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

Singular semilinear elliptic equations in half-spaces

Phuong Le Phuong Le1,2 (ORCID: 0000-0003-4724-7118)
1Faculty of Economic Mathematics, University of Economics and Law, Ho Chi Minh City, Vietnam;
2Vietnam National University, Ho Chi Minh City, Vietnam
phuongl@uel.edu.vn
Abstract.

We prove the monotonicity of positive solutions to the problem Δu=f(u)-\Delta u=f(u) in +N:={(x,xN)NxN>0}\mathbb{R}^{N}_{+}:=\{(x^{\prime},x_{N})\in\mathbb{R}^{N}\mid x_{N}>0\} under zero Dirichlet boundary condition with a possible singular nonlinearity ff. In some situations, we can derive a precise estimate on the blow-up rate of uη\frac{\partial u}{\partial\eta} as xN0+x_{N}\to 0^{+}, where (η,eN)>0(\eta,e_{N})>0, and obtain a classification result. The main tools we use are the method of moving planes and the sliding method.

Key words and phrases:
semilinear elliptic equation, singular nonlinearity, half-space, monotonicity, rigidity
2020 Mathematics Subject Classification:
35J61, 35J75, 35B06, 35B09

1. Introduction

The monotonicity and symmetry of solutions to the semilinear elliptic problem

{Δu=f(u) in +N,u>0 in +N,u=0 on +N,\begin{cases}-\Delta u=f(u)&\text{ in }\mathbb{R}^{N}_{+},\\ u>0&\text{ in }\mathbb{R}^{N}_{+},\\ u=0&\text{ on }\partial\mathbb{R}^{N}_{+},\end{cases} (1)

where

+N:={x:=(x,xN)NxN>0},\mathbb{R}^{N}_{+}:=\{x:=(x^{\prime},x_{N})\in\mathbb{R}^{N}\mid x_{N}>0\},

are well studied in the literature. Berestycki, Caffarelli, and Nirenberg [6, 4] demonstrated that if f:[0,+)f:[0,+\infty)\to\mathbb{R} is a Lipschitz function with f(0)0f(0)\geq 0, then any classical solution of (1) is monotone in the xNx_{N}-direction. When ff is only locally Lipschitz continuous on [0,+)[0,+\infty), a similar monotonicity result can be obtained for solutions that are bounded on all strips Σλ:={(x,xN)N0<xN<λ}\Sigma_{\lambda}:=\{(x^{\prime},x_{N})\in\mathbb{R}^{N}\mid 0<x_{N}<\lambda\} (λ>0\lambda>0), as shown in [17, 27]. The case where f(0)<0f(0)<0 is more complex, and a complete proof of monotonicity for solutions in this scenario is currently only available for dimension N=2N=2, as detailed in [18, 19]. For results on symmetry of solutions, which is usually called rigidity in the literature, we refer to [7, 4, 2, 13, 5] and the references therein.

In this paper, we are mainly interested in problem (1) with singular nonlinearity at zero in the sense that f:(0,+)f:(0,+\infty) is locally Lipschitz continuous and limt0+f(t)=+\lim_{t\to 0^{+}}f(t)=+\infty. A model problem is given by

{Δu=1uγ+g(u) in +N,u>0 in +N,u=0 on +N,\begin{cases}-\Delta u=\dfrac{1}{u^{\gamma}}+g(u)&\text{ in }\mathbb{R}^{N}_{+},\\ u>0&\text{ in }\mathbb{R}^{N}_{+},\\ u=0&\text{ on }\partial\mathbb{R}^{N}_{+},\end{cases} (2)

where γ>0\gamma>0 and g:[0,+)g:[0,+\infty) is a locally Lipschitz continuous function. It’s well established that solutions to problem (2) are generally not smooth up to the boundary. In fact, it was shown in [23] and also in Theorem 2 below that the gradient of solutions becomes unbounded at the boundary. Given the natural regularity behavior of these solutions (as discussed in [14]), we focus on solutions uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) to (1). Consequently, the equation is well-defined in the classical sense within the domain’s interior.

Since the seminal paper [14], singular semilinear elliptic problems

{Δu=1uγ+g(u) in Ω,u>0 in Ω,u=0 on Ω,\begin{cases}-\Delta u=\dfrac{1}{u^{\gamma}}+g(u)&\text{ in }\Omega,\\ u>0&\text{ in }\Omega,\\ u=0&\text{ on }\Omega,\end{cases} (3)

where ΩN\Omega\subset\mathbb{R}^{N} is a bounded domain, have been extensively studied from various perspectives. We specifically reference the works [3, 9, 12, 10, 21, 22, 23, 15], which are closely related to our research. A key focus in the study of these equations is understanding the behavior of solutions near the boundary, where they often lose regularity. A generalized version of the Höpf boundary lemma was obtained in [11]. The symmetry of solutions was studied in [15] (see also [16, 24] and the references therein).

As demonstrated in [11], to obtain the Höpf boundary lemma for (3), one may exploit a scaling argument near the boundary which leads to the study of a limiting problem in the half-space

{Δu=1uγ in +N,u>0 in +N,u=0 on +N,\begin{cases}-\Delta u=\dfrac{1}{u^{\gamma}}&\text{ in }\mathbb{R}^{N}_{+},\\ u>0&\text{ in }\mathbb{R}^{N}_{+},\\ u=0&\text{ on }\partial\mathbb{R}^{N}_{+},\end{cases} (4)

which is exactly problem (1) with f0f\equiv 0. Solutions to problem (4) have been classified recently in elegant papers [25, 26]. These results reveal that all weak solutions to (4) with γ>1\gamma>1 must be either of the form

u(x)(γ+1)2γ+1(2γ2)1γ+1xN2γ+1u(x)\equiv\frac{(\gamma+1)^{\frac{2}{\gamma+1}}}{(2\gamma-2)^{\frac{1}{\gamma+1}}}x_{N}^{\frac{2}{\gamma+1}}

or of the form

u(x)λ2γ+1v(λxN)u(x)\equiv\lambda^{-\frac{2}{\gamma+1}}v(\lambda x_{N})

where λ>0\lambda>0 and vC2(+)C(+¯)v\in C^{2}(\mathbb{R}_{+})\cap C(\overline{\mathbb{R}_{+}}) is the unique solution to

{v′′=1vγ,t>0,v(t)>0,t>0,v(0)=0,limt+v(t)=1.\begin{cases}-v^{\prime\prime}=\dfrac{1}{v^{\gamma}},&t>0,\\ v(t)>0,&t>0,\\ v(0)=0,~{}\lim_{t\to+\infty}v^{\prime}(t)=1.\end{cases}

One notable feature of problem (4) is that its nonlinearity is decreasing on (0,+)(0,+\infty). Hence the weak comparison principle holds in large subdomains of +N\mathbb{R}^{N}_{+} and the monotonicity of solutions follows directly from this principle. In this paper, we consider a more general situation by studying the monotonicity and rigidity results for solutions uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) to (1) with a possible singular nonlinearity ff. In particular, we address the issue that ff is not decreasing in the whole (0,+)(0,+\infty). The main assumption on ff we require is the following:

  1. (FF)

    for any M>0M>0, there exists C(M)>0C(M)>0 such that

    f(s)f(t)C(M)(st) for all 0<tsM.f(s)-f(t)\leq C(M)(s-t)\quad\text{ for all }0<t\leq s\leq M.

Our first result is the following monotonicity result, which holds in a very general setting. Indeed, we require only the behavior of ff near its possible singular point.

Theorem 1.

Assume that f:(0,+)f:(0,+\infty) is a locally Lipschitz continuous function satisfying (FF) and there exist c0,t0>0c_{0},t_{0}>0 such that

f(t)>c0t for all 0<t<t0.f(t)>c_{0}t\quad\text{ for all }0<t<t_{0}.

Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1) with uL(Σλ)u\in L^{\infty}(\Sigma_{\lambda}) for all λ>0\lambda>0. Then

uxN>0 in +N.\frac{\partial u}{\partial x_{N}}>0\quad\text{ in }\mathbb{R}^{N}_{+}.

Theorem 1 applies not only to singular nonlinearities but also the superlinear ones. Since ff is not decreasing, we need to derive a weak comparison principle for the problem in narrow strips and exploit the moving plane method to prove Theorem 1. Similar results for singular problems in bounded domains were obtained in [15, 11]. Theorem 1 can be applied to problem (2) to yield the monotonicity of solutions. Furthermore, the inward derivatives of all such solutions must blow up near the boundary. Indeed, we can provide a precise estimate of the blow-up rate of derivatives as xN0+x_{N}\to 0^{+} in our next result.

Theorem 2.

Assume that γ>1\gamma>1 and g:[0,+)g:[0,+\infty) is a locally Lipschitz continuous function. Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (2) with uL(Σλ¯)u\in L^{\infty}(\Sigma_{\overline{\lambda}}) for some λ¯>0\overline{\lambda}>0. Then

uxN>0 in +N.\frac{\partial u}{\partial x_{N}}>0\quad\text{ in }\mathbb{R}^{N}_{+}.

Moreover, for each β(0,1)\beta\in(0,1), there exist c1,c2,λ0>0c_{1},c_{2},\lambda_{0}>0 such that

c1xN1γγ+1<u(x)η<c2xN1γγ+1 in Σλ0c_{1}x_{N}^{\frac{1-\gamma}{\gamma+1}}<\frac{\partial u(x)}{\partial\eta}<c_{2}x_{N}^{\frac{1-\gamma}{\gamma+1}}\quad\text{ in }\Sigma_{\lambda_{0}} (5)

for all η𝕊+N1\eta\in\mathbb{S}^{N-1}_{+} with (η,eN)β(\eta,e_{N})\geq\beta, where 𝕊+N1:=+NB1(0)\mathbb{S}^{N-1}_{+}:=\mathbb{R}^{N}_{+}\cap\partial B_{1}(0) and eN:=(0,,0,1)e_{N}:=(0,\dots,0,1).

In this paper, we also exploit the techniques and ideas from [25] to establish one-dimensional symmetry of solutions to singular problems whose a model is problem (2), where γ>1\gamma>1 and g:(0,+)g:(0,+\infty)\to\mathbb{R} is a nonnegative locally Lipschitz continuous function such that lim supt+tγg(t)<+\limsup_{t\to+\infty}t^{\gamma}g(t)<+\infty. Notice that gg may have a singularity at zero such as g(t)=1tβg(t)=\frac{1}{t^{\beta}} with βγ\beta\geq\gamma.

Theorem 3.

Assume that γ>1\gamma>1 and f:(0,+)f:(0,+\infty) is a positive locally Lipschitz continuous function satisfying (FF) and

  1. (i)

    there exists c0>0c_{0}>0 such that

    f(t)>c0tγ for all t>0,f(t)>\frac{c_{0}}{t^{\gamma}}\quad\text{ for all }t>0,
  2. (ii)

    there exist c1,t1>0c_{1},t_{1}>0 such that ff is nonincreasing on (t1,+)(t_{1},+\infty) and

    f(t)<c1tγ for all t>t1.f(t)<\frac{c_{1}}{t^{\gamma}}\quad\text{ for all }t>t_{1}.

Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1) with uL(Σλ¯)u\in L^{\infty}(\Sigma_{\overline{\lambda}}) for some λ¯>0\overline{\lambda}>0. Then u(x)v(xN)u(x)\equiv v(x_{N}), where vv is given by the formula

0v(t)dsM+F(s)=2t for all t0\int_{0}^{v(t)}\frac{ds}{\sqrt{M+F(s)}}=\sqrt{2}t\quad\text{ for all }t\geq 0

for some M0M\geq 0, where F(s)=s+f(t)𝑑tF(s)=\int_{s}^{+\infty}f(t)dt.

