1 Introduction

There are some literatures on the nonlinear Schrödinger equations involving the fractional Laplacian

$$\begin{aligned} (-\Delta )^{\alpha }u+A(x)u=f(x,u), \ \ x \in \mathbb {R}^n, \end{aligned}$$
(1.1)

where \(n\ge 2\) and \(\alpha \in (0,1)\). We recall that the fractional Laplacian is defined as

$$\begin{aligned} (-\Delta )^{\alpha }u(x)= C_{n,\alpha }\;\text {P.V.}\int _{\mathbb {R}^n} \frac{u(x)-u(y)}{|x-y|^{n+2\alpha }} dy, \end{aligned}$$
(1.2)

where \(\text {P.V.}\) stands for the Principle Value and \(C_{n,\alpha }>0\). For more backgrounds on \((-\Delta )^{\alpha }\), we refer the readers to [1][7][11][26], and etc.

For instance, Benhassine [4] took

$$\begin{aligned} \psi (x,t)=e^{i\omega t}u(x), \end{aligned}$$

where u is a real-valued function and f is assumed to satisfy \(f(x,e^{i\omega t}u)=e^{-i\omega t}f(x,u)\). He proved the existence of solutions to (1.1) is equivalent to the existence of standing wave solutions for fractional Schrödinger equations of the form:

$$\begin{aligned} i \frac{\partial \psi }{\partial t}=(-\Delta )^\alpha \psi +(A(x)+\omega )\psi -f(x,\psi ), \end{aligned}$$
(1.3)

where \(\omega \) is constant and \(\alpha \in (0,1)\). That is to say, \(\psi (x,t)\) solves (1.3) if and only if u(x) solves (1.1).

When \(\alpha =1\), we have a nonlinear Schrödinger equation with the fraction Laplacian:

$$\begin{aligned} (-\Delta )u+A(x)u=f(x,u),x\in \mathbb {R}^n. \end{aligned}$$

Here, we mention that some earlier work were done by Floer and Weinstein [16], Rabinowitz [28], Wang [32], del Pino and Felmer [27], Azzollini[3], Secchi[29], and etc. In 2012, Felmer-Quaas-Tan [15] considered the case when \(A(x)=1\) and studied the existence, regularity, decay and symmetry properties of positive solutions for the nonlinear Schrödinger equation involving the fractional Laplacian under the conditions \(u>0, \ \text{ in }\ \mathbb {R}^n\) and \(\lim _{|x|\rightarrow \infty }u(x)=0\). In 2016, Secchi [30] proved the existence of radially symmetric solutions for (1.1), where f(xu) is replaced by g(u). In 2018, Servadei studied the nonlinear fractional Schrödinger equation under the condition \(u\in H^{\alpha }(\mathbb {R}^n, \mathbb {R})\). There are also articles about Schrödinger equation under other conditions, such as [2, 28, 20, 24, 18, 19] and [22]. In this paper, we consider \(u \in C^{1,1}_{loc} \cap L_{\alpha }(\mathbb R^n) \), where

$$\begin{aligned} \frac{}{}L_{\alpha }(\mathbb R^n) = \{ u \in L^1_{loc} \mid \int _{\mathbb {R}^n} \frac{|u(x)|}{1+|x|^{n+\alpha }} d x < \infty \}. \end{aligned}$$

On the other hand, Cheng-Huang-Li [14] considered the zero-Dirichlet problem with more generalized nonlinear term,

$$\begin{aligned} {\left\{ \begin{array}{ll} (-\Delta )^{\frac{\alpha }{2}}u=f(x,u,\nabla u) \ \text{ in } \ \Omega \\ u>0 \ \text{ in }\ \Omega ; \ u\equiv 0 \ \text{ in }\ \mathbb R^n\backslash \Omega . \end{array}\right. } \end{aligned}$$
(1.4)

Note that in (1.4), \(\Omega \) has been considered as a convex domain (bounded or unbounded) in \(x_1\)-direction in \(\mathbb R^n\) (we say a domain \(\Omega \) is convex in \(x_1\)-direction if and only if \((x_1,x^{\prime } ), (x_2,x^{\prime })\in \Omega \) imply that \(((1-t)x_1+tx_2,x^{\prime })\in \Omega \) for any \(t\in (0,1)\)). Moreover, the nonlinear term f belongs to the function space \(\digamma \), where \(\digamma \) is defined as the collections of all functions \(f(x,u,p):\mathbb {R}^n\times \mathbb {R}\times \mathbb {R}^n\rightarrow \mathbb {R}\) such that for any \(M>0, u_1,u_2\in [-M,M]\) and any \(x, p\in \mathbb {R}^n\),

$$\begin{aligned} |f(x,u_1,\mathbf{p} )-f(x,u_2,\mathbf{p} )|\le C_M|u_1-u_2|\ \text{ for } \text{ some } \ C_M>0. \end{aligned}$$

Under some other adaptable conditions on u and f, Cheng-Huang-Li[14] showed that the solution u for (1.4) is monotone and symmetric when \(\Omega \) is bounded or unbounded respectively. We generalize the method of moving plane for the fractional Laplacian used by Li [23], Caffarelli and Silvestre [5], Cheng [13], Chen-Li-Li[10], Wang[31], Ma-Chen[25], Garofalo[17], Cao[6], Li [21], Chen [9], and etc.

In this paper our goal is to study the nonlinear Schrödinger equation involving the fractional Laplacian:

$$\begin{aligned} {\left\{ \begin{array}{ll} (-\Delta )^{\frac{\alpha }{2}}u+A(x)u=f(x,u,\nabla u) \ \ \text{ in } \ \Omega \\ u>0 \ \ \text{ in }\ \Omega \\ u\equiv 0 \ \ \text{ in }\ \Omega ^c, \end{array}\right. } \end{aligned}$$
(1.5)

where \(\alpha \in (0,2)\). We shall show that as long as A(x) and \(f(x,u,\nabla u)\) in (1.5) satisfy certain conditions (these conditions will be exhibited in detail in the next theorem), the positive solutions in \(\Omega \) will have symmetry and monotonicity. While in \({\mathbb R}^n_+\) similar results could not exist. Our main results are stated in the following three theorems.

