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

Two Necessary and Sufficient Conditions to the Solvability of the Exterior Dirichlet Problem for the Monge–Ampère Equation

Cong Wang School of Statistics, University of International Business and Economics, Beijing 100029, China. cong_wang@uibe.edu.cn  and  Jiguang Bao School of Mathematical Sciences, Beijing Normal University, Laboratory of Mathematics and Complex Systems, Ministry of Education, Beijing 100875, China. jgbao@bnu.edu.cn
Abstract.

The present paper provides two necessary and sufficient conditions for the existence of solutions to the exterior Dirichlet problem of the Monge–Ampère equation with prescribed asymptotic behavior at infinity. By an adapted smooth approximation argument, we prove that the problem is solvable if and only if the boundary value is semi-convex with respect to the inner boundary, which is our first proposed new concept. Along the lines of Perron’s method for Laplace equation, we obtain the threshold for solvability in the asymptotic behavior at infinity of the solution, and remove the C2C^{2} regularity assumptions on the boundary value and on the inner boundary which are required in the proofs of the corresponding existence theorems in the recent literatures.

Key words and phrases:
Monge–Ampère equation, Exterior Dirichlet problem, Semi-convex boundary value, Enclosing sphere condition, Necessary and sufficient condition, Solvability.
2010 Mathematics Subject Classification:
35J96, 35J25, 35B40
J. Bao was supported by the National Key Research and Development Program of China (No. 2020YFA0712904)

1. Introduction

Let Ω\Omega be a bounded domain of n\mathbb{R}^{n}, n3n\geq 3, and let φ\varphi be a function on Ω\partial\Omega. In this paper, we intend to explore the solvability of the exterior Dirichlet problem for the Monge–Ampère equation

{det(D2u)=1innΩ¯,u=φonΩ,\begin{cases}\det(D^{2}u)=1\quad\text{in}~{}\mathbb{R}^{n}\setminus\overline{\Omega},\\ u=\varphi\qquad\qquad\quad~{}~{}~{}~{}\text{on}~{}\partial\Omega,\end{cases} (1.1)

provided that Ω\Omega and φ\varphi satisfy some general conditions, without the C2C^{2} regularity like in Caffarelli–Li’s work [6].

The prototypical place where Monge–Ampère equations arise is the Minkowski problem (see [25, 28]). Monge–Ampère equations also play a significant role in the studies of affine geometry (see [27, 10, 30]) and optimal transportation (see [12]). The interior Dirichlet problem for Monge–Ampère equations

{det(D2u)=finΩ,u=φonΩ,\begin{cases}\det(D^{2}u)=f&\text{in}~{}\Omega,\\ u=\varphi&\text{on}~{}\partial\Omega,\\ \end{cases} (1.2)

has a long history, and there have been many excellent results, especially on the solvability in different situations. We list several known existence results for solutions of (1.2), under the assumptions that fC(Ω¯)f\in C^{\infty}(\overline{\Omega}), f>0f>0 on Ω¯\overline{\Omega}, and Ω\Omega is a bounded and convex domain with Ω\partial\Omega containing no line segment. Rauch–Taylor [29] proved that (1.2) has a unique convex solution uC0(Ω¯)u\in C^{0}(\overline{\Omega}) by Perron’s method, when φC0(Ω)\varphi\in C^{0}(\partial\Omega); see also Aleksandrov [1] and Cheng–Yau [9]. Pogorelov [26] obtained that the unique convex solution of (1.2) is smooth in Ω\Omega, when ΩC2\partial\Omega\in C^{2} and φC2(Ω)\varphi\in C^{2}(\partial\Omega). After that, Caffarelli–Nirenberg–Spruck [7] further proved that (1.2) has a unique convex solution uC(Ω¯)u\in C^{\infty}(\overline{\Omega}) by the continuity method, when Ω\Omega is strictly convex with ΩC\partial\Omega\in C^{\infty} and φC(Ω)\varphi\in C^{\infty}(\partial\Omega); see also Krylov [18].

By contrast, less results are known for the exterior Dirichlet problem for Monge–Ampère equations. Differently from the interior Dirichlet problem (1.2), in exterior domains we also require the solutions to satisfy appropriate prescribed asymptotic behavior at infinity in order to restore the well-posedness. Such asymptotic behavior comes from Liouville-type theorems. The celebrated Jörgens–Calabi–Pogorelov theorem [17, 8, 27] states that any classical convex solution of

det(D2u)=1\det(D^{2}u)=1 (1.3)

in n\mathbb{R}^{n} must be a quadratic polynomial. Caffarelli [4] generalized this result to viscosity solutions case. Caffarelli–Li [6] extended the Jörgens–Calabi–Pogorelov theorem to exterior domains. Specifically, they showed that if n3n\geq 3 and uu is a viscosity solution of (1.3) outside a bounded set, then there exist A𝒜A\in\mathcal{A}, bnb\in\mathbb{R}^{n} and cc\in\mathbb{R} such that

lim|x|(u(x)(12xAxT+bx+c))=0,\lim_{|x|\to\infty}\bigg{(}u(x)-\bigg{(}\frac{1}{2}xAx^{T}+b\cdot x+c\bigg{)}\bigg{)}=0,

where

𝒜={A|A is a real n×n symmetric positive definite matrix with det(A)=1}.\mathcal{A}=\{A|\ A\text{ is a real }n\times n\text{ symmetric positive definite matrix with }\det(A)=1\}.