We will employ the sliding method, which was introduced by Berestycki and Nirenberg [8], to prove Theorem 3. We stress that in Theorem 3 we do not assume that ff is nonincreasing in the whole domain (0,+)(0,+\infty). If this condition is granted, then we can show that solutions depend only on xNx_{N} without the assumption of their boundedness on strips. This in turn yields a classification result. The proof for the following result is similar to the one in [26].

Theorem 4.

Assume that f:(0,+)f:(0,+\infty) is a nonincreasing positive locally Lipschitz continuous function. Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1). Then uu depends only on xNx_{N}. Consequently, such a solution exists if and only if 1+f(t)𝑑t<+\int_{1}^{+\infty}f(t)dt<+\infty. Moreover, when such a solution exists, it is given by u(x)v(xN)u(x)\equiv v(x_{N}), where vv is determined by the formula

0v(t)dsM+F(s)=2t for all t0\int_{0}^{v(t)}\frac{ds}{\sqrt{M+F(s)}}=\sqrt{2}t\quad\text{ for all }t\geq 0

for some M0M\geq 0, where F(s)=s+f(t)𝑑tF(s)=\int_{s}^{+\infty}f(t)dt.

The rest of this paper is devoted to the proofs of our main results. In Section 2 we prove a comparison principle for narrow strips and derive some a priori estimates for solutions. In Section 3, we prove the monotonicity and rigidity of solutions by means of the moving plane and sliding methods.

2. Preliminaries

2.1. Weak comparison principle for narrow strips

We prove a weak comparison principle which can be applied to problems with singular nonlinearities.

Proposition 5.

Assume that f:(0,+)f:(0,+\infty) is a locally Lipschitz continuous function such that (FF) holds and uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) satisfies

{Δuf(u),u>0 in +N,u=0 on +N,uL(Σλ¯) for some λ¯>0.\begin{cases}-\Delta u\leq f(u),~{}u>0&\text{ in }\mathbb{R}^{N}_{+},\\ u=0&\text{ on }\partial\mathbb{R}^{N}_{+},\\ u\in L^{\infty}(\Sigma_{\overline{\lambda}})&\text{ for some }\overline{\lambda}>0.\end{cases}

Then there exists a small positive number λ=λ(f,uL(Σλ¯))<λ¯\lambda^{*}=\lambda^{*}(f,\|u\|_{L^{\infty}(\Sigma_{\overline{\lambda}})})<\overline{\lambda} such that: if 0<λλ0<\lambda\leq\lambda^{*} and vC2(+N)C(+N¯)v\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) satisfies

{Δvf(v),v>0 in Σλ,v>0 on +N,uv on {xN=λ},\begin{cases}-\Delta v\geq f(v),~{}v>0&\text{ in }\Sigma_{\lambda},\\ v>0&\text{ on }\partial\mathbb{R}^{N}_{+},\\ u\leq v&\text{ on }\{x_{N}=\lambda\},\\ \end{cases}

then uvu\leq v in Σλ\Sigma_{\lambda}.

Proof.

In what follows, we consider λ<λ¯\lambda<\overline{\lambda}. For R>0R>0, let φRC(N1)\varphi_{R}\in C^{\infty}(\mathbb{R}^{N-1}) be such that

{0φR1 in N1,φR=1 in BR,φR=0 in N1B2R,|φR|2R in B2RBR,\begin{cases}0\leq\varphi_{R}\leq 1&\text{ in }\mathbb{R}^{N-1},\\ \varphi_{R}=1&\text{ in }B^{\prime}_{R},\\ \varphi_{R}=0&\text{ in }\mathbb{R}^{N-1}\setminus B^{\prime}_{2R},\\ |\nabla\varphi_{R}|\leq\frac{2}{R}&\text{ in }B^{\prime}_{2R}\setminus B^{\prime}_{R},\end{cases} (6)

where BrB_{r}^{\prime} denotes the ball in N1\mathbb{R}^{N-1} of center 0N10^{\prime}\in\mathbb{R}^{N-1} with radius rr. We set

φ(x)=w+(x)φR2(x)χΣλ(x),\varphi(x)=w^{+}(x)\varphi_{R}^{2}(x^{\prime})\chi_{\Sigma_{\lambda}}(x),

where w+:=max{uv,0}w^{+}:=\max\{u-v,0\}. Since the support of φ\varphi is compactly contained in Σλ{xN=λ}\Sigma_{\lambda}\cup\{x_{N}=\lambda\}, we can use it as a test function in Δuf(u)-\Delta u\leq f(u) and Δvf(v)-\Delta v\geq f(v). Then subtracting, we obtain

Σλ|w+|2φR2\displaystyle\int_{\Sigma_{\lambda}}|\nabla w^{+}|^{2}\varphi_{R}^{2} 2Σλ(w+,φR)w+φR+Σλ(f(u)f(v))w+φR2\displaystyle\leq-2\int_{\Sigma_{\lambda}}(\nabla w^{+},\nabla\varphi_{R})w^{+}\varphi_{R}+\int_{\Sigma_{\lambda}}(f(u)-f(v))w^{+}\varphi_{R}^{2}
2Σλ|w+||φR|w+φR+Σλ(f(u)f(v))w+φR2.\displaystyle\leq 2\int_{\Sigma_{\lambda}}|\nabla w^{+}||\nabla\varphi_{R}|w^{+}\varphi_{R}+\int_{\Sigma_{\lambda}}(f(u)-f(v))w^{+}\varphi_{R}^{2}.

In the set Σλ{w+>0}\Sigma_{\lambda}\cap\{w^{+}>0\} we have

0<v<uuL(Σλ¯).0<v<u\leq\|u\|_{L^{\infty}(\Sigma_{\overline{\lambda}})}.

Hence by exploiting Young’s inequality and using (FF), we have

Σλ|w+|2φR2\displaystyle\int_{\Sigma_{\lambda}}|\nabla w^{+}|^{2}\varphi_{R}^{2} 12Σλ|w+|2φR2+2Σλ|φR|2(w+)2\displaystyle\leq\frac{1}{2}\int_{\Sigma_{\lambda}}|\nabla w^{+}|^{2}\varphi_{R}^{2}+2\int_{\Sigma_{\lambda}}|\nabla\varphi_{R}|^{2}(w^{+})^{2}
+C(uL(Σλ¯))Σλ(w+)2φR2.\displaystyle\qquad+C(\|u\|_{L^{\infty}(\Sigma_{\overline{\lambda}})})\int_{\Sigma_{\lambda}}(w^{+})^{2}\varphi_{R}^{2}.

This is equivalent to

Σλ|w+|2φR24Σλ|φR|2(w+)2+2C(uL(Σλ¯))Σλ(w+)2φR2.\int_{\Sigma_{\lambda}}|\nabla w^{+}|^{2}\varphi_{R}^{2}\leq 4\int_{\Sigma_{\lambda}}|\nabla\varphi_{R}|^{2}(w^{+})^{2}+2C(\|u\|_{L^{\infty}(\Sigma_{\overline{\lambda}})})\int_{\Sigma_{\lambda}}(w^{+})^{2}\varphi_{R}^{2}. (7)

By the classical Poincaré inequality in the interval (0,λ)(0,\lambda), we have

Σλ|w+|2φR2\displaystyle\int_{\Sigma_{\lambda}}|\nabla w^{+}|^{2}\varphi_{R}^{2} N1(0λ(w+xN)2𝑑xN)φR2(x)𝑑x\displaystyle\geq\int_{\mathbb{R}^{N-1}}\left(\int_{0}^{\lambda}\left(\frac{\partial w^{+}}{\partial x_{N}}\right)^{2}dx_{N}\right)\varphi_{R}^{2}(x^{\prime})dx^{\prime}
π2λ2N1(0λ(w+)2𝑑xN)φR2(x)𝑑x\displaystyle\geq\frac{\pi^{2}}{\lambda^{2}}\int_{\mathbb{R}^{N-1}}\left(\int_{0}^{\lambda}(w^{+})^{2}dx_{N}\right)\varphi_{R}^{2}(x^{\prime})dx^{\prime}
=π2λ2Σλ(w+)2φR2.\displaystyle=\frac{\pi^{2}}{\lambda^{2}}\int_{\Sigma_{\lambda}}(w^{+})^{2}\varphi_{R}^{2}.

Therefore, (7) leads to

(π2λ22C(uL(Σλ¯)))Σλ(w+)2φR24Σλ|φR|2(w+)2.\left(\frac{\pi^{2}}{\lambda^{2}}-2C(\|u\|_{L^{\infty}(\Sigma_{\overline{\lambda}})})\right)\int_{\Sigma_{\lambda}}(w^{+})^{2}\varphi_{R}^{2}\leq 4\int_{\Sigma_{\lambda}}|\nabla\varphi_{R}|^{2}(w^{+})^{2}.

Choosing λ0=π[4+2C(uL(Σλ¯))]12\lambda_{0}=\pi[4+2C(\|u\|_{L^{\infty}(\Sigma_{\overline{\lambda}})})]^{-\frac{1}{2}}, then for all λ<min{λ0,λ¯}\lambda<\min\{\lambda_{0},\overline{\lambda}\} we have

Σλ(w+)2φR2Σλ|φR|2(w+)2.\int_{\Sigma_{\lambda}}(w^{+})^{2}\varphi_{R}^{2}\leq\int_{\Sigma_{\lambda}}|\nabla\varphi_{R}|^{2}(w^{+})^{2}.

Using (6), we deduce

ΣλR(w+)24R2Σλ2R(w+)2,\int_{\Sigma_{\lambda}^{R}}(w^{+})^{2}\leq\frac{4}{R^{2}}\int_{\Sigma_{\lambda}^{2R}}(w^{+})^{2},

where ΣλR:=BR×(0,λ)\Sigma_{\lambda}^{R}:=B^{\prime}_{R}\times(0,\lambda). Setting h(R):=ΣλR(w+)2h(R):=\int_{\Sigma_{\lambda}^{R}}(w^{+})^{2}, we have

h(R)4R2h(2R) and h(R)CλRN1 for all R>0.h(R)\leq\frac{4}{R^{2}}h(2R)\quad\text{ and }\quad h(R)\leq C_{\lambda}R^{N-1}\quad\text{ for all }R>0.

Hence h(R)12Nh(2R)h(R)\leq\frac{1}{2^{N}}h(2R) for all R>2N+22R>2^{\frac{N+2}{2}}. By iteration of this inequality, we obtain

h(R)12Nkh(2kR)Cλ2Nk(2kR)N1=Cλ2kRN1h(R)\leq\frac{1}{2^{Nk}}h(2^{k}R)\leq\frac{C_{\lambda}}{2^{Nk}}(2^{k}R)^{N-1}=\frac{C_{\lambda}}{2^{k}}R^{N-1}

for all kk\in\mathbb{N} and R>2N+22R>2^{\frac{N+2}{2}}. Letting kk\to\infty, we deduce h(R)0h(R)\equiv 0. This implies Σλ(w+)2=0\int_{\Sigma_{\lambda}}(w^{+})^{2}=0, which means uvu\leq v in Σλ\Sigma_{\lambda} for all λ<λ:=min{λ0,λ¯}\lambda<\lambda^{*}:=\min\{\lambda_{0},\overline{\lambda}\}. ∎

2.2. A priori estimates

We need the following property on solutions so that we can carry out the moving plane method to prove Theorem 1.

Lemma 6.