Theorem 1.1

Let \(\Omega \) be a bounded domain in \(\mathbb {R}^n\) which is convex in \(x_1\)-direction and symmetric about \(x_1=0\). We suppose that \(u\in C(\mathbb {R}^n)\bigcap \mathcal C_{loc}^{1,1}(\Omega )\) solves (1.5) for \(\alpha \in (0,2)\), \(f(x,v,\mathbf{p} )\in \digamma \) satisfies

$$\begin{aligned} f(x_1,x',v,p_1,p_2,\cdots ,p_n)\le f(\overline{x}_1,x',v,-p_1,p_2,\cdots ,p_n), \ \forall -x_1, p_1\ge 0, \end{aligned}$$

and A(x) is bounded below and symmetric in \(x_1\)-direction and satisfies

$$\begin{aligned} A(x_1,x')<A(\overline{x}_1,x'), \; \forall x_1>\overline{x}_1,\ (x_1,x'), (\overline{x}_1,x')\in \Omega . \end{aligned}$$

Then \(u(x_1,x')\) is strictly increasing in the left half of \(\Omega \) in \(x_1\)-direction and

$$\begin{aligned} u(x_1,x')\le u(-x_1,x'), \; \forall x_1<0,\ (x_1,x')\in \Omega . \end{aligned}$$

Moreover, if \(f(x_1,x',v,p_1,p_2,\cdot \cdot \cdot ,p_n)= f(-x_1,x',v,-p_1,p_2,\cdot \cdot \cdot ,p_n)\), then

$$\begin{aligned} u(x_1,x')= u(-x_1,x'). \end{aligned}$$

To state our second theorem in unbounded domain \(\Omega \), we need to impose a growth condition on \(f(x,v,\mathbf{p} )\):

$$\begin{aligned} \frac{|f(x,u_1,\mathbf{p} )-f(x,u_2,\mathbf{p} )|}{|u_1-u_2|}\le C(|u_1\mid ^s+|u_2|^s)\ as\ u_1,u_2\rightarrow 0,\ for\ some\ s>0. \end{aligned}$$
(1.6)

Theorem 1.2

Let \(\Omega \) be an unbounded domain in \(\mathbb {R}^n\), which is convex in \(x_1\)-direction and symmetric about \(x_1=0\). Suppose that \(u\in C(\mathbb {R}^n)\bigcap \mathcal C_{loc}^{1,1}(\Omega )\cap L_\alpha \) solves (1.5) for \(\alpha \in (0,2)\) and \(f(x,v,\mathbf{p} )\in \digamma \). If \(f(x,v,\mathbf{p} )\) satisfies (??) (1.6) and u(x) satisfies

$$\begin{aligned} u(x)=o({|x|^{-\alpha / s}})\ \text{ as } \ |x|\rightarrow +\infty ,\; x_1<0 \end{aligned}$$
(1.7)

and A(x) is bounded below and symmetric in \(x_1\)-direction and satisfies

$$\begin{aligned} A(x_1,x')<A(\overline{x}_1,x'), \forall x_1>\overline{x}_1,\ (x_1,x'), (\overline{x}_1,x')\in \Omega . \end{aligned}$$

then there exists \(u_0\le 0\) such that \(u(x_1,x')\) is strictly increasing in \(\Omega \cap \{x_1<\mu _0\}\) in \(x_1\)-direction and

$$\begin{aligned} u(x_1,x')\le u(2\mu _0-x_1,x'), (x_1,x')\in \Omega \cap \{x_1<\mu _0\}, \end{aligned}$$
(1.8)

where

$$\begin{aligned} \mu _0=\sup _{\mu \le 0}\{\mu \mid u(x_1,x')\le u(2\lambda -x_1,x'),\forall (x_1,x')\in \Omega \cap \{x_1<\lambda \},\forall \lambda \le \mu \}. \end{aligned}$$

Furthermore, if \(\mu _0<0\), then \(u(x_1,x')= u(2\mu -x_1,x'),(x_1,x')\in \Omega \cap \{x_1<\mu _0\}\).

We next turn our attention to the follow equation

$$\begin{aligned} {\left\{ \begin{array}{ll} (-\Delta )^{\frac{\alpha }{2}}u+A(x)u=f(x,u,\nabla u), x\in \mathbb {R}^n_+,\\ u\equiv 0, x\in \mathbb {R}^n\backslash \mathbb {R}^n_+. \end{array}\right. } \end{aligned}$$
(1.9)

We have the next theorem.

Theorem 1.3

Suppose that \(u\in C(\mathbb {R}^n_+)\bigcap \mathcal C_{loc}^{1,1}(\mathbb {R}^n_+)\) is a nonnegative solution of (1.9), \(f(x,v,\mathbf{p} )\in \digamma \) satisfies

$$\begin{aligned} f(x',x_n,v,p_1,\cdots ,p_{n-1},p_n)\le f(x',\overline{x}_n,v,p_1,\cdots ,p_{n-1},-p_n), \end{aligned}$$

for any \(x_n,\; p_n\ge 0\), \(x_n\le \overline{x}_n\), and A(x) is bounded below and symmetric in \(x_1\)-direction and satisfies

$$\begin{aligned} A(x_1,x')<A(\overline{x}_1,x'), \forall x_1>\overline{x}_1,\ (x_1,x'), (\overline{x}_1,x')\in \Omega . \end{aligned}$$

Suppose that

$$\begin{aligned} \lim _{|x|\rightarrow \infty }u(x)=0 \end{aligned}$$
(1.10)

and

$$\begin{aligned} f(x,v,\mathbf {p})=0,\ if\ v=0. \end{aligned}$$
(1.11)

Then \(u\equiv 0\).

This paper is organized as follows. In Sect. 2, we give the proof of Theorem 1.1. Section 3 is devoted to the proof of Theorem 1.2. While in Sect. 4, we will refer to the paper [8][12] and give the proof of Theorem 1.3.

