rr_{j}}^{+}|}>\mathcal{D}.\end{split} (4.2)

We can see that the computations done in (4.2), in fact shows that the density property remains invariant under blowup of any function vv,. This way, we conclude that the function (u0)0(u_{0})_{0} which is the blowup limit of (u0)(u_{0}) (in particular (u0)0:=limr0(u0)r(u_{0})_{0}:=\lim_{r\to 0}(u_{0})_{r}) also satisfies

|{(u0)0>0}BR+||BR+|>𝒟,R>0.\frac{|\{(u_{0})_{0}>0\}\cap B_{R}^{+}|}{|B_{R}^{+}|}>\mathcal{D},\qquad\forall R>0. (4.3)

Now, we note that from Lemma 3.8 u0𝒫u_{0}\in\mathcal{P}_{\infty}. Moreover, from (4.2) u00u_{0}\not\equiv 0 and from [14, Theorem 4.2, Lemma 4.3] we have u00u_{0}\geq 0 also, from (4.3), we conclude (u0)00(u_{0})_{0}\not\equiv 0. This way, again by [14, Theorem 4.9], we have u0(x)=cxN+u_{0}(x)=c\,x_{N}^{+} for all x+Nx\in\mathbb{R}^{N}_{+} for some constant c>0c>0.

Hence the function u0u_{0} cannot be equal to zero at any point in +N\mathbb{R}^{N}_{+}. But we have x0B1+Kεx_{0}\in\partial B_{1}^{+}\cap K_{{\varepsilon}} and u0(x0)=0u_{0}(x_{0})=0. This leads to a contradiction. ∎

Remark 4.1.

Interested readers may check that the modulus of continuity σ\sigma mentioned in the statement of Theorem 2.4 can be written as

σ(r)=sup(xnx12++xN12:x(Bρ+{0})F(u),ρr).\sigma(r)=\sup\left(\frac{x_{n}}{\sqrt{x_{1}^{2}+...+x_{N-1}^{2}}}\,:\,x\in(B_{\rho}^{+}\setminus\{0\})\cap F(u),\rho\leq r\right).

References

  • [1] H. Alt and L. Caffarelli, Existence and regularity for a minimum problem with free boundary., Journal für die reine und angewandte Mathematik, 325 (1981), pp. 105–144.
  • [2] H. Alt, L. Caffarelli, and A. Friedman, Variational problems with two phases and their free boundaries, Transactions of the American Mathematical Society, 282 (1984), pp. 431–431.
  • [3] H. W. Alt and G. Gilardi, The behavior of the free boundary for the dam problem, Annali della Scuola Normale Superiore di Pisa - Classe di Scienze, Ser. 4, 9 (1982), pp. 571–626.
  • [4] M. Borsuk, Dini continuity of the first order derivatives of solutions to the problem for linear second-order dirichlet elliptic equations in a nonsmooth domain, Siberian Mathematical Journal, 39 (1998), pp. 261–280.
  • [5] F. Demengel and G. Demengel, Functional Spaces for the Theory of Elliptic Partial Differential Equations, Springer, Berlin, 2012.
  • [6] C. Dunn, Introduction to Analysis, Textbooks in Mathematics, CRC Press, 2017.
  • [7] Z. Eduardo and G. B. H, Jets, Wakes, and Cavities, Elsevier Science, 1957.
  • [8] L. C. Evans and R. F. Gariepy, Measure Theory and Fine Properties of Functions, Revised Edition, Textbooks in Mathematics, CRC Press, 2015.
  • [9] E. Giusti, Direct methods in the calculus of variations, World Scientific, 2003.
  • [10] G. Gravina and L. Giovanni, On the behavior of the free boundary for a one-phase bernoulli problem with mixed boundary conditions, Communications on Pure and Applied Analysis, 19 (2019), pp. 4853–4878.
  • [11] M. Grüter and K.-O. Widman, The green function for uniformly elliptic equations, Manuscripta Matematica, 37 (1982), pp. 303–342.
  • [12] E. Indrei, Free boundary regularity near the fixed boundary for the fully nonlinear obstacle problem, Submitted article.
  • [13]  , Boundary regularity and non-transversal intersection for the fully nonlinear obstacle problem, Communications on Pure and Applied Mathematics, 72 (2019), pp. 1459–1473.
  • [14] A. Karakhanyan, C. Kenig, and H. Shahgholian, The behavior of the free boundary near the fixed boundary for a minimization problem, Calculus of Variations and Partial Differential Equations, 28 (2006), pp. 15–31.
  • [15] D. Moreira and J. E. M. Braga, Upto the boundary gradient estimates for viscosity solutions to nonlinear free boundary problems with unbounded measurable ingredients, Submitted Article.
  • [16] D. Moreira and H. Shrivastava, Optimal regularity for variational solutions of free transmission problems., Submitted paper, (2021).