Assume that f:(0,+)f:(0,+\infty) is a locally Lipschitz continuous function such that (FF) holds. Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1) with uL(Σλ¯)u\in L^{\infty}(\Sigma_{\overline{\lambda}}) for some λ¯>0\overline{\lambda}>0. Then

limxN0+u(x,xN)=0 uniformly in x+N1.\lim_{x_{N}\to 0^{+}}u(x^{\prime},x_{N})=0\text{ uniformly in }x^{\prime}\in\mathbb{R}^{N-1}_{+}.
Proof.

Let λ<λ¯\lambda^{*}<\overline{\lambda} be defined as in Proposition 5 and choose some ρ>uL(Σλ)\rho>\|u\|_{L^{\infty}(\Sigma_{\lambda^{*}})}. Let h:(0,+)h:(0,+\infty)\to\mathbb{R} be a C1C^{1} function such that

h(t)>max{f(t),0}\displaystyle h(t)>\max\{f(t),0\}  in (0,ρ),\displaystyle\quad\text{ in }(0,\rho),
h(t)=ct2\displaystyle h(t)=\frac{c}{t^{2}}  in [ρ,+)\displaystyle\quad\text{ in }[\rho,+\infty)

for some c>0c>0. We set H(t)=ρth(s)𝑑sH(t)=\int_{\rho}^{t}h(s)ds for t>0t>0, then HH is strictly increasing in (0,+)(0,+\infty) and H(t)<ρ+h(s)𝑑s=cρH(t)<\int_{\rho}^{+\infty}h(s)ds=\frac{c}{\rho}. For each μcρ\mu\geq\frac{c}{\rho}, one can check that

0+dsμH(s)>ρ+dsμ=+\int_{0}^{+\infty}\frac{ds}{\sqrt{\mu-H(s)}}>\int_{\rho}^{+\infty}\frac{ds}{\sqrt{\mu}}=+\infty

and

0tdsμH(s)<tμH(t)<+ for 0<t<+.\int_{0}^{t}\frac{ds}{\sqrt{\mu-H(s)}}<\frac{t}{\sqrt{\mu-H(t)}}<+\infty\quad\text{ for }0<t<+\infty.

Hence the formula

0vμ(t)dsμH(s)=2t for all t0\int_{0}^{v_{\mu}(t)}\frac{ds}{\sqrt{\mu-H(s)}}=\sqrt{2}t\quad\text{ for all }t\geq 0

uniquely determine a function vμC2(+)C(+¯)v_{\mu}\in C^{2}(\mathbb{R}_{+})\cap C(\overline{\mathbb{R}_{+}}), which is a solution to the ODE problem

{v′′=h(v) in +,v(t)>0 in +,v(0)=0.\begin{cases}-v^{\prime\prime}=h(v)&\text{ in }\mathbb{R}_{+},\\ v^{\prime}(t)>0&\text{ in }\mathbb{R}_{+},\\ v(0)=0.\end{cases}

Moreover, limμ+vμ(t)=+\lim_{\mu\to+\infty}v_{\mu}(t)=+\infty for all t>0t>0.

We fix some μ>0\mu>0 such that vμ(λ)>ρv_{\mu}(\lambda^{*})>\rho. Then we choose λ0<λ\lambda_{0}<\lambda^{*} satisfying uL(Σλ)<vμ(λ0)<ρ\|u\|_{L^{\infty}(\Sigma_{\lambda^{*}})}<v_{\mu}(\lambda_{0})<\rho. By abuse of notation, we will write vμ(x,xN):=vμ(xN)v_{\mu}(x^{\prime},x_{N}):=v_{\mu}(x_{N}). Then 0<vμ<ρ0<v_{\mu}<\rho in Σλ0\Sigma_{\lambda_{0}} and u<vμu<v_{\mu} on {xN=λ0}\{x_{N}=\lambda_{0}\}.

For small ε>0\varepsilon>0 such that vμ(λ0+ε)<ρv_{\mu}(\lambda_{0}+\varepsilon)<\rho, we define

vμ,ε(x):=vμ(x+εeN),v_{\mu,\varepsilon}(x):=v_{\mu}(x+\varepsilon e_{N}),

Then

{vμ(ε)<vμ,ε<vμ(λ0+ε)<ρ in Σλ0,Δvμ,ε=h(vμ,ε)>f(vμ,ε) in Σλ0,uvμ,ε on Σλ0.\begin{cases}v_{\mu}(\varepsilon)<v_{\mu,\varepsilon}<v_{\mu}(\lambda_{0}+\varepsilon)<\rho&\text{ in }\Sigma_{\lambda_{0}},\\ -\Delta v_{\mu,\varepsilon}=h(v_{\mu,\varepsilon})>f(v_{\mu,\varepsilon})&\text{ in }\Sigma_{\lambda_{0}},\\ u\leq v_{\mu,\varepsilon}&\text{ on }\partial\Sigma_{\lambda_{0}}.\end{cases}

Now Proposition 5 implies uvμ,εu\leq v_{\mu,\varepsilon} in Σλ0\Sigma_{\lambda_{0}}. Letting ε0\varepsilon\to 0, we have uvμu\leq v_{\mu} in Σλ0\Sigma_{\lambda_{0}} and the conclusion follows from that fact that limt0+vμ(t)=0\lim_{t\to 0^{+}}v_{\mu}(t)=0. ∎

We prove some a priori estimates for solutions to (1) in what follows. The next lemma improves the upper bound on uu near the boundary in Lemma 6 when an explicit upper bound on ff is given.

Lemma 7.

Assume that f:(0,+)f:(0,+\infty) is a locally Lipschitz continuous function such that (FF) holds and f(t)<c0tγf(t)<\frac{c_{0}}{t^{\gamma}} for all 0<t<t00<t<t_{0}, where c0,t0>0c_{0},t_{0}>0 and γ>1\gamma>1. Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1) with uL(Σλ¯)u\in L^{\infty}(\Sigma_{\overline{\lambda}}) for some λ¯>0\overline{\lambda}>0. Then

u(x)CxN2γ+1 in Σλ0u(x)\leq Cx_{N}^{\frac{2}{\gamma+1}}\quad\text{ in }\Sigma_{\lambda_{0}}

for some constants C,λ0>0C,\lambda_{0}>0.

Proof.

Let λ<λ¯\lambda^{*}<\overline{\lambda} be defined as in Proposition 5 and choose some ρ>uL(Σλ)\rho>\|u\|_{L^{\infty}(\Sigma_{\lambda^{*}})}. By compactness, there exists cc0c_{\geq}c_{0} such that f(t)<c1tγf(t)<\frac{c_{1}}{t^{\gamma}} for all 0<t<ρ0<t<\rho. Let

v(t)(γ+1)2γ+1(2γ2)1γ+1t2γ+1,v(t)\equiv\frac{(\gamma+1)^{\frac{2}{\gamma+1}}}{(2\gamma-2)^{\frac{1}{\gamma+1}}}t^{\frac{2}{\gamma+1}},

then vμ(x)=μv(xN)v_{\mu}(x)=\mu v(x_{N}) solves Δvμ=μγ+1vμγ-\Delta v_{\mu}=\frac{\mu^{\gamma+1}}{v_{\mu}^{\gamma}} in +N\mathbb{R}^{N}_{+}. By abuse of notation, we will write vμ(xN):=vμ(x,xN)v_{\mu}(x_{N}):=v_{\mu}(x^{\prime},x_{N}). We choose μ\mu large such that μγ+1>c1\mu^{\gamma+1}>c_{1} and vμ(λ)>ρv_{\mu}(\lambda^{*})>\rho. Then we choose λ0<λ\lambda_{0}<\lambda^{*} satisfying uL(Σλ)<vμ(λ0)<ρ\|u\|_{L^{\infty}(\Sigma_{\lambda^{*}})}<v_{\mu}(\lambda_{0})<\rho.  Now we have 0<vμ<ρ0<v_{\mu}<\rho in Σλ0\Sigma_{\lambda_{0}} and Δvμf(vμ)-\Delta v_{\mu}\geq f(v_{\mu}) in Σλ0\Sigma_{\lambda_{0}}.

For small ε>0\varepsilon>0 such that vμ(λ0+ε)<ρv_{\mu}(\lambda_{0}+\varepsilon)<\rho, we define

vμ,ε(x)=vμ(x+εeN).v_{\mu,\varepsilon}(x)=v_{\mu}(x+\varepsilon e_{N}).

Then

{Δvμ,εf(vμ,ε) in Σλ0,vμ,ε=vμ(ε)>0 on {xN=0},u<vμ,ε on {xN=λ0}.\begin{cases}-\Delta v_{\mu,\varepsilon}\geq f(v_{\mu,\varepsilon})&\text{ in }\Sigma_{\lambda_{0}},\\ v_{\mu,\varepsilon}=v_{\mu}(\varepsilon)>0&\text{ on }\{x_{N}=0\},\\ u<v_{\mu,\varepsilon}&\text{ on }\{x_{N}=\lambda_{0}\}.\end{cases}

Now Proposition 5 implies uvμ,εu\leq v_{\mu,\varepsilon} in Σλ0\Sigma_{\lambda_{0}}. Letting ε0\varepsilon\to 0 we conclude the proof. ∎

The next lemma concerns a lower bound on solutions.

Lemma 8.

Assume that f:(0,+)f:(0,+\infty) is a locally Lipschitz continuous function and f(t)>c0tγf(t)>\frac{c_{0}}{t^{\gamma}} for all 0<t<t00<t<t_{0}, where c0,t0>0c_{0},t_{0}>0 and γ0\gamma\geq 0. Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1). Then

u(x)min{CxN2γ+1,t0} in +Nu(x)\geq\min\{Cx_{N}^{\frac{2}{\gamma+1}},t_{0}\}\quad\text{ in }\mathbb{R}^{N}_{+}

for some constant C>0C>0 independent of t0t_{0}.

Proof.

Let λ1>0\lambda_{1}>0 and ϕ1C2(B1(0)¯)\phi_{1}\in C^{2}(\overline{B_{1}(0)}) be the first eigenvalue and a corresponding positive eigenfunction of the Laplacian in B1(0)B_{1}(0), namely,

{Δϕ1=λ1ϕ1 in B1(0),ϕ1>0 in B1(0),ϕ1=0 on B1(0).\begin{cases}-\Delta\phi_{1}=\lambda_{1}\phi_{1}&\text{ in }B_{1}(0),\\ \phi_{1}>0&\text{ in }B_{1}(0),\\ \phi_{1}=0&\text{ on }\partial B_{1}(0).\end{cases}

Setting

w=βϕ12γ+1,w=\beta\phi_{1}^{\frac{2}{\gamma+1}},

where β>0\beta>0 will be chosen later. Direct calculation yields that

Δw=α(x)wγ in B1(0),-\Delta w=\frac{\alpha(x)}{w^{\gamma}}\quad\text{ in }B_{1}(0),

where

α(x):=2βγ+1γ+1(γ1γ+1|ϕ1(x)|2+λ1ϕ12(x)).\alpha(x):=\frac{2\beta^{\gamma+1}}{\gamma+1}\left(\frac{\gamma-1}{\gamma+1}|\nabla\phi_{1}(x)|^{2}+\lambda_{1}\phi_{1}^{2}(x)\right).

Now we fix β>0\beta>0 such that supxB1(0)α(x)=c0\sup_{x\in B_{1}(0)}\alpha(x)=c_{0} and hence

Δwc0wγ in B1(0).-\Delta w\leq\frac{c_{0}}{w^{\gamma}}\text{ in }B_{1}(0).