2 Proof of Theorem1.1

Firstly, we define some frequently used notations in this paper. For \(x'=(x_2,x_3,\cdots ,x_n)\) and \(\lambda \in \mathbb {R}\), write

$$\begin{aligned} x=(x_{1},x')\in \mathbb {R}^n,\; x^{\lambda }=(2\lambda -x_{1},x{'})\in \mathbb {R}^n. \end{aligned}$$

Let \(T_\lambda \) be a hyperplane in \(\mathbb {R}^n\), which is defined as

$$\begin{aligned} T_\lambda =\{x=(x_1,x')\in \mathbb {R}^n|x_1=\lambda \} \end{aligned}$$

and

$$\begin{aligned} \Sigma _{\lambda }=\{(x_{1},x')\in \mathbb {R}^n\mid {x_{1}<\lambda }\}, \;\Sigma ^{c}_{\lambda }=\{(x_{1},x{'})\in \mathbb {R}^n\mid {x_{1}\ge \lambda }\}. \end{aligned}$$

We set

$$\begin{aligned} u_\lambda (x)=u(x^\lambda ),\ \omega _\lambda (x)=u_\lambda (x)-u(x),\ \forall x\in \mathbb {R}^n. \end{aligned}$$

Sometimes, we may write \(\omega _\lambda (x)\) as \(\omega (x)\) if there is no ambiguity. Now we introduce a lemma, which was established in [14].

Lemma 2.1

Let \(\omega \in L_{\alpha }\) be a so called \(\lambda \)-antisymmetric function, i.e. \(\omega (x_1,x')=-\omega (2\lambda -x_1,x')\). Suppose that there exists \(x\in \Sigma _{\lambda }\) such that

$$\begin{aligned} \omega (x)=\underset{\Sigma _{\lambda }}{\inf } \omega (y)<0. \end{aligned}$$

If \(\omega \in C^{1,1}\)at x, then there exists some positive constant \(C_{n,\alpha }\) such that

$$\begin{aligned} (-\Delta )^{\frac{\alpha }{2}}\omega (x)\le C_{n,\alpha }\left( \delta ^{-\alpha }\omega (x)-\delta \int _{\Sigma _{\lambda }} \frac{({\omega }(y)-\omega (x))(\lambda -y_{1})}{|x-y^{\lambda }|^{n+\alpha +2}} dy\right) , \end{aligned}$$

for all \(\alpha \in (0,2)\), where \(\delta =d(x,T_{\lambda })=|x_1-\lambda |.\)

Proof of Theorem 1.1 We divide the proof into two steps. Firstly, we shall show that for \(\lambda >-1\) which is sufficiently close to \(-1\), we have

$$\begin{aligned} \omega _\lambda (x)\ge 0,\ \forall x\in \Sigma _\lambda . \end{aligned}$$
(2.1)

Next, we shall move the plane \(T_\lambda \) along the \(x_1\)-axis to the right in the left half of \(\Omega \) as long as inequality (2.1) holds. The plane will eventually stop at the limiting position \(\lambda =0\).

Step 1. Since \(\Omega \) is bounded and convex in \(x_1\)-direction, without loss of generality, we may assume that

$$\begin{aligned} \Omega \subset \{|x_1|\le 1\} \; \, \text{ and } \;\, \partial \Omega \cap \{x_1=-1\}\ne \emptyset . \end{aligned}$$

We claim that there exists \(\delta >0\) small enough such that for all \(\lambda \in [-1,-1+\delta )\), then

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _\lambda . \end{aligned}$$
(2.2)

If not, we set

$$\begin{aligned} \ell =\inf _{x\in \overline{\Sigma }_{ \begin{array}{c} \lambda \\ \lambda \end{array}\in [-1,-1+\delta )}}\omega _\lambda (x)<0. \end{aligned}$$

Since \(0\le u(x)\in C(\mathbb R^n)\) and \(u(x)\equiv 0, x\in \Omega ^c\), we know that for any \(\delta \in (0,1)\), \(\ell \) can obtained the infimum for some

$$\begin{aligned} (\lambda _0,x_0)\in \{(\lambda , x)|(\lambda , x)\in [-1, -1+\delta ]\times \overline{\Sigma _{\lambda }\cap \Omega }\}. \end{aligned}$$

Thus, \(\omega _{\lambda _0}\ge 0\) on \(\partial (\Omega \cap \Sigma _{\lambda _0})\), which yields that \(x_0\in \Sigma _{\lambda _0}\cap \Omega \). Since \(\lambda =-1\) implies \(\Sigma _{-1}\cap \Omega =\partial \Sigma _{-1}\cap \Omega \), one gets \(\omega _1(x)\ge 0\) in \(\Sigma _{-1}\cap \Omega \). Therefore, \(\lambda _0>-1\).

Now we set \(u_{x_i}(x)=\frac{\partial u}{\partial {x_i}}(x)\) for \(i=1,2,\ldots ,n\). Since \((\lambda _0,x_0)\) is point of minimum, we know that for \(\lambda _0\in (-1,-1+\delta )\),

$$\begin{aligned} \left. \frac{\partial \omega _\lambda (x)}{\partial \lambda }\right| _{(\lambda _0,x_0)}=0, \;\text {and}\; \left. \frac{\partial \omega _\lambda (x)}{\partial \lambda }\right| _{(-1+\delta ,x_0)}\le 0. \end{aligned}$$

Moreover, we also have (a) \(u_{x_1}(x_0^{\lambda _0})\le 0\). This is because that

$$\begin{aligned}&0\ge \left. \frac{\partial \omega _\lambda (x)}{\partial \lambda }\right| _{(\lambda _0,x_0)}\\&\quad =\left. \frac{\partial }{\partial \lambda }(u(x^\lambda )-u(x))\right| _{(\lambda _0,x_0)}\\&\quad =2u_{x_1}(x^\lambda )|_{(\lambda _0,x_0)}; \end{aligned}$$