We refer to [2, 21, 33] for more information about the Liouville-type theorems for Hessian equations.

Based on the asymptotic behavior above, Caffarelli–Li [6] proposed the following exterior Dirichlet problem

{det(D2u)=1innΩ¯,u=φonΩ,lim|x|(u(x)(12xAxT+bx+c))=0.\begin{cases}\det(D^{2}u)=1\quad\text{in}~{}\mathbb{R}^{n}\setminus\overline{\Omega},\\ u=\varphi\quad\text{on}~{}\partial\Omega,\\ \lim_{|x|\to\infty}\Big{(}u(x)-\Big{(}\frac{1}{2}xAx^{T}+b\cdot x+c\Big{)}\Big{)}=0.\end{cases} (1.4)

Under the conditions of ΩC2\partial\Omega\in C^{2} and φC2(Ω)\varphi\in C^{2}(\partial\Omega), Caffarelli–Li [6] proved the existence result by an adapted Perron’s method, and then Li–Lu [22] gave the nonexistence result in terms of the asymptotic behavior. Therefore the characterization of solvability of (1.4) is completed.

Theorem 1.1 (Caffarelli–Li [6] and Li–Lu [22]).

Let Ω\Omega be a bounded, strictly convex domain of n\mathbb{R}^{n}, n3n\geq 3, ΩC2\partial\Omega\in C^{2} and let φC2(Ω)\varphi\in C^{2}(\partial\Omega). Then for any A𝒜A\in\mathbb{\mathcal{A}} and bnb\in\mathbb{R}^{n}, there exists some constant cc_{*}, such that (1.4) has a viscosity solution in C0(nΩ)C^{0}(\mathbb{R}^{n}\setminus\Omega) if and only if ccc\geq c_{*}, where cc_{*} depends only on nn, AA, bb, Ω\Omega and φ\varphi.

Moreover, when the problem (1.4) has a viscosity solution, it is unique by the comparison principle, and interior smooth by [4].

Under the same assumptions on the domain and on the boundary value as in Theorem 1.1, the solvability of the exterior Dirichlet problem was also exploited for kk-Hessian equations [3], Hessian quotient equations [19], special Lagrangian equation [23], and general Hessian-type equations [20, 16]. The ideas therein are similar in spirit to [6].

The aim of this paper is to improve the C2C^{2} regularity condition of both domains and boundary values in Theorem 1.1. We are interested in studying the solvability of (1.4) under the geometry and regularity conditions which are corresponding to that of Laplace equation. For the interior Dirichlet problem for Laplace equation, to guarantee the continuity up to the boundary of the solution, the continuous boundary value is necessary, and the domain needs to be regular (i.e. there exists a barrier function at each boundary point); see [14, Chapter 2.8]. A known sufficient condition that makes the domain be regular is the exterior sphere condition. For the exterior Dirichlet problem for Laplace equation with prescribed limit at infinity, there is a unique solution when n3n\geq 3, the domain is smooth and the boundary is continuous; see for instance Meyers–Serrin [24].

We firstly focus on the boundary value φ\varphi on Ω\partial\Omega. For interior Dirichlet problem (1.2), the solution exists if and only if φC0(Ω)\varphi\in C^{0}(\partial\Omega) due to Rauch–Taylor [29]. While for the exterior Dirichlet problem (1.4), the convexity of nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega} is opposite to that of Ω\Omega, which may lead to that φ\varphi in (1.4) has different and even stronger structure from φC0(Ω)\varphi\in C^{0}(\partial\Omega) in (1.2). Based on such observation, we are naturally motivated to investigate the necessary condition of φ\varphi for the existence of solutions. Unexpectedly, we derive that the “semi-convexity” condition below, which is a different phenomenon from interior Dirichlet problem (1.2).

In order to introduce the new concept of “semi-convexity” clearly, we first introduce the local coordinate system at a boundary point. For any fixed ξΩ\xi\in\partial\Omega, we choose a coordinate system (x,xn)(x^{\prime},x_{n}) such that x=0x=0 at ξ\xi, the positive xnx_{n}-axis directs to the interior of Ω\Omega, and Ω\partial\Omega can be locally represented by the graph of

xn=ρ(x)for|x|<δ(ξ)x_{n}=\rho(x^{\prime})\quad\text{for}~{}|x^{\prime}|<\delta(\xi) (1.5)

for some constant δ(ξ)>0\delta(\xi)>0, where ρ:Bδ(ξ)(0)={xn1||x|<δ(ξ)}\rho:B^{\prime}_{\delta(\xi)}(0^{\prime})=\{x^{\prime}\in\mathbb{R}^{n-1}|\,|x^{\prime}|<\delta(\xi)\}\to\mathbb{R} is a function with ρ(0)=0\rho(0^{\prime})=0. We call such coordinate system the local coordinate system at ξ\xi.