Let R0>0R_{0}>0 be such that R02γ+1w(0)=t0R_{0}^{\frac{2}{\gamma+1}}w(0)=t_{0}. For any 0<RR00<R\leq R_{0} and x0=(x0,x0,N)Nx_{0}=(x_{0}^{\prime},x_{0,N})\in\mathbb{R}^{N} with x0,NR+εx_{0,N}\geq R+\varepsilon, where ε\varepsilon is sufficiently small, we set

wx0,R(x):=R2γ+1w(xx0R) in BR(x0).w_{x_{0},R}(x):=R^{\frac{2}{\gamma+1}}w\left(\frac{x-x_{0}}{R}\right)\quad\text{ in }B_{R}(x_{0}).

Then

wx0,Rt0 and Δwx0,Rc0wx0,Rγ in BR(x0).w_{x_{0},R}\leq t_{0}\quad\text{ and }\quad-\Delta w_{x_{0},R}\leq\frac{c_{0}}{w_{x_{0},R}^{\gamma}}\quad\text{ in }B_{R}(x_{0}).

On the other hand, since wx0,R=0<uw_{x_{0},R}=0<u on BR(x0)\partial B_{R}(x_{0}), we can use (wx0,Ru)+χBR(x0)(w_{x_{0},R}-u)^{+}\chi_{B_{R}(x_{0})} as a test function in

Δu=f(u) and Δwx0,Rc0wx0,Rγ-\Delta u=f(u)\quad\text{ and }\quad-\Delta w_{x_{0},R}\leq\frac{c_{0}}{w_{x_{0},R}^{\gamma}}

to obtain

BR(x0)|(wx0,Ru)+|2BR(x0)(c0wx0,Rγf(u))(wx0,Ru)+.\int_{B_{R}(x_{0})}\left|\nabla(w_{x_{0},R}-u)^{+}\right|^{2}\leq\int_{B_{R}(x_{0})}\left(\frac{c_{0}}{w_{x_{0},R}^{\gamma}}-f(u)\right)(w_{x_{0},R}-u)^{+}.

In BR(x0){wx0,R>u}B_{R}(x_{0})\cap\{w_{x_{0},R}>u\} we have f(u)c0uγf(u)\geq\frac{c_{0}}{u^{\gamma}}. Hence

BR(x0)|(wx0,Ru)+|2BR(x0)(c0wx0,Rγc0uγ)(wx0,Ru)+ 0.\int_{B_{R}(x_{0})}\left|\nabla(w_{x_{0},R}-u)^{+}\right|^{2}\leq\int_{B_{R}(x_{0})}\left(\frac{c_{0}}{w_{x_{0},R}^{\gamma}}-\frac{c_{0}}{u^{\gamma}}\right)(w_{x_{0},R}-u)^{+} \leq 0.

This implies uwx0,Ru\geq w_{x_{0},R} in BR(x0)B_{R}(x_{0}) with x0,NR+εx_{0,N}\geq R+\varepsilon. Since ε>0\varepsilon>0 is arbitrary, we deduce

uwx0,R in BR(x0) for all 0<RR0 and all x0N with x0,NR.u\geq w_{x_{0},R}\text{ in }B_{R}(x_{0})\text{ for all }0<R\leq R_{0}\text{ and all }x_{0}\in\mathbb{R}^{N}\text{ with }x_{0,N}\geq R.

In particular, if x0,N=R<R0x_{0,N}=R<R_{0}, then

u(x0)wx0,R(x0)=w(0)R2γ+1=w(0)x0,N2γ+1.u(x_{0})\geq w_{x_{0},R}(x_{0})=w(0)R^{\frac{2}{\gamma+1}}=w(0)x_{0,N}^{\frac{2}{\gamma+1}}.

If x0,NR=R0x_{0,N}\geq R=R_{0}, then

u(x0)wx0,R(x0)=w(0)R02γ+1=t0.u(x_{0})\geq w_{x_{0},R}(x_{0})=w(0)R_{0}^{\frac{2}{\gamma+1}}=t_{0}.

The conclusion follows from the fact that x0x_{0} is chosen arbitrarily in N\mathbb{R}^{N}. ∎

Under a weaker assumption on ff, we can still obtain a lower bound of uu, which is useful in many situations.

Lemma 9.

Assume that f:(0,+)f:(0,+\infty) is a locally Lipschitz continuous function and f(t)>c0tf(t)>c_{0}t for all 0<t<t00<t<t_{0}, where c0,t0>0c_{0},t_{0}>0. Let uC2(+N)C(+N¯)u\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}) be a solution to (1). Then

u(x)min{CxN,t0} in +N.u(x)\geq\min\{Cx_{N},t_{0}\}\quad\text{ in }\mathbb{R}^{N}_{+}.

for some constant C>0C>0.

An similar result was obtained by Berestycki et al. [7] for C2(+N¯)C^{2}(\overline{\mathbb{R}^{N}_{+}}) solutions and ff being locally Lipschitz continuous on [0,+)[0,+\infty). We provide a proof that works in our more general situation.

Proof of Lemma 9.

Let λ1>0\lambda_{1}>0 and ϕ1C2(B1(0)¯)\phi_{1}\in C^{2}(\overline{B_{1}(0)}) be the first eigenvalue and the corresponding positive eigenfunction of the Laplacian in B1(0)B_{1}(0) such that ϕ1(0)=t0\phi_{1}(0)=t_{0}. We take R=λ1c0R=\sqrt{\frac{\lambda_{1}}{c_{0}}} and set ϕR(x)=ϕ1(xR)\phi_{R}(x)=\phi_{1}\left(\frac{x}{R}\right), then

{ΔϕR=c0ϕRf(ϕR) in BR(0),ϕR>0 in BR(0),ϕR=0 on BR(0).\begin{cases}-\Delta\phi_{R}=c_{0}\phi_{R}\leq f(\phi_{R})&\text{ in }B_{R}(0),\\ \phi_{R}>0&\text{ in }B_{R}(0),\\ \phi_{R}=0&\text{ on }\partial B_{R}(0).\end{cases}

Since ϕR\phi_{R} is radially symmetric and by abuse of notion, we may write ϕR(x)=ϕR(|x|)\phi_{R}(x)=\phi_{R}(|x|). For each x0+NΣRx_{0}\in\mathbb{R}^{N}_{+}\setminus\Sigma_{R} we set

ϕRx0(x)=ϕR(xx0) for xBR(x0).\phi_{R}^{x_{0}}(x)=\phi_{R}(x-x_{0})\quad\text{ for }x\in B_{R}(x_{0}).

We will show that

uϕRx0 in BR(x0) for every x0+NΣR.u\geq\phi_{R}^{x_{0}}\text{ in }B_{R}(x_{0})\quad\text{ for every }x_{0}\in\mathbb{R}^{N}_{+}\setminus\Sigma_{R}. (8)

To this end, we let any x0:=(x0,x0,N)+NΣRx_{0}:=(x_{0}^{\prime},x_{0,N})\in\mathbb{R}^{N}_{+}\setminus\Sigma_{R}.

We only consider the case x0,N>Rx_{0,N}>R since the case x0,N=Rx_{0,N}=R can be derived by continuity. We define the set

Λ:={s(0,1]u>tϕRx0 in BR(x0) for all t(0,s)}.\Lambda:=\{s\in(0,1]\mid u>t\phi_{R}^{x_{0}}\text{ in }B_{R}(x_{0})\text{ for all }t\in(0,s)\}.

Since uu is positive on compact set BR(x0)¯\overline{B_{R}(x_{0})}, we have Λ\Lambda\neq\emptyset. We denote s0=supΛs_{0}=\sup\Lambda. To derive (8), we have to show that s0=1s_{0}=1. Assume by contradiction that s0<1s_{0}<1. We set ϕ~:=s0ϕRx0\tilde{\phi}:=s_{0}\phi_{R}^{x_{0}}. Then

uϕ~ in BR(x0)u\geq\tilde{\phi}\quad\text{ in }B_{R}(x_{0})
u(x^)=ϕ~(x^) for some x^BR(x0).u(\hat{x})=\tilde{\phi}(\hat{x})\quad\text{ for some }\hat{x}\in B_{R}(x_{0}). (9)

From the boundary data of ϕ~\tilde{\phi}, we can choose ε>0\varepsilon>0 small such that

x^BRε(x0) and u>ϕ~ on BRε(x0).\hat{x}\in B_{R-\varepsilon}(x_{0})\quad\text{ and }\quad u>\tilde{\phi}\quad\text{ on }\partial B_{R-\varepsilon}(x_{0}). (10)

Since uu and ϕ~\tilde{\phi} are positive in the set BRε(x0)¯\overline{B_{R-\varepsilon}(x_{0})} and ff is locally Lipschitz continuous in (0,+)(0,+\infty), the strong comparison principle can be applied in BRε(x0)B_{R-\varepsilon}(x_{0}) to yield either ϕ~<u\tilde{\phi}<u in BRε(x0)B_{R-\varepsilon}(x_{0}) or ϕ~=u\tilde{\phi}=u in BRε(x0)B_{R-\varepsilon}(x_{0}). However, the former contradicts (9), while the latter contradicts (10). Hence (8) holds. This implies

u(x){ϕR(RxN) if xN<R,ϕR(0) if xNR.u(x)\geq\begin{cases}\phi_{R}(R-x_{N})&\text{ if }x_{N}<R,\\ \phi_{R}(0)&\text{ if }x_{N}\geq R.\end{cases}

The conclusion follows immediately from the fact that ϕR(R)<0\phi_{R}^{\prime}(R)<0 and ϕR(0)=t0\phi_{R}(0)=t_{0}. ∎

3. Qualitative properties of solutions

3.1. Monotonicity of solutions

As in the previous works, the main tool we use in proving the monotonicity of solutions is the method of moving planes, which was introduced by Alexandrov [1] in the context of differential geometry and by Serrin [29] in the PDE framework, for an overdetermined problem. We recall some familiar notions related to this method. For each λ>0\lambda>0, we denote

xλ=(x1,x2,,2λxn),x_{\lambda}=(x_{1},x_{2},\dots,2\lambda-x_{n}),

which is the reflection of xx through the hyperplane Σλ\partial\Sigma_{\lambda}. Let uu be a solution to (1). We set

uλ(x)=u(xλ),u_{\lambda}(x)=u(x_{\lambda}),

then uλu_{\lambda} satisfies Δuλ=f(uλ)-\Delta u_{\lambda}=f(u_{\lambda}) in Σ2λ\Sigma_{2\lambda}.

We are ready to prove the main results in this section.

Proof of Theorem 1.

Applying Proposition 5 with vuλv\equiv u_{\lambda}, we find λ>0\lambda^{*}>0 such that uuλu\leq u_{\lambda} in Σλ\Sigma_{\lambda} for all λλ\lambda\leq\lambda^{*}. Hence the set

Λ={λ(0,+)uuμ in Σμ for all μ(0,λ]}\Lambda=\{\lambda\in(0,+\infty)\mid u\leq u_{\mu}\text{ in }\Sigma_{\mu}\text{ for all }\mu\in(0,\lambda]\}

is not empty. Therefore, we can define

λ0=supΛ.\lambda_{0}=\sup\Lambda.

We show that λ0=+\lambda_{0}=+\infty.