(b) \( \nabla _x\omega (x_0^{\lambda _0}) =0\), where \(\nabla _x\) denotes the gradient operator with respective to x. From (b), we know that

$$\begin{aligned} \nabla _x u(x_0^{\lambda _0})=\nabla _x u(x_0) , \end{aligned}$$

which yields

$$\begin{aligned} u_{x_i}(x_0^{\lambda _0})=u_{x_i}(x_0) \ \text{ for }\ i=2, \cdots ,n \end{aligned}$$
(2.3)

and

$$\begin{aligned} u_{x_1}(x_0^{\lambda _0})=-u_{x_1}(x_0). \end{aligned}$$
(2.4)

Moreover, (a) implies that

$$\begin{aligned} u_{x_n}(x_0)\ge 0. \end{aligned}$$
(2.5)

By using Property (1.6) of \(f(x,u,\mathbf{p} )\), Conditions (2.3)-(2.5) and the fact that A(x) satisfies \(A(x_1,x')<A(\overline{x}_1,x')\) for all \(x_1>\overline{x}_1, \ (x_1,x'), (\overline{x}_1,x')\in \Omega ,\) we know from (1.5) that

$$\begin{aligned} \nonumber (-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})&=f(x_0^{\lambda _0},u_{\lambda _0}(x_0),\nabla _xu(x_0^{\lambda _0}))-f(x_0,u(x_0),\nabla _xu(x_0))\\ \nonumber&\quad +A(x_{0})u(x_{0})-A(x_0^{\lambda _{0}})u(x_0^{\lambda _{0}})\\ \nonumber&\ge f(x_0,u_{\lambda _0}(x_0),\nabla _xu(x_0))-f(x_0,u(x_0),\nabla _xu(x_0))\\ \nonumber&\quad +A(x_{0})u(x_{0})-A(x_0)u(x_0^{\lambda _{0}})\\&=c(x_0)\omega _{\lambda _0}(x_0)-A(x_0)\omega _{\lambda _{0}}(x_{0}), \end{aligned}$$
(2.6)

where

$$\begin{aligned} c(x_0)=\frac{f(x_0,u_{\lambda _0}(x_0),\nabla _xu(x_0))-f(x_0,u(x_0),\nabla _xu(x_0))}{u_{\lambda _0}(x_0)-u(x_0)}. \end{aligned}$$

Since \(u(x)\in C(\mathbb {R}^n)\) with compact support and \(f(x,u,\mathbf{p} )\in \digamma \), we get that c(x) is uniformly bounded.

Moreover, since A(x) is bounded from below, by using Lemma 1 we have

$$\begin{aligned} (-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0) \nonumber \\\le (\frac{C}{\delta ^\alpha }+A(x_{0})-c(x_0))\omega _{\lambda _0}(x_0)<0. \end{aligned}$$
(2.7)

Take \(\delta \) small enough, we arrive at a contradiction with (2.6). Hence, Claim (2.2) holds.

Step 2. Set

$$\begin{aligned} \lambda _0=\sup _{-1\le \lambda \le 0}\{\lambda \mid \omega _\mu (x)\ge 0,\forall x\in \Sigma _\mu ,\forall \mu \le \lambda \}. \end{aligned}$$

Then we have \(\lambda _0\le 0\). Indeed, we must have \(\lambda _0=0\). Otherwise, suppose that \(\lambda _0<0\), since \(\omega _{\lambda _0}\not \equiv 0\), we will have

$$\begin{aligned} \omega _{\lambda _0}>0,\; x\in \Sigma _{\lambda _0}\cap \Omega . \end{aligned}$$
(2.8)

If not, there would exist \(x_0\in \Omega \cap \Sigma _{\lambda _0}\) such that \(\omega _{\lambda _0}(x_0)=0\). By using (2.6) we have

$$\begin{aligned} (-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0) \ge 0. \end{aligned}$$

On the other hand, we have

$$\begin{aligned}&(-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})\nonumber \\&=C_{n,\alpha } \text {P.V.} \int _{\mathbb {R}^n} \frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)}{|x-y|^{n+\alpha }} dy\nonumber \\&=C_{n,\alpha }P.V.\int _{ \Sigma _{\lambda }}\frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)}{|x-y|^{n+\alpha }} dy+C_{n,\alpha }P.V.\int _{ \Sigma _{\lambda }^c}\frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)}{|x-y|^{n+\alpha }} dy\nonumber \\&=C_{n,\alpha }P.V.\int _{ \Sigma _{\lambda }}\frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)}{|x-y|^{n+\alpha }} dy+C_{n,\alpha }P.V.\int _{ \Sigma _{\lambda }}\frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y^{\lambda })}{|x-y^{\lambda }|^{n+\alpha }} dy\nonumber \\&=C_{n,\alpha }P.V.\int _{ \Sigma _{\lambda }}\frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)}{|x-y|^{n+\alpha }} dy+C_{n,\alpha }P.V.\int _{ \Sigma _{\lambda }}\frac{\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)}{|x-y^{\lambda }|^{n+\alpha }} dy\nonumber \\&=C_{n,\alpha }P.V.\int _{\Sigma _{\lambda }}\Big (\frac{1}{|x-y|^{n+\alpha }}-\frac{1}{|x-y^{\lambda }|^{n+\alpha }}\Big ) \Big (\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)\Big )dy\nonumber \\&\quad +C_{n,\alpha }\int _{\Sigma _{\lambda }}\frac{2\omega _{\lambda _{0}}(x_{0})}{|x-y^{\lambda }|^{n+\alpha }}dy\nonumber \\&:=I_1+I_2. \end{aligned}$$
(2.9)

The fact \(\omega _{\lambda _0}(x_0)=0\) gives \(I_2=0\). Then

$$\begin{aligned}&(-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0)\\&\le C_{n,\alpha }P.V.\int _{\Sigma _{\lambda }}\Big (\frac{1}{|x-y|^{n+\alpha }}-\frac{1}{|x-y^{\lambda }|^{n+\alpha }}\Big ) \Big (\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)\Big )dy <0, \end{aligned}$$

where

$$\begin{aligned} \frac{1}{|x-y|^{n+\alpha }}>\frac{1}{|x-y^{\lambda }|^{n+\alpha }},\;\forall x,y\in \Sigma _{\lambda }. \end{aligned}$$

Since \(\omega _{\lambda _0}(y)\ge 0\) and \(\omega _{\lambda _0}(y)\not \equiv 0\), which yields a contradiction, thus (2.8) holds.