Definition 1.2.

Let φ\varphi be a function defined on Ω\partial\Omega. We say that φ\varphi is semi-convex with respect to Ω\partial\Omega at ξ\xi, if under the local coordinate system at ξ\xi, the function

ψ(x):=φ(x,ρ(x))\psi(x^{\prime}):=\varphi(x^{\prime},\rho(x^{\prime}))

is semi-convex in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}), that is, there exists a constant K(ξ)>0K(\xi)>0, such that ψ(x)+K(ξ)2|x|2\psi(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2} is convex in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}). We say that φ\varphi is semi-convex with respect to Ω\partial\Omega, if φ\varphi is semi-convex with respect to Ω\partial\Omega at each ξΩ\xi\in\partial\Omega.

Our first main result is that the boundary value is necessarily semi-convex with respect to the boundary for the existence of solutions of (1.4).

Theorem 1.3.

Let Ω\Omega be a bounded convex domain of n\mathbb{R}^{n}, n3n\geq 3, ΩC3\partial\Omega\in C^{3}. Let uC0(nΩ)u\in C^{0}(\mathbb{R}^{n}\setminus\Omega) be a viscosity solution of (1.1), then φ\varphi is semi-convex with respect to Ω\partial\Omega.

With the necessity in hand, we are inspired to study the solvability of (1.4) under the semi-convexity condition. We further focus on the domain Ω\Omega. Following [14, Chapter 14.2], we use the geometry concept of enclosing sphere condition below, which was also used in Urbas’s work on studying prescribed Gauss curvature problem [31, 32]. Such geometry condition on the domain is much weaker than C2C^{2} regularity condition.

Definition 1.4.

We say that Ω\Omega satisfies an enclosing sphere condition at ξΩ\xi\in\partial\Omega, if there exists a ball B=Br(ξ)(y(ξ))ΩB=B_{r(\xi)}(y(\xi))\supset\Omega satisfying ξΩB\xi\in\partial\Omega\cap\partial B. We say that Ω\Omega satisfies an enclosing sphere condition, if Ω\Omega satisfies an enclosing sphere condition at each ξΩ\xi\in\partial\Omega. Moreover, we say that Ω\Omega satisfies a uniform enclosing sphere condition, if r(ξ)r(\xi) is bounded on Ω\partial\Omega.

Suppose that the semi-convexity with respect to the boundary and a uniform enclosing sphere conditions hold. Our second main result is a necessary and sufficient condition to the existence of solution of (1.4) in terms of the asymptotic behavior near infinity.

Theorem 1.5.

Let Ω\Omega be a domain of n\mathbb{R}^{n} satisfying a uniform enclosing sphere condition, n3n\geq 3, ΩC1\partial\Omega\in C^{1}. Let φ\varphi be semi-convex with respect to Ω\partial\Omega. Then for any A𝒜A\in\mathcal{A} and bnb\in\mathbb{R}^{n}, there exists some constant cc_{*}, such that (1.4) has a viscosity solution in C0(nΩ)C^{0}(\mathbb{R}^{n}\setminus\Omega) if and only if ccc\geq c_{*}, where cc_{*} depends only on nn, AA, bb, Ω\Omega and φ\varphi.

If Ω\Omega is a bounded and strictly convex domain with ΩC2\partial\Omega\in C^{2}, then Ω\Omega satisfies a uniform enclosing sphere condition (see Proposition A in Appendix). The strict convexity of a C2C^{2} domain Ω\Omega throughout the paper refers to Caffarelli–Li [6]. Namely, principal curvatures of Ω\partial\Omega are positive. Also, if φC2(Ω)\varphi\in C^{2}(\partial\Omega), then direct calculation shows that φ\varphi is semi-convex with respect to Ω\partial\Omega. Therefore, Theorem 1.5 is more general than Theorem 1.1 of Caffarelli–Li [6] and Li–Lu [22].

By combining Theorem 1.3, we conclude the following necessary and sufficient condition to the existence of solution of (1.4) in terms of the boundary value.

Theorem 1.6.

Let Ω\Omega be a bounded, strictly convex domain of n\mathbb{R}^{n}, n3n\geq 3, ΩC3\partial\Omega\in C^{3}. Then for any A𝒜A\in\mathcal{A}, bnb\in\mathbb{R}^{n}, there exists a constant cc such that (1.4) has a viscosity solution in C0(nΩ)C^{0}(\mathbb{R}^{n}\setminus\Omega) if and only if φ\varphi is semi-convex with respect to Ω\partial\Omega.