By contradiction, we assume that λ0<+\lambda_{0}<+\infty. By continuity, we know that

uuλ0 in Σλ0.u\leq u_{\lambda_{0}}\quad\text{ in }\Sigma_{\lambda_{0}}. (11)

By Lemmas 6 and 9, there exist λ~,δ~>0\tilde{\lambda},\tilde{\delta}>0 sufficiently small such that

u+δ~<uλ in Σλ~ for all λλ0.u+\tilde{\delta}<u_{\lambda}\text{ in }\Sigma_{\tilde{\lambda}}\quad\text{ for all }\lambda\geq\lambda_{0}. (12)

We will reach a contradiction by showing that for some small ε0>0\varepsilon_{0}>0,

uuλ in Σλ for all λ[λ0,λ0+ε0].u\leq u_{\lambda}\text{ in }\Sigma_{\lambda}\quad\text{ for all }\lambda\in[\lambda_{0},\lambda_{0}+\varepsilon_{0}].

If this is not true, then there exist λnλ0\lambda_{n}\searrow\lambda_{0} and xn:=(xn,xn,N)ΣλnΣλ~x_{n}:=(x_{n}^{\prime},x_{n,N})\in\Sigma_{\lambda_{n}}\setminus\Sigma_{\tilde{\lambda}} such that

u(xn)>uλn(xn).u(x_{n})>u_{\lambda_{n}}(x_{n}). (13)

Up to a subsequence, we may assume that xn,Ny0[λ~,λ0]x_{n,N}\to y_{0}\in[\tilde{\lambda},\lambda_{0}] as nn\to\infty. Now we set

un(x,xN):=u(x+xn,xN).u_{n}(x^{\prime},x_{N}):=u(x^{\prime}+x_{n}^{\prime},x_{N}).

By Lemma 9, we know that

min{Cλ~,t0}unuL(Σλ) in ΣλΣλ~ for λ>0.\min\{C\tilde{\lambda},t_{0}\}\leq u_{n}\leq\|u\|_{L^{\infty}(\Sigma_{\lambda})}\quad\text{ in }\Sigma_{\lambda}\setminus\Sigma_{\tilde{\lambda}}\text{ for }\lambda>0.

Hence f(un)f(u_{n}) are also bounded on each strip ΣλΣλ~\Sigma_{\lambda}\setminus\Sigma_{\tilde{\lambda}}. By standard regularity, Ascoli-Arzelà’s theorem and a diagonal process, we deduce that

unv in Cloc2(+NΣλ~)u_{n}\to v\quad\text{ in }C^{2}_{\rm loc}(\mathbb{R}^{N}_{+}\setminus\Sigma_{\tilde{\lambda}})

up to a subsequence, where vv weakly solves Δu=f(u)-\Delta u=f(u) in +NΣλ~¯\mathbb{R}^{N}_{+}\setminus\overline{\Sigma_{\tilde{\lambda}}}. Moreover, (11), (13) imply vvλ0v\leq v_{\lambda_{0}} in Σλ0Σλ~\Sigma_{\lambda_{0}}\setminus\Sigma_{\tilde{\lambda}} and v(0,y0)vλ0(0,y0)v(0^{\prime},y_{0})\geq v_{\lambda_{0}}(0^{\prime},y_{0}). Hence

v(0,y0)=vλ0(0,y0).v(0^{\prime},y_{0})=v_{\lambda_{0}}(0^{\prime},y_{0}).

Therefore,

Δ(vλ0v)+C(vλ0v)=f(vλ0)f(v)+C(vλ0v)0-\Delta(v_{\lambda_{0}}-v)+C(v_{\lambda_{0}}-v)=f(v_{\lambda_{0}})-f(v)+C(v_{\lambda_{0}}-v)\geq 0 (14)

in any compact set of {λ~<xNλ0}\{\tilde{\lambda}<x_{N}\leq\lambda_{0}\} with sufficiently large CC. By the strong maximum principle, we deduce v<vλ0v<v_{\lambda_{0}} in Σλ0Σλ~¯\Sigma_{\lambda_{0}}\setminus\overline{\Sigma_{\tilde{\lambda}}}. (The case vvλ0v\equiv v_{\lambda_{0}} in Σλ0\Sigma_{\lambda_{0}} cannot happen due to v<vλ0v<v_{\lambda_{0}} on {xN=λ~}\{x_{N}=\tilde{\lambda}\} deduced from (12).) This implies y0=λ0y_{0}=\lambda_{0}. By the mean value theorem, there exists ξn(xn,N,2λnxn,N)\xi_{n}\in(x_{n,N},2\lambda_{n}-x_{n,N}) such that

unxN(0,ξn)=un(0,2λnxn,N)un(0,xn,N)2λn2xn,N=uλn(xn)u(xn)2λn2xn,N<0.\frac{\partial u_{n}}{\partial x_{N}}(0^{\prime},\xi_{n})=\frac{u_{n}(0^{\prime},2\lambda_{n}-x_{n,N})-u_{n}(0^{\prime},x_{n,N})}{2\lambda_{n}-2x_{n,N}}=\frac{u_{\lambda_{n}}(x_{n})-u(x_{n})}{2\lambda_{n}-2x_{n,N}}<0.

Letting nn\to\infty, we obtain

vxN(0,λ0)0.\frac{\partial v}{\partial x_{N}}(0^{\prime},\lambda_{0})\leq 0.

Hence

(vλ0v)xN(0,λ0)=2vxN(0,λ0)0.\frac{\partial(v_{\lambda_{0}}-v)}{\partial x_{N}}(0^{\prime},\lambda_{0})=-2\frac{\partial v}{\partial x_{N}}(0^{\prime},\lambda_{0})\geq 0.

However, this contradicts the Höpf lemma [20] for (14) in Σλ0Σλ~\Sigma_{\lambda_{0}}\setminus\Sigma_{\tilde{\lambda}}.

Therefore, λ0=+\lambda_{0}=+\infty. Hence uuλu\leq u_{\lambda} in Σλ\Sigma_{\lambda} for all λ>0\lambda>0. Exploiting the strong maximum principle and the Höpf lemma for uλuu_{\lambda}-u as above we deduce

uxN>0 in N,\frac{\partial u}{\partial x_{N}}>0\quad\text{ in }\mathbb{R}^{N},

which is what we have to prove. ∎

Proof of Theorem 2.

Since g:[0,+)g:[0,+\infty) is a locally Lipschitz continuous, there exist t0,c1,c2>0t_{0},c_{1},c_{2}>0 such that the function f(t)=1tγ+g(u)f(t)=\frac{1}{t^{\gamma}}+g(u) is decreasing on (0,t0)(0,t_{0}) and

c1tγ<f(t)<c2tγ in (0,t0).\frac{c_{1}}{t^{\gamma}}<f(t)<\frac{c_{2}}{t^{\gamma}}\quad\text{ in }(0,t_{0}).

Hence Lemmas 7 and 8 imply the existence of λ0>0\lambda_{0}>0 such that

cxN2γ+1u(x)CxN2γ+1 in Σλ0.cx_{N}^{\frac{2}{\gamma+1}}\leq u(x)\leq Cx_{N}^{\frac{2}{\gamma+1}}\quad\text{ in }\Sigma_{\lambda_{0}}. (15)

The monotonicity of uu follows from Theorem 1. So we only prove (5). Our proof is motivated by an idea from [28].

Let any A>a>0A>a>0 and a positive sequence (εn)(\varepsilon_{n}) such that εn0\varepsilon_{n}\to 0 as nn\to\infty. We define

wn(x):=εn2γ+1u(εnx) for x+N.w_{n}(x):=\varepsilon_{n}^{-\frac{2}{\gamma+1}}u(\varepsilon_{n}x)\quad\text{ for }x\in\mathbb{R}^{N}_{+}.

For nn sufficiently large, we deduce from (15)

ca2γ+1wn(x)CA2γ+1 in ΣAΣaca^{\frac{2}{\gamma+1}}\leq w_{n}(x)\leq CA^{\frac{2}{\gamma+1}}\quad\text{ in }\Sigma_{A}\setminus\Sigma_{a} (16)

and

wn(x)Ca2γ+1 on {xN=a}.w_{n}(x)\leq Ca^{\frac{2}{\gamma+1}}\quad\text{ on }\{x_{N}=a\}. (17)

Moreover, wnw_{n} solves

Δwn=1wnγ+εn2γγ+1g(εn2γ+1wn) in +N.-\Delta w_{n}=\frac{1}{w_{n}^{\gamma}}+\varepsilon_{n}^{\frac{2\gamma}{\gamma+1}}g(\varepsilon_{n}^{\frac{2}{\gamma+1}}w_{n})\quad\text{ in }\mathbb{R}^{N}_{+}. (18)

Since the right hand side of (18) is uniformly bounded in ΣAΣa\Sigma_{A}\setminus\Sigma_{a} and by the standard regularity [20], (wn)(w_{n}) is uniformly bounded in C2,α(ΣAΣa¯)C^{2,\alpha}(\overline{\Sigma_{A}\setminus\Sigma_{a}}), for some 0<α<10<\alpha<1. Since

|wn(x)|=εnγ1γ+1|u(εnx)|εnγ1γ+1u(εnx)η,|\nabla w_{n}(x)|=\varepsilon_{n}^{\frac{\gamma-1}{\gamma+1}}|\nabla u(\varepsilon_{n}x)|\geq\varepsilon_{n}^{\frac{\gamma-1}{\gamma+1}}\frac{\partial u(\varepsilon_{n}x)}{\partial\eta},

for εn\varepsilon_{n} sufficiently small we get the estimate from above in (5).

Now we prove the estimate from below. Suppose by contradiction that there exist β>0\beta>0, a sequence of normal vectors ηn𝕊+N1\eta_{n}\in\mathbb{S}^{N-1}_{+} with (ηn,eN)β(\eta_{n},e_{N})\geq\beta and a sequence of points xn=(xn,xn,N)+Nx_{n}=(x_{n}^{\prime},x_{n,N})\in\mathbb{R}^{N}_{+} such that

xn,Nγ1γ+1u(xn)ηn0 and xn,N0 as n.x_{n,N}^{\frac{\gamma-1}{\gamma+1}}\frac{\partial u(x_{n})}{\partial\eta_{n}}\to 0\text{ and }x_{n,N}\to 0\quad\text{ as }n\to\infty. (19)

Passing to a subsequence, we may assume ηnη𝕊+N1\eta_{n}\to\eta\in\mathbb{S}^{N-1}_{+} with (η,eN)β(\eta,e_{N})\geq\beta as nn\to\infty. We define wnw_{n} as above with εn=xn,N\varepsilon_{n}=x_{n,N} and w~n(x,xN)=wn(x+εn1xn,xN)\tilde{w}_{n}(x^{\prime},x_{N})=w_{n}(x^{\prime}+\varepsilon_{n}^{-1}x^{\prime}_{n},x_{N}), namely,

w~n(x):=xn,N2γ+1u(xn,Nx+xn,xn,NxN) for x+N.\tilde{w}_{n}(x):=x_{n,N}^{-\frac{2}{\gamma+1}}u(x_{n,N}x^{\prime}+x_{n}^{\prime},x_{n,N}x_{N})\quad\text{ for }x\in\mathbb{R}^{N}_{+}.

Then (16), (17) and (18) still hold for w~n\tilde{w}_{n}. Since (w~n)(\tilde{w}_{n}) is uniformly bounded in C2,α(ΣAΣa¯)C^{2,\alpha}(\overline{\Sigma_{A}\setminus\Sigma_{a}}), up to a subsequence, we have

w~nwa,A in Cloc2(ΣAΣa¯).\tilde{w}_{n}\to w_{a,A}\quad\text{ in }C^{2}_{\rm loc}(\overline{\Sigma_{A}\setminus\Sigma_{a}}).

Moreover, passing (18) to the limit, we get

Δwa,A=1wa,Aγ in ΣAΣa.-\Delta w_{a,A}=\frac{1}{w_{a,A}^{\gamma}}\quad\text{ in }\Sigma_{A}\setminus\Sigma_{a}.