We claim that there exists \(\epsilon >0\) small enough such that

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _\lambda \cap \Omega ,\forall \lambda \in [\lambda _0,\lambda _0+\epsilon ). \end{aligned}$$
(2.10)

Now we prove (2.10) is true by contradiction. Indeed, combining (2.8) and the fact that \(\omega _\lambda (x)\) is lower semi-continuous in \(\Omega \), for any \(\delta >0\) we have

$$\begin{aligned} \omega _{\lambda _0}(x)\ge C_\delta >0,\forall x\in \Sigma _{\lambda _0-\delta }\cap \Omega . \end{aligned}$$

By the continuity of \(\omega _\lambda \) with respect to \(\lambda \), there exist \(\epsilon >0\) such that

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _{\lambda _0-\delta }\cap \Omega ,\forall \lambda \in [\lambda _0,\lambda _0+\epsilon ). \end{aligned}$$
(2.11)

Suppose (2.10) is not true, then for all \(\epsilon >0\) we have

$$\begin{aligned} \ell _\epsilon =\inf _{x\in \overline{\begin{array}{c} {\Sigma _{\lambda }}\\ \lambda \end{array}}\in [\lambda _0,\lambda _0+\epsilon ]}\omega _\lambda (x)<0. \end{aligned}$$

Since \(\Omega \) is a bounded domain, then by (2.11), \(\ell _\epsilon \) can obtain the infimum for some point \((\mu ,x_0)\in \{(\lambda ,x)|(\lambda ,x)\in [\lambda _0,\lambda _0+\epsilon ]\times \overline{(\Sigma _\lambda \backslash \Sigma _{{\lambda _0}-\delta })\cap \Omega }\}\), i.e.,

$$\begin{aligned} \omega _\mu (x_0)=\ell _\epsilon <0. \end{aligned}$$

Obviously we have \(\omega _{\mu }\ge 0\) on \(\partial (\Sigma _\mu \backslash \Sigma _{{\lambda _0}-\delta }\cap \Omega )\), thus \(x_0\in \Sigma _\mu \backslash \Sigma _{{\lambda _0}-\delta }\cap \Omega \). By similar arguments as in Step 1 (2.6)-(2.7), we have

$$\begin{aligned} \nonumber 0&\le (-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0)\\&\nonumber \le (\frac{C}{(\delta +\epsilon )^\alpha }+A(x_{0})-c(x_0))\omega _\mu (x_0)<0, \end{aligned}$$

if \(\delta \), \(\epsilon \) are chosen small enough. This yields a contradiction, so (2.10) holds. Therefore, we must have \(\lambda _0=0\). It follows that

$$\begin{aligned} \omega _0(x)\ge 0,\; x\in \Sigma _0, \end{aligned}$$

which implies

$$\begin{aligned} u(x_1,x')\le u(-x_1,x'), \forall x_1\le 0,\ (x_1,x')\in \Omega . \end{aligned}$$

The proof of Theorem 1.1 is completed. \(\square \)

3 Proof of Theorem 1.2

We divide the proof into two steps. To begin with, we shall show that for \(\lambda <0\),

$$\begin{aligned} \omega _\lambda (x)\ge 0,\ \forall x\in \Sigma _\lambda . \end{aligned}$$
(3.1)

Then we move the plane \(T_\lambda \) along the \(x_1\)-axis to the right in the left half of \(\Omega \) as long as inequality (3.1) holds. The plane will eventually stop at some limit position \(\lambda =\lambda _0<0\) or \(\lambda =0\).

Step 1. Start moving the plane \(T_\lambda \) along the \(x_1\)-axis from near \(-\infty \) to the right.

Claim: There exists \(R_0>0\) large enough such that

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _\lambda ,\forall \lambda \le -R_0. \end{aligned}$$
(3.2)

Suppose the claim is not true, then there exist \(\lambda _k\rightarrow -\infty \) such that

$$\begin{aligned} \ell =\inf _{x\in \overline{\begin{array}{c} {\Sigma }_{\mu _k}\\ \lambda \end{array}}\in (-\infty ,\lambda _k]}\omega _\lambda (x)<0. \end{aligned}$$

Since u(x) vanishes at infinity, then we have \(\omega _\lambda (x)=u_\lambda (x)-u(x)\ge -u(x)\ge \frac{\ell }{2}\) for any \(|x|\ge R_k\) and \(R_k\) large enough. Hence, we can get

$$\begin{aligned} \omega _{\mu _k}(x^k)=\ell \ \text {for some}\ \mu _k\in [-|x^k|,\lambda _k]\;\text {and}\;|x^k|\le R_k. \end{aligned}$$

On the one hand, similarly as (2.9), we may write \((-\Delta )^{\frac{\alpha }{2}}\omega _{\mu _k}(x^{k}):=I_1+I_2\). We also have \(I_1\le 0\), then by similar arguments as in (2.9)-(2.11), we get

$$\begin{aligned} (-\Delta )^{\frac{\alpha }{2}}\omega _{\mu _k}(x^{k}) \le I_2 \le 2c_0C_{n,\alpha }\int _{\Sigma _{\mu _k}}\frac{\omega _{\mu _k}(x^k)}{|x^k-y^{\mu _k}|^{n+\alpha }}dy \le \frac{C\omega _{\mu _k}(x^k)}{|x^k|^\alpha }. \end{aligned}$$
(3.3)