We see from the above theorem that for the Monge–Ampère equation, the semi-convex boundary value in the exterior Dirichlet problem is in the same position as the continuous boundary value in the interior Dirichlet problem.

We turn to point out that there exist many functions which satisfy the semi-convexity condition but φC2(Ω)\varphi\notin C^{2}(\partial\Omega); see Example 1 in Appendix. Also, there exist many domains Ω\Omega which satisfy a uniform enclosing sphere condition and make Theorem 1.5 work but even not C1C^{1}; see Example 3. It would be interesting to see if Theorem 1.5 remains valid under weaker assumptions on Ω\partial\Omega. In addition, we mention that the “uniform” in a uniform enclosing sphere condition could not be dropped; see Example 2.

The constant cc_{*} in Theorem 1.5 and Theorem 1.1 depends on different quantities of Ω\Omega and φ\varphi. Precisely, the former depends on C1C^{1} regularity of Ω\partial\Omega, the uniform radius of enclosing sphere of Ω\Omega and the semi-convexity of φ\varphi, while the latter depends on the diameter and the strict convexity of Ω\Omega, the C2C^{2} norm of Ω\partial\Omega and φC2(Ω)\|\varphi\|_{C^{2}(\partial\Omega)}.

We now comment the proof of Theorem 1.5. The proof of existence part is based on Perron’s method, and processes the equation, the boundary value and the asymptotic behavior in (1.4) separately. Along the lines of Perron’s method for Laplace equation, our proof is traditional and elementary. We first introduce the lifting function with respect to the Monge–Ampère equation and prove that Perron’s solution satisfies the Monge–Ampère equation in the viscosity sense. We then verify that such solution also satisfies the asymptotic behavior at infinity when cc is sufficiently large. It is worth to note that Perron’s solution satisfies the equation and the asymptotic behavior under fairly weak assumptions on domains and boundary values. To handle the interior boundary behavior, compared with [6], key changes need to be made due to the weaker conditions by reproving the barrier lemma (see Lemma 5.1). Once the existence part is established, we can follow almost without change as in [22, Theorem 1.2] to obtain the nonexistence part.

The rest of this paper is organized as follows. In Section 2, we establish Theorem 1.3 using an adapted smooth approximation argument. In Section 3, we prove that the Perron’s solution is a viscosity solution of the Monge–Ampère equation. In Section 4, we further check the Perron’s solution satisfies the asymptotic quadratic behavior at infinity. In Section 5, we investigate the boundary behavior of the Perron’s solution. Section 6 is devoted to prove Theorem 1.5.

We fix some notations throughout this paper. For x=(x1,,xn)nx=(x_{1},\cdots,x_{n})\in\mathbb{R}^{n}, we denote x=(x1,,xn1)n1x^{\prime}=(x_{1},\cdots,x_{n-1})\in\mathbb{R}^{n-1}. For r>0r>0, we denote by Br(x)B_{r}(x) the open ball in n\mathbb{R}^{n} centered at xx of radius rr, and by Br(x)B^{\prime}_{r}(x^{\prime}) the open ball in n1\mathbb{R}^{n-1} centered at xx^{\prime} of radius rr. For a function u(x)u(x), we denote by uiu_{i} the partial derivative with respect to xix_{i}, and by uiju_{ij} the second derivative with respect to xix_{i} and xjx_{j}. We also use DuDu and D2uD^{2}u to denote (u1,,un)(u_{1},\cdots,u_{n}) and the n×nn\times n matrix (uij)(u_{ij}), respectively. Similarly, for a function v(x)v(x^{\prime}), we used DvD^{\prime}v and D2vD^{\prime 2}v to denote (v1,,vn1)(v_{1},\cdots,v_{n-1}) and the (n1)×(n1)(n-1)\times(n-1) matrix (vij)(v_{ij}), respectively.

2. Proof of Theorem 1.3

For the reader’s convenience, we recall that uC0(D)u\in C^{0}(D) is locally convex in an open set DnD\subset\mathbb{R}^{n} if for any xDx\in D, there is an open ball BDB\subset D centered at xx such that the restriction of uu to BB is convex, which is equivalent to that uu is convex in any open ball BDB\subset D (see for instance [34]).

In this section, we derive the necessity of semi-convexity condition of the boundary value in a very general setting through an adapted smooth approximation argument, which implies Theorem 1.3.

Proposition 2.1.

Let Ω\Omega be a bounded convex domain of n\mathbb{R}^{n}, n3n\geq 3, ΩC3\partial\Omega\in C^{3}. Let uC0(nΩ)u\in C^{0}(\mathbb{R}^{n}\setminus\Omega) be locally convex, and φ(x):=u(x)\varphi(x):=u(x) for xΩx\in\partial\Omega. Then φ\varphi is semi-convex with respect to Ω\partial\Omega.