Now we take a=1ja=\frac{1}{j} and A=jA=j, for large jj\in\mathbb{N} and we construct w1j,jw_{\frac{1}{j},j} as above. For jj\to\infty, using a standard diagonal process, we can construct a limiting profile wCloc2(+N)w_{\infty}\in C^{2}_{\rm loc}(\mathbb{R}^{N}_{+}) so that

Δw=1wγ in +N-\Delta w_{\infty}=\frac{1}{w_{\infty}^{\gamma}}\quad\text{ in }\mathbb{R}^{N}_{+}

and w1j,j=ww_{\frac{1}{j},j}=w_{\infty} in ΣjΣ1j\Sigma_{j}\setminus\Sigma_{\frac{1}{j}}. Moreover, from (17) we know that

limxN0+w(x)=0 uniformly in xN1.\lim_{x_{N}\to 0^{+}}w_{\infty}(x)=0\quad\text{ uniformly in }x^{\prime}\in\mathbb{R}^{N-1}.

Hence ww_{\infty} is a solution to (4). By [25, Theorem 1], ww_{\infty} depends only on xNx_{N} and w>0w_{\infty}^{\prime}>0 in +\mathbb{R}_{+}.

On the other hand, (19) gives w~n(eN)ηn=xn,Nγ1γ+1u(xn)ηn0\frac{\partial\tilde{w}_{n}(e_{N})}{\partial\eta_{n}}=x_{n,N}^{\frac{\gamma-1}{\gamma+1}}\frac{\partial u(x_{n})}{\partial\eta_{n}}\to 0 as nn\to\infty. This is a contradiction since w~n(eN)ηnw(eN)η=w(1)ηN>0\frac{\partial\tilde{w}_{n}(e_{N})}{\partial\eta_{n}}\to\frac{\partial w_{\infty}(e_{N})}{\partial\eta}=w_{\infty}^{\prime}(1)\eta_{N}>0. ∎

Remark 1.

The proof indicates the following estimate which is stronger than the upper bound in (5)

|u(x)|<c2xN1γγ+1 in Σλ1|\nabla u(x)|<c_{2}x_{N}^{\frac{1-\gamma}{\gamma+1}}\quad\text{ in }\Sigma_{\lambda_{1}}

for some λ1,c2>0\lambda_{1},c_{2}>0 independent of β\beta.

3.2. Rigidity of solutions

In this subsection, we prove Theorem 3. We will make use of the following version of the maximum principle in unbounded domains which is due to Berestycki, Caffarelli and Nirenberg.

Lemma 10 (Lemma 2.1 in [7]).

Let DD be a domain (open connected set) in N\mathbb{R}^{N}, possibly unbounded. Assume that D¯\overline{D} is disjoint from the closure of an infinite open connected cone Σ\Sigma. Suppose there is a function wC2(D)C(D¯)w\in C^{2}(D)\cap C(\overline{D}) that is bounded above and satisfies for some continuous function c(x)c(x),

Δwc(x)w0\displaystyle-\Delta w-c(x)w\leq 0 in D with c(x)0,\displaystyle\text{ in }D\text{ with }c(x)\leq 0,
w0\displaystyle w\leq 0 on D.\displaystyle\text{ on }\partial D.

Then w0w\leq 0 in DD.

Lemma 10 allows us to derive a weak comparison principle. Notice that in the following result, we do not assume that vv is bounded from above.

Proposition 11.

Let f:(0,+)f:(0,+\infty) be a locally Lipschitz continuous function which is non-increasing on (t1,+)(t_{1},+\infty) for some t1>0t_{1}>0. Let u,vC2(Ω)C(Ω¯)u,v\in C^{2}(\Omega)\cap C(\overline{\Omega})  be solutions to

{Δw=f(w) in Ω,w>0 in Ω,\begin{cases}-\Delta w=f(w)&\text{ in }\Omega,\\ w>0&\text{ in }\Omega,\end{cases}

where ΩN\Omega\subset\mathbb{R}^{N} is an open connected set such that NΩ¯\mathbb{R}^{N}\setminus\overline{\Omega} contains an infinite open connected cone. Assume that

uv on Ω,v(x)t1 in Ωu\leq v\text{ on }\partial\Omega,\quad v(x)\geq t_{1}\text{ in }\Omega

and

supΩ(uv)<+.\sup_{\Omega}(u-v)<+\infty.

Then uvu\leq v in Ω\Omega.

Proof.

Assume by contradiction that u>vu>v somewhere in Ω\Omega. Let DD be a connected component of the set where u>vu>v. Setting

w=uv,w=u-v,

then

{Δwc(x)w=0 in D,w=0 on D,\begin{cases}-\Delta w-c(x)w=0&\text{ in }D,\\ w=0&\text{ on }\partial D,\end{cases}

where

c(x)=f(u(x))f(v(x))u(x)v(x).c(x)=\frac{f(u(x))-f(v(x))}{u(x)-v(x)}.

Moreover, since t1v<ut_{1}\leq v<u in DD and ff is non-increasing on (t1,+)(t_{1},+\infty), we deduce c(x)0c(x)\leq 0 in DD. Hence Lemma 10 applies to yield w0w\leq 0 in DD, a contradiction. Therefore, uvu\leq v in Ω\Omega. ∎

We employ the technique from [25, Proposition 5] to show that solutions to problem (1) grow at most at a linear rate as xN+x_{N}\to+\infty.

Lemma 12.

Under the assumptions of Theorem 3, there exists a constant C>0C>0 such that

u(x)CxN in +NΣλ¯.u(x)\leq Cx_{N}\quad\text{ in }\mathbb{R}^{N}_{+}\setminus\Sigma_{\overline{\lambda}}.
Proof.

If uu is a solution to (1), then

v(x):=(λ¯2)2γ+1u(λ¯2x)v(x):=\left(\frac{\overline{\lambda}}{2}\right)^{-\frac{2}{\gamma+1}}u\left(\frac{\overline{\lambda}}{2}x\right)

is bounded in Σ2\Sigma_{2} and vv satisfies

Δv=(λ¯2)2γγ+1f((λ¯2)2γ+1v).-\Delta v=\left(\frac{\overline{\lambda}}{2}\right)^{\frac{2\gamma}{\gamma+1}}f\left(\left(\frac{\overline{\lambda}}{2}\right)^{\frac{2}{\gamma+1}}v\right).

Moreover, the function g(t):=(λ¯2)2γγ+1f((λ¯2)2γ+1t)g(t):=\left(\frac{\overline{\lambda}}{2}\right)^{\frac{2\gamma}{\gamma+1}}f\left(\left(\frac{\overline{\lambda}}{2}\right)^{\frac{2}{\gamma+1}}t\right) still satisfies (i) and (ii) in Theorem 3 with possible different parameters c0,c1,t1c_{0},c_{1},t_{1}. Therefore, without loss of generality, we may assume that our solution uu is bounded in the strip Σ2\Sigma_{2}.

From (i) and  Lemma 8, we have

u(x)C~xN2γ+1 in +N.u(x)\geq\tilde{C}x_{N}^{\frac{2}{\gamma+1}}\quad\text{ in }\mathbb{R}^{N}_{+}. (20)

From (ii) and the compactness, there exists c2>0c_{2}>0 such that

f(t)<c2tγ for all tC~.f(t)<\frac{c_{2}}{t^{\gamma}}\quad\text{ for all }t\geq\tilde{C}. (21)

Let any x0=(x0,x0,N)+Nx_{0}=(x_{0}^{\prime},x_{0,N})\in\mathbb{R}^{N}_{+} with x0,N:=4R>2x_{0,N}:=4R>2. We set

uR(x):=R2γ+1u(x0+R(xx0)),u_{R}(x):=R^{-\frac{2}{\gamma+1}}u(x_{0}+R(x-x_{0})),

then uR>0u_{R}>0 in B4(x0)B_{4}(x_{0}) and

ΔuR(x)=R2γγ+1f(u(x0+R(xx0)))=R2f(u(x0+R(xx0)))u(x0+R(xx0)uR(x).-\Delta u_{R}(x)=R^{\frac{2\gamma}{\gamma+1}}f(u(x_{0}+R(x-x_{0})))=\frac{R^{2}f(u(x_{0}+R(x-x_{0})))}{u(x_{0}+R(x-x_{0})}u_{R}(x).

Remark that x0+R(xx0)+NΣ2R+NΣ1x_{0}+R(x-x_{0})\in\mathbb{R}^{N}_{+}\setminus\Sigma_{2R}\subset\mathbb{R}^{N}_{+}\setminus\Sigma_{1} for xB2(x0)x\in B_{2}(x_{0}). Hence (20) and (21) give

R2f(u(x0+R(xx0))u(x0+R(xx0)c2R2u(x0+R(xx0))γ+1c24C~γ+1 in B2(x0).\frac{R^{2}f(u(x_{0}+R(x-x_{0}))}{u(x_{0}+R(x-x_{0})}\leq\frac{c_{2}R^{2}}{u(x_{0}+R(x-x_{0}))^{\gamma+1}}\leq\frac{c_{2}}{4\tilde{C}^{\gamma+1}}\quad\text{ in }B_{2}(x_{0}).

By Harnack’s inequality, we have

supB1(x0)uRCHinfB1(x0)uR,\sup_{B_{1}(x_{0})}u_{R}\leq C_{H}\inf_{B_{1}(x_{0})}u_{R},

which implies

supBR(x0)uCHinfBR(x0)u,\sup_{B_{R}(x_{0})}u\leq C_{H}\inf_{B_{R}(x_{0})}u,

where CH>0C_{H}>0 is independent of x0x_{0}. From this point, we can proceed as in the proof of [25, Proposition 5] to get the thesis. ∎

Given the previous asymptotic bound on uu, we can apply the scaling technique as in [25, Proposition 7] to establish a bound on the gradient.

Lemma 13.

Under the assumptions of Theorem 3, there exists a constant C>0C>0 such that

|u(x)|C in +NΣλ¯.|\nabla u(x)|\leq C\quad\text{ in }\mathbb{R}^{N}_{+}\setminus\Sigma_{\overline{\lambda}}.
Proof.

As in Lemma 12, we may assume λ¯=2\overline{\lambda}=2. Let x0+NΣ2x_{0}\in\mathbb{R}^{N}_{+}\setminus\Sigma_{2} and set R=x0,N2R=x_{0,N}\geq 2. We define

uR(x):=u(Rx)R in B12(x0R).u_{R}(x):=\frac{u(Rx)}{R}\quad\text{ in }B_{\frac{1}{2}}\left(\frac{x_{0}}{R}\right).

By Lemma 12 we have uRCu_{R}\leq C. Moreover, from (20) and (21), we deduce

ΔuR=Rf(u(Rx))c2Ru(Rx)γ4γγ+1c2C~γRγ1γ+12c2C~γ in B12(x0R).-\Delta u_{R}=Rf(u(Rx))\leq\frac{c_{2}R}{u(Rx)^{\gamma}}\leq\frac{4^{\frac{\gamma}{\gamma+1}}c_{2}}{\tilde{C}^{\gamma}}R^{-\frac{\gamma-1}{\gamma+1}}\leq\frac{2c_{2}}{\tilde{C}^{\gamma}}\quad\text{ in }B_{\frac{1}{2}}\left(\frac{x_{0}}{R}\right).

By the standard gradient estimate, we have |uR|C|\nabla u_{R}|\leq C^{\prime} in B14(x0R)B_{\frac{1}{4}}\left(\frac{x_{0}}{R}\right). This indicates |u|C|\nabla u|\leq C^{\prime} in BR4(x0)B_{\frac{R}{4}}(x_{0}). The thesis follows from the arbitrariness of x0x_{0}. ∎

We are ready to prove Theorem 3 by employing the sliding method.