To obtain the last inequality, we notice that \(0<u_{\mu _k}(x^k)<u(x^k)\). For \(x^k\in \Sigma _{\mu _k}\) and \(|x^k|\) sufficiently large, let \(\Sigma _{\mu _k}^c=\mathbb {R}^n\backslash \Sigma _{\mu _k},\) choose a point in \(\Sigma _{\mu _k}^c:x^m=(3|x^k|+x_1^k,(x^k)')\), then \(B_{|x^k|}(x^m)\subset \Sigma _{\mu _k}^c\). There exists a \(C>0\) such that

$$\begin{aligned} \int _{\Sigma _{\mu _k}}\frac{1}{|x^k-y^{\mu _k}|^{n+\alpha }}dy&\ge \int _{\Sigma _{\mu _k}^c}\frac{1}{|x^k-y|^{n+\alpha }}dy\\ \nonumber&\ge \int _{B_{|x^k|}(x^m)}\frac{1}{|x^k-y|^{n+\alpha }}dy\sim \frac{C}{|x^k|^\alpha }. \end{aligned}$$
(3.4)

On the other hand, from (2.6), we have

$$\begin{aligned} c(x_k)\omega _{\mu _k}(x_k)-A(x_k)\omega _{\mu _k}(x_{k})\le (-\Delta )^{\frac{\alpha }{2}}\omega _{\mu _k}(x^{k}), \end{aligned}$$

where

$$\begin{aligned} c(x^k)&=\frac{f(x^k,u_{\mu _k}(x^k),\nabla u(x^k))-f(x_0,u(x^k),\nabla u(x^k))}{u_{\mu _k}(x^k)-u(x^k)}\\&\le C(|u_{\mu _k}(x^k)|^s+|u(x^k)|^s)\le C|u(x^k)|^s, \end{aligned}$$

and A(x) is bounded below, then we have

$$\begin{aligned} c(x^k)-A(x^k)&\le C|u(x^k)|^s. \end{aligned}$$
(3.5)

Then from (1.7) and (3.3)-(3.4), we can get

$$\begin{aligned} C|u(x^k)|^s\omega _{\mu _k}(x^k)\le (-\Delta )^{\frac{\alpha }{2}}\omega _{\mu _k}(x^{k})\le \frac{C\omega _{\mu _k}(x^k)}{|x^k|^\alpha }, \end{aligned}$$

that is

$$\begin{aligned} o(1)=C|x^k|^\alpha u^s(x^k)\ge \widetilde{C}_{n,\alpha }, \end{aligned}$$

since \(|x^k|\ge |\lambda _k|\rightarrow \infty \). This yields a contradiction and ends the proof of Step 1.

Step 2. Set

$$\begin{aligned} \lambda _0=\sup _{\lambda \le 0}\{\lambda \mid \omega _\mu (x)\ge 0,\forall x\in \Sigma _\mu ,\forall \mu \le \lambda \}. \end{aligned}$$

By the definition of \(\lambda _0\) and the continuity of u(x), we have \(\omega _{\lambda _0}(x)\ge 0\) for all \(x\in \Sigma _{\lambda _0}.\) Then we must have

$$\begin{aligned} \lambda _0=0 \end{aligned}$$

or

$$\begin{aligned} \ \lambda _0<0\ \text {and} \ \omega _{\lambda _0}(x)\equiv 0. \end{aligned}$$

The second case happens only when \(\sup _{x\in \Omega }|x_1|=+\infty \). If not, we suppose that \(\lambda _0<0\) but \(\omega _{\lambda _0}(x)\not \equiv 0\). We claim that there exists \(\delta >0\) small enough such that

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _\lambda ,\forall \lambda \in [\lambda _0,\lambda _0+\delta ). \end{aligned}$$
(3.6)

Indeed, for \(x\in \Sigma _{\lambda _0}\cap \Omega ^c\), it is easy to see that \(\omega _{\lambda _0}(x)\ge 0\). Then by similar arguments as in the proof of (2.8), we have

$$\begin{aligned} \omega _{\lambda _0}>0,\; x\in \Sigma _{\lambda _0}\cap B_{R_0}(0)\cap \Omega . \end{aligned}$$

It follows that for any positive real number \(\sigma \),

$$\begin{aligned} \omega _{\lambda _0}(x)\ge C_\delta >0,\forall x\in \Sigma _{\lambda _0-\sigma }\cap B_{R_0}(0)\cap \Omega , \end{aligned}$$
(3.7)

where \(R_0\) is defined in Step 1. Since \(\omega _\lambda \) is continuous with respect to \(\lambda \), for all sufficiently small \(\delta >0\), we have

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _{\lambda _0-\sigma }\cap B_{R_0}(0)\cap \Omega ,\ \forall \lambda \in [\lambda _0,\lambda _0+\delta ). \end{aligned}$$
(3.8)

Suppose that (3.6) is false, then we have for any \(\delta >0\),

$$\begin{aligned} \ell =\inf _{x\in \overline{\begin{array}{c} {\Sigma _\lambda }\\ \lambda \end{array}}\in [\lambda _0,\lambda _0+\delta ]}\omega _\lambda (x)<0. \end{aligned}$$

By Step 1, we know that the point of minimum can not be obtained in \(B_{R_0}^c\). By using (3.8), \(\ell \) can obtain the minimum at some point

$$\begin{aligned} (\mu _k,x_0)\in \{(\lambda ,x)|(\lambda ,x)\in [\lambda _0,\lambda _0+\epsilon ]\times \overline{(\Sigma _\lambda \backslash \Sigma _{{\lambda _0}-\sigma })\cap B_{R_0}(0)}\}. \end{aligned}$$

Obviously we have \(\omega _{\lambda _0}\ge \frac{\ell }{2}\) on \(\partial ((\Sigma _\lambda \backslash \Sigma _{{\lambda _0}-\delta })\cap B_{R_0})\), thus

$$\begin{aligned} x_0\in (\Sigma _\lambda \backslash \Sigma _{{\lambda _0}-\delta })\cap B_{R_0}(0). \end{aligned}$$

By virtue of similar arguments as in Theorem 1 (2.6)-(2.7), one can see that

$$\begin{aligned} 0\le (-\Delta )^{\frac{\alpha }{2}}\omega _{\mu _0}(x^{0})\le (\frac{C}{{(\delta +\sigma )}^\alpha }+A(x^{0})-c(x^0))\omega _{\mu _0}(x^0)<0, \end{aligned}$$

if \(\delta \), \(\sigma \) are chosen small enough. This yields a contradiction, hence (3.6) holds, which contradicts with the definition of \(\lambda _0\). The strict monotonicity follows from the fact that \(\omega _\lambda (x)>0\) in \(\Sigma _\lambda \cap \Omega \) for all \(\lambda <\lambda _0\).