Proof.

For ε>0\varepsilon>0, denote Ωε={xn|dist(x,Ω)<ε}\Omega^{\varepsilon}=\{x\in\mathbb{R}^{n}|\,\text{dist}(x,\Omega)<\varepsilon\}. Define for 0<ε<10<\varepsilon<1,

uε(x)=B1(0)η(z)u(xεz)dzforxnΩε¯,u^{\varepsilon}(x)=\int_{B_{1}(0)}\eta(z)u(x-\varepsilon z)\mathrm{d}z\quad\text{for}~{}x\in\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}},

where ηC(n)\eta\in C^{\infty}(\mathbb{R}^{n}) is the standard mollifier given by

η(x)={C0exp(1|x|21)if|x|<1,0if|x|1,\eta(x)=\begin{cases}C_{0}\exp{\Big{(}\frac{1}{|x|^{2}-1}\Big{)}}&\quad\text{if}~{}|x|<1,\\ 0&\quad\text{if}~{}|x|\geq 1,\end{cases}

with the constant C0C_{0} satisfying nη(x)dx=1\int_{\mathbb{R}^{n}}\eta(x)\mathrm{d}x=1. Clearly, uεC(nΩε¯)u^{\varepsilon}\in C^{\infty}(\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}}). It can be verified that uεu^{\varepsilon} is locally convex in nΩε¯\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}}. Indeed, given a ball Br(x0)nΩε¯B_{r}(x_{0})\subset\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}}, x,yBr(x0)x,y\in B_{r}(x_{0}) and zB1(0)z\in B_{1}(0), we have

xεz,yεzBr+ε(x0)nΩ¯.x-\varepsilon z,~{}y-\varepsilon z\in B_{r+\varepsilon}(x_{0})\subset\mathbb{R}^{n}\setminus\overline{\Omega}.

It follows from the convexity of uu in Br+ε(x0)B_{r+\varepsilon}(x_{0}) and η0\eta\geq 0 that for 0<t<10<t<1,

uε(tx+(1t)y)=B1(0)η(z)u(tx+(1t)yεz)dz=B1(0)η(z)u(t(xεz)+(1t)(yεz))dztB1(0)η(z)u(xεz)dz+(1t)B1(0)η(z)u(yεz)dz=tuε(x)+(1t)uε(y).\begin{split}u^{\varepsilon}(tx+(1-t)y)&=\int_{B_{1}(0)}\eta(z)u(tx+(1-t)y-\varepsilon z)\mathrm{d}z\\ &=\int_{B_{1}(0)}\eta(z)u(t(x-\varepsilon z)+(1-t)(y-\varepsilon z))\mathrm{d}z\\ &\leq t\int_{B_{1}(0)}\eta(z)u(x-\varepsilon z)\mathrm{d}z+(1-t)\int_{B_{1}(0)}\eta(z)u(y-\varepsilon z)\mathrm{d}z\\ &=tu^{\varepsilon}(x)+(1-t)u^{\varepsilon}(y).\end{split}

Thus we have proved uεu^{\varepsilon} is locally convex in nΩε¯\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}}.

Fix ξΩ\xi\in\partial\Omega and assume Ω\partial\Omega can be locally represented by the graph of

xn=ρ(x)0for|x|<δ(ξ),x_{n}=\rho(x^{\prime})\geq 0\quad\text{for}\,|x^{\prime}|<\delta(\xi),

for some δ(ξ)>0\delta(\xi)>0. Denote ν(x)=(ν(x),ν(n)(x))\nu(x)=(\nu^{\prime}(x),\nu^{(n)}(x)) the unit inner normal vector of Ω\partial\Omega at x=(x,ρ(x))x=(x^{\prime},\rho(x^{\prime})). Since Ω\Omega is convex, we have for x=(x,ρ(x))x=(x^{\prime},\rho(x^{\prime})),

x2εν(x)=(x2εν(x,ρ(x)),ρ(x)2εν(n)(x,ρ(x)))Ω2ε.x-2\varepsilon\nu(x)=(x^{\prime}-2\varepsilon\nu^{\prime}(x^{\prime},\rho(x^{\prime})),\rho(x^{\prime})-2\varepsilon\nu^{(n)}(x^{\prime},\rho(x^{\prime})))\in\partial\Omega^{2\varepsilon}.