Proof of Theorem 3.

For each λ>0\lambda>0 and ν𝕊+N1\nu\in\mathbb{S}^{N-1}_{+}, we define

uλν(x):=u(x+λν).u_{\lambda}^{\nu}(x):=u(x+\lambda\nu).

We aim to show that

uuλν in +N for all λ>0.u\leq u_{\lambda}^{\nu}\,\text{ in }\mathbb{R}^{N}_{+}\quad\text{ for all }\lambda>0. (22)

By Lemma 8, there exists λ>0\lambda^{*}>0 such that

u(x)>t1 for xN>λ.u(x)>t_{1}\text{ for }x_{N}>\lambda^{*}.

Let λ>λν\lambda>\lambda_{\nu}^{*}, where λν:=λ(ν,eN)\lambda_{\nu}^{*}:=\frac{\lambda^{*}}{(\nu,e_{N})}, then uλν>t1u_{\lambda}^{\nu}>t_{1} in +N\mathbb{R}^{N}_{+}. Moreover, from Lemmas 12, 13 and the mean value theorem, we deduce

sup+N(uuλν)<+.\sup_{\mathbb{R}^{N}_{+}}(u-u_{\lambda}^{\nu})<+\infty.

Since

{Δu=f(u) in +N,Δuλν=f(uλν) in +N,u>0 in +N,uλν>t1 in +N,uuλν on +N,\begin{cases}-\Delta u=f(u)&\text{ in }\mathbb{R}^{N}_{+},\\ -\Delta u_{\lambda}^{\nu}=f(u_{\lambda}^{\nu})&\text{ in }\mathbb{R}^{N}_{+},\\ u>0&\text{ in }\mathbb{R}^{N}_{+},\\ u_{\lambda}^{\nu}>t_{1}&\text{ in }\mathbb{R}^{N}_{+},\\ u\leq u_{\lambda}^{\nu}&\text{ on }\partial\mathbb{R}^{N}_{+},\end{cases}

we can apply Proposition 11 to derive

uuλν in +N for all λ>λν.u\leq u_{\lambda}^{\nu}\,\text{ in }\mathbb{R}^{N}_{+}\quad\text{ for all }\lambda>\lambda_{\nu}^{*}.

Now that the set

Λ={λ>0uuμν in +N for all μ>λ}\Lambda=\{\lambda>0\mid u\leq u_{\mu}^{\nu}\text{ in }\mathbb{R}^{N}_{+}\text{ for all }\mu>\lambda\}

is nonempty, we can define λ0=infΛ\lambda_{0}=\inf\Lambda. We will show that

λ0=0.\lambda_{0}=0.

Assume on contrary that λ0>0\lambda_{0}>0. By continuity of uu, we have uuλ0νu\leq u_{\lambda_{0}}^{\nu} in +N\mathbb{R}^{N}_{+}. In order to reach a contradiction, we will search for some ε0\varepsilon_{0} small such that

uuλν in +Nu\leq u_{\lambda}^{\nu}\quad\text{ in }\mathbb{R}^{N}_{+} (23)

for all λ(λ0ε0,λ0]\lambda\in(\lambda_{0}-\varepsilon_{0},\lambda_{0}].

\circ Due to Lemmas 6 and 8, there exist λ~,δ~>0\tilde{\lambda},\tilde{\delta}>0 sufficiently small such that

u+δ~uλν in Σλ~u+\tilde{\delta}\leq u_{\lambda}^{\nu}\quad\text{ in }\Sigma_{\tilde{\lambda}} (24)

for all λ>λ0/2\lambda>\lambda_{0}/2.

\circ We claim that

uuλν in ΣλΣλ~u\leq u_{\lambda}^{\nu}\quad\text{ in }\Sigma_{\lambda^{*}}\setminus\Sigma_{\tilde{\lambda}} (25)

for all λ(λ0ε0,λ0)\lambda\in(\lambda_{0}-\varepsilon_{0},\lambda_{0}), where ε0>0\varepsilon_{0}>0 is sufficiently small.

Assume that (25) does not hold. Then there exist two sequences λnλ0\lambda_{n}\nearrow\lambda_{0} and xn:=(xn,xn,N)N1×[λ~,λ)x_{n}:=(x^{\prime}_{n},x_{n,N})\in\mathbb{R}^{N-1}\times[\tilde{\lambda},\lambda^{*}) such that

u(xn)>uλnν(xn).u(x_{n})>u_{\lambda_{n}}^{\nu}(x_{n}). (26)

Moreover, we may assume xn,Ny0[λ~,λ]x_{n,N}\to y_{0}\in[\tilde{\lambda},\lambda^{*}]. Now we set

un(x,xN)=u(x+xn,xN).u_{n}(x^{\prime},x_{N})=u(x^{\prime}+x^{\prime}_{n},x_{N}).

Since Cλ~2γ+1unuL(Σλ)C\tilde{\lambda}^{\frac{2}{\gamma+1}}\leq u_{n}\leq\|u\|_{L^{\infty}(\Sigma_{\lambda})} in ΣλΣλ~\Sigma_{\lambda}\setminus\Sigma_{\tilde{\lambda}}, we have that f(un)f(u_{n}) is bounded in ΣλΣλ~\Sigma_{\lambda}\setminus\Sigma_{\tilde{\lambda}} for each λ>λ~\lambda>\tilde{\lambda}. The standard regularity gives unC2,α(ΣλΣλ~¯)<Cλ\|u_{n}\|_{C^{2,\alpha}(\overline{\Sigma_{\lambda}\setminus\Sigma_{\tilde{\lambda}}})}<C_{\lambda} for some 0<α<10<\alpha<1. By the Arzelà–Ascoli theorem, via a standard diagonal process, we have

unv in Cloc2(+NΣλ~¯)u_{n}\to v\quad\text{ in }C^{2}_{\rm loc}(\overline{\mathbb{R}^{N}_{+}\setminus\Sigma_{\tilde{\lambda}}})

up to a subsequence. Moreover, vv weakly solves Δv=f(v)-\Delta v=f(v) in +NΣλ~¯\mathbb{R}^{N}_{+}\setminus\overline{\Sigma_{\tilde{\lambda}}}. Using the definition of λ0\lambda_{0} and passing (26) to the limit, we have

vvλ0ν in +NΣλ~,\displaystyle v\leq v_{\lambda_{0}}^{\nu}\quad\text{ in }\mathbb{R}^{N}_{+}\setminus\Sigma_{\tilde{\lambda}},
v(x0)=vλ0ν(x0),\displaystyle v(x_{0})=v_{\lambda_{0}}^{\nu}(x_{0}),

where x0=(0,y0)x_{0}=(0^{\prime},y_{0}). On the other hand, by (24) we have v+δ~vλ0νv+\tilde{\delta}\leq v_{\lambda_{0}}^{\nu} on {xN=λ~}\{x_{N}=\tilde{\lambda}\}. Hence the strong comparison principle implies v<vλ0νv<v_{\lambda_{0}}^{\nu} in +NΣλ~\mathbb{R}^{N}_{+}\setminus\Sigma_{\tilde{\lambda}}. This contradicts the fact that v(x0)=vλ0ν(x0)v(x_{0})=v_{\lambda_{0}}^{\nu}(x_{0}). Therefore, (25) must hold.

\circ Next, we show that

uuλν in +NΣλu\leq u_{\lambda}^{\nu}\quad\text{ in }\mathbb{R}^{N}_{+}\setminus\Sigma_{\lambda^{*}} (27)

for all λ(λ0ε0,λ0)\lambda\in(\lambda_{0}-\varepsilon_{0},\lambda_{0}).

From (25) and the continuity, we already have uuλνu\leq u_{\lambda}^{\nu} on {xN=λ}\{x_{N}=\lambda^{*}\}. Moreover, uλν(x)t1u_{\lambda}^{\nu}(x)\geq t_{1} for each x+NΣλx\in\mathbb{R}^{N}_{+}\setminus\Sigma_{\lambda^{*}}. Hence (27) follows by applying Proposition 11 with uu and v:=uλνv:=u_{\lambda}^{\nu} on +NΣλ\mathbb{R}^{N}_{+}\setminus\Sigma_{\lambda^{*}}.

Combining (24), (25) and (27), we obtain (23). This contradicts the definition of λ0\lambda_{0} and hence (22) is proved.

Therefore, uu is monotone increasing in direction ν\nu for all ν𝕊+N1\nu\in\mathbb{S}^{N-1}_{+}. That is,

uν:=(u,ν)0 in +N.\frac{\partial u}{\partial\nu}:=(\nabla u,\nu)\geq 0\quad\text{ in }\mathbb{R}^{N}_{+}.

To deduce the one-dimensional symmetry of uu, we take ζ\zeta be any direction in {xB1(0)xN=0}\{x\in\partial B_{1}(0)\mid x_{N}=0\}. Let νn𝕊+N1\nu_{n}\in\mathbb{S}^{N-1}_{+} be a sequence converging to ζ\zeta, we have uνn0\frac{\partial u}{\partial\nu_{n}}\geq 0. By sending nn\to\infty, we deduce uζ0\frac{\partial u}{\partial\zeta}\geq 0 in +N\mathbb{R}^{N}_{+}. Similarly, let another sequence τn𝕊+N1\tau_{n}\in\mathbb{S}^{N-1}_{+} converging to ζ-\zeta, we obtain uζ0\frac{\partial u}{\partial\zeta}\leq 0 in +N\mathbb{R}^{N}_{+}. Therefore, uu is constant in direction ζ\zeta. Since ζ\zeta is arbitrary, we deduce that uu does not depend on xx^{\prime}. Hence uu depends only on xNx_{N} and monotone increasing in xNx_{N}.

By the Höpf lemma [20], we have actually uxN>0\frac{\partial u}{\partial x_{N}}>0 in +N\mathbb{R}^{N}_{+}. By writing v(xN)=u(x)v(x_{N})=u(x), problem (1) reduces to

{v′′=f(v) in +,v(t)>0 in +,v(0)=0.\begin{cases}-v^{\prime\prime}=f(v)&\text{ in }\mathbb{R}_{+},\\ v^{\prime}(t)>0&\text{ in }\mathbb{R}_{+},\\ v(0)=0.\end{cases}

Hence for every t>0t>0, we have

12(v)2F(v)=M,\frac{1}{2}(v^{\prime})^{2}-F(v)=M, (28)

which is a constant. Letting t+t\to+\infty and noticing that v(t)+v(t)\to+\infty by Lemma 8, we deduce M0M\geq 0. By integrating (28) and using v(0)=0v(0)=0, this gives

0v(t)dsM+F(s)=2t for all t0.\int_{0}^{v(t)}\frac{ds}{\sqrt{M+F(s)}}=\sqrt{2}t\quad\text{ for all }t\geq 0. (29)

Conversely, for every M0M\geq 0 we have

0+dsM+F(s)=+ and 0tdsM+F(s)<+ for all t>0.\int_{0}^{+\infty}\frac{ds}{\sqrt{M+F(s)}}=+\infty\text{ and }\int_{0}^{t}\frac{ds}{\sqrt{M+F(s)}}<+\infty\text{ for all }t>0.

Therefore, for each M0M\geq 0, formula (29) uniquely determines a function v:=vMv:=v_{M} which is a solution to (1). It is also clear from (28) that each solution vMv_{M} is characterized by the property limt+vM(t)=2M\lim_{t\to+\infty}v_{M}^{\prime}(t)=\sqrt{2M}. ∎

Finally, we discuss the case that ff is nonincreasing in the whole domain (0,+)(0,+\infty).