The proof of Theorem 1.2 is completed.

4 Proof of Theorem 1.3

We first claim that

$$\begin{aligned} u(x)>0,x\in \mathbb {R}^n_+ \ or\ u(x)\equiv 0,x\in \mathbb {R}^n_+. \end{aligned}$$
(4.1)

To prove (4.1), we assume that there exist \(x_0\in \mathbb {R}^n_+ \) such that \(u(x_0)=0\), then by (1.11), we have

$$\begin{aligned} (-\Delta )^{\frac{\alpha }{2}}u(x_{0})+A(x_{0})u(x_{0})=f(x,u,\nabla u)=0. \end{aligned}$$

On the other hand, by the definition (1.2), we have

$$\begin{aligned} (-\Delta )^{\frac{\alpha }{2}}u(x_{0})&= C_{n,\alpha } P.V. \int _{\mathbb {R}^n} \frac{u(x_0)-u(y)}{|x_0-y|^{n+\alpha }} dy\\&=-C_{n,\alpha }P.V.\int _{\mathbb {R}^n_+}\frac{u(y)}{|x_0-y|^{n+\alpha }}dy <0, \end{aligned}$$

where the last inequality follows from the fact that \(u(y)\ge 0\) and \(u(y)\not \equiv 0\) with \(u\in C(\mathbb {R}^n_{+})\). This yields a contradiction, so (4.1) holds. Now we assume that

$$\begin{aligned} u(x)>0,x\in \mathbb {R}^n_+. \end{aligned}$$
(4.2)

We carry out the method of moving planes on the solution u along \(x_n\)-direction. Let \(T_\lambda \) be a hyperplane in \(\mathbb {R}^n\) defined by

$$\begin{aligned} T_\lambda =\{x=(x',x_n)\in \mathbb {R}^n|x_n=\lambda ,\lambda >0\}, \end{aligned}$$

where \(x'=(x_1,x_2,\cdots ,x_{n-1})\). Let \( x^\lambda =(x_1,\cdots ,x_{n-1},2\lambda -x_n) \) be the reflection of x with respect to the hyperplane \(T_\lambda \). Set

$$\begin{aligned} \Sigma _\lambda = \{x\in \mathbb {R}^n|0<x_n<\lambda \} \end{aligned}$$

and

$$\begin{aligned} u_\lambda (x)=u(x^\lambda ),\ \omega _\lambda (x)=u_\lambda (x)-u(x),\ \forall x\in \mathbb {R}^n. \end{aligned}$$

We divide the proof into two steps: Step 1. We claim that there exists \(\delta > 0\) small enough such that

$$\begin{aligned} \omega _\lambda (x)\ge 0, \,\ \forall x\in \Sigma _\lambda ,\ \forall \lambda \in [0,\delta ]. \end{aligned}$$
(4.3)

Suppose that the claim is not true, then we have for any \( \delta >0\) (without loss of generality, we can restrict \(\delta \in (0,1)\)),

$$\begin{aligned} \ell =\inf _{x\in \overline{\begin{array}{c} {\Sigma _\lambda }\\ \lambda \end{array}}\in [0,\delta ]}\omega _\lambda (x)<0. \end{aligned}$$
(4.4)

First we know that \(\omega _\lambda (x)=u_\lambda (x)-u(x)\ge -u(x)\) uniformly with respect to \(\lambda \), so by (1.10), we have

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\omega _\lambda (x)\ge \lim _{|x|\rightarrow \infty }-u(x)=0 \end{aligned}$$

uniformly with respect to \(\lambda \). Then there must exist \(R_0>0\) large enough such that \(\omega _{\lambda }(x)\ge \frac{\ell }{2}, x\in B_{R_0}^c(0)\) holds uniformly with respect to \(\lambda \). So \(\ell \) in (4.4) can obtain the infimum at some point

$$\begin{aligned} (\lambda _0,x_0)\in \{(\lambda ,x)|(\lambda ,x)\in [0,\delta ]\times \overline{(\Sigma _\lambda \backslash \Sigma _0)\cap B_{R_0}(0)}\}. \end{aligned}$$

Notice that \(u\equiv 0 \ \text{ in }\ \mathbb {R}^n_{-},\; u\ge 0 \ \text{ in } \ \mathbb {R}^n\) and \( \omega _{\lambda _0}(x)\ge \frac{\ell }{2}, x\in B_{R_0}^c(0)\), we have \(\omega _{\lambda _0}\ge \frac{\ell }{2}\) on \(\partial ((\Sigma _{\lambda _0}\backslash \Sigma _0)\cap B_{R_0}(0))\), so \(x_0\) must be in \(B_{R_0}(0)\cap (\Sigma _{\lambda _0}\backslash \Sigma _0)\). On the other hand, it is easy to see that \(\lambda _0\in (0, \delta ]\). Since \((\lambda _0,x_0)\) is a point of minimum, we can get (a) \(u_{x_n}(x_0^{\lambda _0})\le 0\); (b) \(\nabla _x\omega (x_0^{\lambda _0}) =0\).

Then by similar arguments as in the proof of Theorem 1 Step 1 (2.6)-(2.7), we get

$$\begin{aligned} 0&\le (-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0)\nonumber \\&\le (\frac{C}{\delta ^\alpha }+A(x_{0})-c(x_0))\omega _{\lambda _0}(x_0)<0, \end{aligned}$$
(4.5)

if we choose \(\delta \) small enough. This yields a contradiction, so (4.3) is true.