Define

ψε(x)=uε(x2εμ(x),ρ(x)2εμ(n)(x))for|x|<δ(ξ),\psi^{\varepsilon}(x^{\prime})=u^{\varepsilon}(x^{\prime}-2\varepsilon\mu^{\prime}(x^{\prime}),\rho(x^{\prime})-2\varepsilon\mu^{(n)}(x^{\prime}))\quad\text{for}\,|x^{\prime}|<\delta(\xi),

where μ(x)=(μ(1)(x),,μ(n1)(x))=ν(x,ρ(x))\mu^{\prime}(x^{\prime})=(\mu^{(1)}(x^{\prime}),\cdots,\mu^{(n-1)}(x^{\prime}))=\nu^{\prime}(x^{\prime},\rho(x^{\prime})), μ(n)(x)=ν(n)(x,ρ(x))\mu^{(n)}(x^{\prime})=\nu^{(n)}(x^{\prime},\rho(x^{\prime})). Direct calculation shows for i,j=1,,n1i,j=1,\cdots,n-1,

ψiε=k=1n1ukε(δik2εμi(k))+unε(ρi2εμi(n)),\psi^{\varepsilon}_{i}=\sum_{k=1}^{n-1}u^{\varepsilon}_{k}\left(\delta_{ik}-2\varepsilon\mu^{(k)}_{i}\right)+u^{\varepsilon}_{n}\left(\rho_{i}-2\varepsilon\mu^{(n)}_{i}\right),

and

ψijε=k=1n1((l=1n1uklε(δjl2εμj(l))+uknε(ρj2εμj(n)))(δik2εμi(k))+ukε(2εμij(k)))+(k=1n1unkε(δkj2εμj(k))+unnε(ρj2εμj(n)))(ρi2εμi(n))+unε(ρij2εμij(n))=k,l=1n1uklε(δik2εμi(k))(δjl2εμj(l))+k=1n1uknε(δik2εμi(k))(ρj2εμj(n))+k=1n1unkε(δkj2εμj(k))(ρi2εμi(n))+unnε(ρi2εμi(n))(ρj2εμj(n))+unερij2ε(k=1n1ukεμij(k)+unεμij(n)),\begin{split}\psi^{\varepsilon}_{ij}=&\,\sum_{k=1}^{n-1}\left(\left(\sum_{l=1}^{n-1}u^{\varepsilon}_{kl}\left(\delta_{jl}-2\varepsilon\mu^{(l)}_{j}\right)+u^{\varepsilon}_{kn}\left(\rho_{j}-2\varepsilon\mu^{(n)}_{j}\right)\right)\left(\delta_{ik}-2\varepsilon\mu^{(k)}_{i}\right)+u^{\varepsilon}_{k}\left(-2\varepsilon\mu^{(k)}_{ij}\right)\right)\\ &\,+\left(\sum_{k=1}^{n-1}u^{\varepsilon}_{nk}\left(\delta_{kj}-2\varepsilon\mu^{(k)}_{j}\right)+u^{\varepsilon}_{nn}\left(\rho_{j}-2\varepsilon\mu^{(n)}_{j}\right)\right)\left(\rho_{i}-2\varepsilon\mu^{(n)}_{i}\right)+u^{\varepsilon}_{n}\left(\rho_{ij}-2\varepsilon\mu^{(n)}_{ij}\right)\\ =&\,\sum_{k,l=1}^{n-1}u^{\varepsilon}_{kl}\left(\delta_{ik}-2\varepsilon\mu^{(k)}_{i}\right)\left(\delta_{jl}-2\varepsilon\mu^{(l)}_{j}\right)\\ &\,+\sum_{k=1}^{n-1}u^{\varepsilon}_{kn}\left(\delta_{ik}-2\varepsilon\mu^{(k)}_{i}\right)\left(\rho_{j}-2\varepsilon\mu^{(n)}_{j}\right)+\sum_{k=1}^{n-1}u^{\varepsilon}_{nk}\left(\delta_{kj}-2\varepsilon\mu^{(k)}_{j}\right)\left(\rho_{i}-2\varepsilon\mu^{(n)}_{i}\right)\\ &\,+u^{\varepsilon}_{nn}\left(\rho_{i}-2\varepsilon\mu^{(n)}_{i}\right)\left(\rho_{j}-2\varepsilon\mu^{(n)}_{j}\right)+u^{\varepsilon}_{n}\rho_{ij}-2\varepsilon\Bigg{(}\sum_{k=1}^{n-1}u^{\varepsilon}_{k}\mu^{(k)}_{ij}+u^{\varepsilon}_{n}\mu^{(n)}_{ij}\Bigg{)},\end{split}

where δij=1\delta_{ij}=1 if i=ji=j and δij=0\delta_{ij}=0 if iji\neq j. For ζn1\zeta\in\mathbb{R}^{n-1}, it follows that