Proof of Theorem 4.

The proof is almost the same as that of [26, Theorem 6], so we only comment on the difference. By employing the Kelvin transform

u^(x):=1|x|N2u(x|x|2),\hat{u}(x):=\frac{1}{|x|^{N-2}}u\left(\frac{x}{|x|^{2}}\right),

we deduce that u^C2(+N)C(+N¯{0})\hat{u}\in C^{2}(\mathbb{R}^{N}_{+})\cap C(\overline{\mathbb{R}^{N}_{+}}\setminus\{0\}) and u^\hat{u} is a solution to

{Δu^=1|x|N+2f(|x|N2u^) in +N,u^>0 in +N,u^=0 on +N{0}.\begin{cases}-\Delta\hat{u}=\dfrac{1}{|x|^{N+2}}f(|x|^{N-2}\hat{u})&\text{ in }\mathbb{R}^{N}_{+},\\ \hat{u}>0&\text{ in }\mathbb{R}^{N}_{+},\\ \hat{u}=0&\text{ on }\partial\mathbb{R}^{N}_{+}\setminus\{0\}.\end{cases}

As in [26] we denote Σλ={(x1,x)+Nx1<λ}\Sigma_{\lambda}=\{(x_{1},x^{\prime})\in\mathbb{R}^{N}_{+}\mid x_{1}<\lambda\}, xλ=(2λx1,x)x_{\lambda}=(2\lambda-x_{1},x^{\prime}) and u^λ(x)=u^(xλ)\hat{u}_{\lambda}(x)=\hat{u}(x_{\lambda}). Then for test functions of type w=(u^u^λτ)+ψχΣλw=(\hat{u}-\hat{u}_{\lambda}-\tau)^{+}\psi\chi_{\Sigma_{\lambda}} with compact support in Σλ\Sigma_{\lambda} and τ>0\tau>0 we have

Σλ((u^u^λ),w)=Σλ(1|x|N+2f(|x|N2u^)1|xλ|N+2f(|xλ|N2u^λ))w0\int_{\Sigma_{\lambda}}(\nabla(\hat{u}-\hat{u}_{\lambda}),\nabla w)=\int_{\Sigma_{\lambda}}\left(\frac{1}{|x|^{N+2}}f(|x|^{N-2}\hat{u})-\frac{1}{|x_{\lambda}|^{N+2}}f(|x_{\lambda}|^{N-2}\hat{u}_{\lambda})\right)w\leq 0

since u^u^λ\hat{u}\geq\hat{u}_{\lambda} on the support of ww and |x||xλ||x|\geq|x_{\lambda}| in Σλ\Sigma_{\lambda}. From this inequality, we can argue as in the proof of [26, Theorem 6] and repeat the arguments there to get u(x)=v(xN)u(x)=v(x_{N}), where vv is a solution to

{v′′=f(v) in +,v(t)>0 in +,v(0)=0.\begin{cases}-v^{\prime\prime}=f(v)&\text{ in }\mathbb{R}_{+},\\ v(t)>0&\text{ in }\mathbb{R}_{+},\\ v(0)=0.\end{cases}

By Theorem 1 we have v(t)>0v^{\prime}(t)>0 in +\mathbb{R}_{+}. Moreover, since v′′0-v^{\prime\prime}\geq 0 we have that vv^{\prime} is nondecreasing and hence limt+v(t)=+\lim_{t\to+\infty}v(t)=+\infty. From v′′=f(v)-v^{\prime\prime}=f(v) we deduce

12(v)2+F1(v)=M1,\frac{1}{2}(v^{\prime})^{2}+F_{1}(v)=M_{1},

where F1(s)=1sf(t)𝑑tF_{1}(s)=\int_{1}^{s}f(t)dt and M1M_{1} is a constant. Sending t+t\to+\infty, we obtain M11+f(t)𝑑tM_{1}\geq\int_{1}^{+\infty}f(t)dt. In particular, 1+f(t)𝑑t\int_{1}^{+\infty}f(t)dt is finite. Hence also F(s)=s+f(t)𝑑tF(s)=\int_{s}^{+\infty}f(t)dt is finite for all s>0s>0. Similar to the proof of Theorem 3 we deduce

0v(t)dsM+F(s)=2t for all t0 and some M0\int_{0}^{v(t)}\frac{ds}{\sqrt{M+F(s)}}=\sqrt{2}t\quad\text{ for all }t\geq 0\text{ and some }M\geq 0 (30)

and (30) indeed provides a solution to our problem. ∎

Conflict of interest The author declares no conflict of interest.

Data Availability Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

References

  • [1] A. D. Alexandrov. A characteristic property of spheres. Ann. Mat. Pura Appl. (4), 58:303–315, 1962. doi:10.1007/BF02413056.
  • [2] S. B. Angenent. Uniqueness of the solution of a semilinear boundary value problem. Math. Ann., 272(1):129–138, 1985. doi:10.1007/BF01455933.
  • [3] D. Arcoya, L. Boccardo, T. Leonori, and A. Porretta. Some elliptic problems with singular natural growth lower order terms. J. Differential Equations, 249(11):2771–2795, 2010. doi:10.1016/j.jde.2010.05.009.
  • [4] H. Berestycki, L. Caffarelli, and L. Nirenberg. Further qualitative properties for elliptic equations in unbounded domains. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 25(1-2):69–94 (1998), 1997. Dedicated to Ennio De Giorgi. URL: http://www.numdam.org/item?id=ASNSP_1997_4_25_1-2_69_0.
  • [5] H. Berestycki, L. A. Caffarelli, and L. Nirenberg. Symmetry for elliptic equations in a half space. In Boundary value problems for partial differential equations and applications, volume 29 of RMA Res. Notes Appl. Math., pages 27–42. Masson, Paris, 1993.
  • [6] H. Berestycki, L. A. Caffarelli, and L. Nirenberg. Inequalities for second-order elliptic equations with applications to unbounded domains. I. Duke Math. J., 81(2):467–494, 1996. A celebration of John F. Nash, Jr. doi:10.1215/S0012-7094-96-08117-X.
  • [7] H. Berestycki, L. A. Caffarelli, and L. Nirenberg. Monotonicity for elliptic equations in unbounded Lipschitz domains. Comm. Pure Appl. Math., 50(11):1089–1111, 1997. doi:10.1002/(SICI)1097-0312(199711)50:11<1089::AID-CPA2>3.0.CO;2-6.
  • [8] H. Berestycki and L. Nirenberg. On the method of moving planes and the sliding method. Bol. Soc. Brasil. Mat. (N.S.), 22(1):1–37, 1991. doi:10.1007/BF01244896.
  • [9] L. Boccardo. A Dirichlet problem with singular and supercritical nonlinearities. Nonlinear Anal., 75(12):4436–4440, 2012. doi:10.1016/j.na.2011.09.026.
  • [10] L. Boccardo and L. Orsina. Semilinear elliptic equations with singular nonlinearities. Calc. Var. Partial Differential Equations, 37(3-4):363–380, 2010. doi:10.1007/s00526-009-0266-x.
  • [11] A. Canino, F. Esposito, and B. Sciunzi. On the Höpf boundary lemma for singular semilinear elliptic equations. J. Differential Equations, 266(9):5488–5499, 2019. doi:10.1016/j.jde.2018.10.039.
  • [12] A. Canino, M. Grandinetti, and B. Sciunzi. Symmetry of solutions of some semilinear elliptic equations with singular nonlinearities. J. Differential Equations, 255(12):4437–4447, 2013. doi:10.1016/j.jde.2013.08.014.
  • [13] P. Clément and G. Sweers. Existence and multiplicity results for a semilinear elliptic eigenvalue problem. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 14(1):97–121, 1987. URL: http://www.numdam.org/item?id=ASNSP_1987_4_14_1_97_0.
  • [14] M. G. Crandall, P. H. Rabinowitz, and L. Tartar. On a Dirichlet problem with a singular nonlinearity. Comm. Partial Differential Equations, 2(2):193–222, 1977. doi:10.1080/03605307708820029.
  • [15] F. Esposito, A. Farina, and B. Sciunzi. Qualitative properties of singular solutions to semilinear elliptic problems. J. Differential Equations, 265(5):1962–1983, 2018. doi:10.1016/j.jde.2018.04.030.
  • [16] F. Esposito and B. Sciunzi. The moving plane method for doubly singular elliptic equations involving a first-order term. Adv. Nonlinear Stud., 21(4):905–916, 2021. doi:10.1515/ans-2021-2151.
  • [17] A. Farina. Some results about semilinear elliptic problems on half-spaces. Math. Eng., 2(4):709–721, 2020. doi:10.3934/mine.2020033.
  • [18] A. Farina and B. Sciunzi. Qualitative properties and classification of nonnegative solutions to Δu=f(u)-\Delta u=f(u) in unbounded domains when f(0)<0f(0)<0. Rev. Mat. Iberoam., 32(4):1311–1330, 2016. doi:10.4171/RMI/918.
  • [19] A. Farina and B. Sciunzi. Monotonicity and symmetry of nonnegative solutions to Δu=f(u)-\Delta u=f(u) in half-planes and strips. Adv. Nonlinear Stud., 17(2):297–310, 2017. doi:10.1515/ans-2017-0010.
  • [20] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [21] N. Hirano, C. Saccon, and N. Shioji. Existence of multiple positive solutions for singular elliptic problems with concave and convex nonlinearities. Adv. Differential Equations, 9(1-2):197–220, 2004.
  • [22] B. Kawohl. On a class of singular elliptic equations. In Progress in partial differential equations: elliptic and parabolic problems (Pont-à-Mousson, 1991), volume 266 of Pitman Res. Notes Math. Ser., pages 156–163. Longman Sci. Tech., Harlow, 1992.
  • [23] A. C. Lazer and P. J. McKenna. On a singular nonlinear elliptic boundary-value problem. Proc. Amer. Math. Soc., 111(3):721–730, 1991. doi:10.2307/2048410.
  • [24] P. Le. Doubly singular elliptic equations involving a gradient term: symmetry and monotonicity. Results Math., 79(1):Paper No. 3, 13, 2024. doi:10.1007/s00025-023-02025-y.
  • [25] L. Montoro, L. Muglia, and B. Sciunzi. Classification of solutions to Δu=uγ-\Delta u=u^{-\gamma} in the half-space. Math. Ann., 389(3):3163–3179, 2024. doi:10.1007/s00208-023-02717-4.
  • [26] L. Montoro, L. Muglia, and B. Sciunzi. The Classification of all weak solutions to Δu=uγ-\Delta u={u^{-\gamma}} in the half-space. arXiv e-prints, page arXiv:2404.03343, Apr. 2024. arXiv:2404.03343, doi:10.48550/arXiv.2404.03343.
  • [27] A. Quaas and B. Sirakov. Existence results for nonproper elliptic equations involving the Pucci operator. Comm. Partial Differential Equations, 31(7-9):987–1003, 2006. doi:10.1080/03605300500394421.
  • [28] B. Sciunzi. Classification of positive 𝒟1,p(N)\mathcal{D}^{1,p}(\mathbb{R}^{N})-solutions to the critical pp-Laplace equation in N\mathbb{R}^{N}. Adv. Math., 291:12–23, 2016. doi:10.1016/j.aim.2015.12.028.
  • [29] J. Serrin. A symmetry problem in potential theory. Arch. Rational Mech. Anal., 43:304–318, 1971. doi:10.1007/BF00250468.