\(\mathbf{Step}\ 2. \) Set

$$\begin{aligned} \lambda _0=\sup \{\lambda >0|\omega _\mu (x)\ge 0,\forall x\in \Sigma _\mu ,\forall \mu \le \lambda \}. \end{aligned}$$

We claim that

$$\begin{aligned} \lambda _0=\infty . \end{aligned}$$
(4.6)

Otherwise, if \(\lambda _0<\infty \), then by the definition of \(\lambda _0\), we have

$$\begin{aligned} \omega _{\lambda _0}>0,x\in \Sigma _{\lambda _0}\ or\ \omega _{\lambda _0}\equiv 0,x\in \Sigma _{\lambda _0}. \end{aligned}$$
(4.7)

To prove (4.7), we assume there exists \(x_0\in \Sigma _{\lambda _0}\), such that \(\omega _{\lambda _0}(x_0)=0\). Then by using (2.6), we have

$$\begin{aligned}&(-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0)\ge 0. \end{aligned}$$

On the other hand, by using Lemma 2.1 and \(\omega _{\lambda _0}\ge 0\), we can write \((-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})=I_1+I_2\). The fact \(\omega _{\lambda _0}(x_0)=0\) gives \(I_2=0\). We have

$$\begin{aligned}&(-\Delta )^{\frac{\alpha }{2}}\omega _{\lambda _{0}}(x_{0})+A(x_0)\omega _{\lambda _{0}}(x_{0})-c(x_0)\omega _{\lambda _0}(x_0)\\&\quad \le C_{n,\alpha }P.V.\int _{\Sigma _{\lambda }}\Big (\frac{1}{|x-y|^{n+\alpha }}-\frac{1}{|x-y^{\lambda }|^{n+\alpha }}\big ) \Big (\omega _{\lambda _{0}}(x_{0})-\omega _{\lambda _{0}}(y)\Big )dy <0. \end{aligned}$$

Since \(\omega _{\lambda _0}(y)\not \equiv 0\), it yields a contradiction. So (4.7) holds.

Now we suppose that

$$\begin{aligned} \omega _{\lambda _0}(x)>0,\; x\in \Sigma _{\lambda _0}. \end{aligned}$$
(4.8)

We first claim that there exists \(\epsilon >0\) small enough such that

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \Sigma _\lambda ,\forall \lambda \in [\lambda _0,\lambda _0+\epsilon ). \end{aligned}$$
(4.9)

Indeed, since \(\omega _\lambda (x)\) is lower semi-continuous in \(\Omega \), it follows by (4.8) that for any \(\delta >0\) and \(R>0\),

$$\begin{aligned} \omega _{\lambda _0}(x)\ge C_\delta >0,\forall x\in \overline{\Sigma _{\lambda _0-\delta }\cap B_R(0)}. \end{aligned}$$

By the continuity of \(\omega _\lambda \) with respect to \(\lambda \), there exist \(\epsilon >0\) such that

$$\begin{aligned} \omega _\lambda (x)\ge 0,\forall x\in \overline{\Sigma _{\lambda _0-\delta }\cap B_R(0)},\forall \lambda \in [\lambda _0,\lambda _0+\epsilon ). \end{aligned}$$
(4.10)

Suppose (4.9) is not true, then for any \(\epsilon >0\) we have

$$\begin{aligned} \ell =\inf _{x\in \overline{\begin{array}{c} {\Sigma _\lambda }\\ \lambda \end{array}}\in [\lambda _0,\lambda _0+\epsilon ]}\omega _\lambda (x)<0. \end{aligned}$$

Since \(\omega _\lambda (x)=u_\lambda (x)-u(x)\ge -u(x)\), so by (1.10), we have

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\omega _\lambda (x)\ge \lim _{|x|\rightarrow \infty }-u(x)=0. \end{aligned}$$

So there must exist \(R_0>0\) large enough such that

$$\begin{aligned} \lim _{|x|\rightarrow \infty }\omega _{\lambda _0}(x)\ge 0, x\in B_{R_0}^c(0)\cap \Sigma _\lambda . \end{aligned}$$

By virtue of (4.10), \(\ell \) can obtain the infimum at some point

$$\begin{aligned} (\mu _k,x_0)\in \{(\lambda ,x)|(\lambda ,x)\in [\lambda _0,\lambda _0+\epsilon ]\times \overline{(\Sigma _\lambda \backslash \Sigma _{{\lambda _0}-\delta })\cap B_{R_0}(0)}\}.\end{aligned}$$

By the same arguments as in Theorem 1 (2.6)-(2.7), it is easy to see that

$$\begin{aligned} 0\le (-\Delta )^{\frac{\alpha }{2}}\omega _{\mu _k}(x_0)-c(x_0)\omega _{\mu _k}(x_0)\le \Big (\frac{C}{(\delta +\epsilon )^\alpha }+A(x_{0})-c(x_0)\Big )\omega _{\mu _k}(x_0)<0, \end{aligned}$$

if \(\delta \), \(\epsilon \) are chosen small enough. This yields a contradiction and proves claim (4.9). However, this contradicts with the definition of \(\lambda _0\). So (4.8) is not true. Therefore, we must have

$$\begin{aligned} \omega _{\lambda _0}\equiv 0,\;&x\in \Sigma _{\lambda _0}, \end{aligned}$$

or equivalently,

$$\begin{aligned} u_{\lambda _0}(x)=u(x),\;&x\in \Sigma _{\lambda _0}. \end{aligned}$$

If we choose the point \(\bar{x}=(x_1,x_2,\cdots ,x_{n-1},0) \) in the hyperplane \(\{x_n=0\}\), then \(\bar{x}^{\lambda }\in \mathbb R^{n}_+\) and

$$\begin{aligned} u(\bar{x}^{\lambda })=u(x_1,x_2,\cdots ,x_{n-1},2\lambda _0)=u(x_1,x_2,\cdots ,x_{n-1},0)=0. \end{aligned}$$

This contradict with the assumption (4.2).

Therefore, we have proved Claim (4.6): \(\lambda _0=\infty \). Consequently, the solution u(x) is monotonically increasing with respect to \(x_n\). Recall that the condition (1.10) tells us \(\lim _{|x|\rightarrow \infty }u(x)=0\). So Claim (4.6) is not true, thus \(u(x)\equiv 0, x\in \mathbb {R}^n_+\).

The proof of Theorem 1.3 is completed. \(\square \)