i,j=1n1ψijεζiζj=k,l=1n1uklε(i=1n1(δik2εμi(k))ζi)(j=1n1(δjl2εμj(l))ζj)+2k=1n1uknε(i=1n1(δik2εμi(k))ζi)(j=1n1(ρj2εμi(n))ζj)+unnε(i=1n1(ρi2εμi(n))ζi)(j=1n1(ρj2εμj(n))ζj)+unεi,j=1n1ρijζiζj2ε(k=1n1ukεi,j=1n1μij(k)ζiζj+unεi,j=1n1μij(n)ζiζj).\begin{split}\sum_{i,j=1}^{n-1}\psi^{\varepsilon}_{ij}\zeta_{i}\zeta_{j}=&\,\sum_{k,l=1}^{n-1}u^{\varepsilon}_{kl}\Bigg{(}\sum_{i=1}^{n-1}\left(\delta_{ik}-2\varepsilon\mu^{(k)}_{i}\right)\zeta_{i}\Bigg{)}\Bigg{(}\sum_{j=1}^{n-1}\left(\delta_{jl}-2\varepsilon\mu^{(l)}_{j}\right)\zeta_{j}\Bigg{)}\\ &\,+2\sum_{k=1}^{n-1}u^{\varepsilon}_{kn}\Bigg{(}\sum_{i=1}^{n-1}\left(\delta_{ik}-2\varepsilon\mu^{(k)}_{i}\right)\zeta_{i}\Bigg{)}\Bigg{(}\sum_{j=1}^{n-1}\left(\rho_{j}-2\varepsilon\mu^{(n)}_{i}\right)\zeta_{j}\Bigg{)}\\ &\,+u^{\varepsilon}_{nn}\Bigg{(}\sum_{i=1}^{n-1}\left(\rho_{i}-2\varepsilon\mu^{(n)}_{i}\right)\zeta_{i}\Bigg{)}\Bigg{(}\sum_{j=1}^{n-1}\left(\rho_{j}-2\varepsilon\mu^{(n)}_{j}\right)\zeta_{j}\Bigg{)}\\ &\,+u^{\varepsilon}_{n}\sum_{i,j=1}^{n-1}\rho_{ij}\zeta_{i}\zeta_{j}-2\varepsilon\Bigg{(}\sum_{k=1}^{n-1}u^{\varepsilon}_{k}\sum_{i,j=1}^{n-1}\mu^{(k)}_{ij}\zeta_{i}\zeta_{j}+u^{\varepsilon}_{n}\sum_{i,j=1}^{n-1}\mu^{(n)}_{ij}\zeta_{i}\zeta_{j}\Bigg{)}.\end{split}

Since D2uε0D^{2}u^{\varepsilon}\geq 0, D2ρ0D^{\prime 2}\rho\geq 0 and (μij(1)),,(μij(n))(\mu^{(1)}_{ij}),\cdots,(\mu^{(n)}_{ij}) are bounded due to ΩC3\partial\Omega\in C^{3},

i,j=1n1ψijεζiζjunεi,j=1n1ρijζiζj2ε(k=1n1ukεi,j=1n1μij(k)ζiζj+unεi,j=1n1μij(n)ζiζj)C(min{minΓ2εunε,0}2maxΓ2εε|Duε|)|ζ|2,\begin{split}\sum_{i,j=1}^{n-1}\psi^{\varepsilon}_{ij}\zeta_{i}\zeta_{j}&\,\geq u^{\varepsilon}_{n}\sum_{i,j=1}^{n-1}\rho_{ij}\zeta_{i}\zeta_{j}-2\varepsilon\Bigg{(}\sum_{k=1}^{n-1}u^{\varepsilon}_{k}\sum_{i,j=1}^{n-1}\mu^{(k)}_{ij}\zeta_{i}\zeta_{j}+u^{\varepsilon}_{n}\sum_{i,j=1}^{n-1}\mu^{(n)}_{ij}\zeta_{i}\zeta_{j}\Bigg{)}\\ &\,\geq C\left(\min\{\min_{\Gamma^{2\varepsilon}}u^{\varepsilon}_{n},0\}-2\max_{\Gamma^{2\varepsilon}}\varepsilon|Du^{\varepsilon}|\right)|\zeta|^{2},\end{split} (2.1)

where C>0C>0 is a constant depending only on nn and the C3C^{3} norm of Ω\partial\Omega, and

Γ2ε={x2εv(x)|x=(x,ρ(x)),|x|δ(ξ)}Ω2ε.\Gamma^{2\varepsilon}=\{x-2\varepsilon v(x)|\,x=(x^{\prime},\rho(x^{\prime})),~{}|x^{\prime}|\leq\delta(\xi)\}\subset\partial\Omega^{2\varepsilon}.

To prove that ψε\psi^{\varepsilon} is semi-convex in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}), we first estimate unεu_{n}^{\varepsilon} on Γ2ε\Gamma^{2\varepsilon} from below. Take R>0R>0 such that Ω3BR(0)\Omega^{3}\subset B_{R}(0). For x0Γ2εx_{0}\in\Gamma^{2\varepsilon}, we denote by y0y_{0} the intersection of ll and BR(0)\partial B_{R}(0), where ll is the ray along the direction of negative xnx_{n}-axis with the endpoint x0x_{0}. Then the open line segment x0y0¯\overline{x_{0}y_{0}} connecting x0x_{0} and y0y_{0} is contained in nΩε¯\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}}. Hence for

x^=|y0x^||y0x0|x0+|x^x0|y0x0y0x0y0¯,\hat{x}=\frac{|y_{0}-\hat{x}|}{|y_{0}-x_{0}|}x_{0}+\frac{|\hat{x}-x_{0}|}{y_{0}-x_{0}}y_{0}\in\overline{x_{0}y_{0}},

we have by the convexity of uεu^{\varepsilon},

uε(x^)|y0x^||y0x0|uε(x0)+|x^x0||y0x0|uε(y0).u^{\varepsilon}(\hat{x})\leq\frac{|y_{0}-\hat{x}|}{|y_{0}-x_{0}|}u^{\varepsilon}(x_{0})+\frac{|\hat{x}-x_{0}|}{|y_{0}-x_{0}|}u^{\varepsilon}(y_{0}).

Together with |y0x0|1|y_{0}-x_{0}|\geq 1, we get

uε(x^)uε(x0)|x^x0|uε(y0)uε(x0)|y0x0|2maxBR(0)Ω2ε¯|uε|.\frac{u^{\varepsilon}(\hat{x})-u^{\varepsilon}(x_{0})}{|\hat{x}-x_{0}|}\leq\frac{u^{\varepsilon}(y_{0})-u^{\varepsilon}(x_{0})}{|y_{0}-x_{0}|}\leq 2\max_{\overline{B_{R}(0)\setminus\Omega^{2\varepsilon}}}|u^{\varepsilon}|.

The boundedness of uu on BR+1(0)Ω¯\overline{B_{R+1}(0)\setminus\Omega} yields that

|uε(x)|B1(0)η(z)|u(xεz)|dzmaxzB1(0)¯|u(xεz)||u^{\varepsilon}(x)|\leq\int_{B_{1}(0)}\eta(z)|u(x-\varepsilon z)|\mathrm{d}z\leq\max_{z\in\overline{B_{1}(0)}}|u(x-\varepsilon z)|

holds uniformly for xBR(0)Ω2ε¯x\in\overline{B_{R}(0)\setminus\Omega^{2\varepsilon}}. Then

maxBR(0)Ω2ε¯|uε|maxBR+1(0)Ω¯|u|.\max_{\overline{B_{R}(0)\setminus\Omega^{2\varepsilon}}}|u^{\varepsilon}|\leq\max_{\overline{B_{R+1}(0)\setminus\Omega}}|u|.

Hence,

uε(x^)uε(x0)|x^x0|2maxBR+1(0)Ω¯|u|.\frac{u^{\varepsilon}(\hat{x})-u^{\varepsilon}(x_{0})}{|\hat{x}-x_{0}|}\leq 2\max_{\overline{B_{R+1}(0)\setminus\Omega}}|u|.

That is,

minΩ2εunε2maxBR+1(0)Ω¯|u|.\min_{\partial\Omega^{2\varepsilon}}u^{\varepsilon}_{n}\geq-2\max_{\overline{B_{R+1}(0)\setminus\Omega}}|u|. (2.2)

We continue to estimate ε|Duε|\varepsilon|Du^{\varepsilon}| on Ω2ε\partial\Omega^{2\varepsilon} from above. It is clear that

uε(x)=nΩ¯1εnη(xzε)u(z)dzforxnΩε¯.u^{\varepsilon}(x)=\int_{\mathbb{R}^{n}\setminus\overline{\Omega}}\frac{1}{\varepsilon^{n}}\eta\bigg{(}\frac{x-z}{\varepsilon}\bigg{)}u(z)\mathrm{d}z\quad\text{for}~{}x\in\mathbb{R}^{n}\setminus\overline{\Omega^{\varepsilon}}.

Then direct calculation gives

|uiε(x)|=|nΩ¯1εnηxi(xzε)u(z)dz|=1ε|B1(0)ηi(z)u(xεz)dz|1εmaxzB1(0)¯|u(xεz)|B1(0)|Dη(z)|dz=:C(n)εmaxzB1(0)¯|u(xεz)|.\begin{split}|u^{\varepsilon}_{i}(x)|&=\bigg{|}\int_{\mathbb{R}^{n}\setminus\overline{\Omega}}\frac{1}{\varepsilon^{n}}\frac{\partial\eta}{\partial x_{i}}\bigg{(}\frac{x-z}{\varepsilon}\bigg{)}u(z)\mathrm{d}z\bigg{|}\\ &=\frac{1}{\varepsilon}\bigg{|}\int_{B_{1}(0)}\eta_{i}(z)u(x-\varepsilon z)\mathrm{d}z\bigg{|}\\ &\leq\frac{1}{\varepsilon}\max_{z\in\overline{B_{1}(0)}}|u(x-\varepsilon z)|\int_{B_{1}(0)}|D\eta(z)|\mathrm{d}z\\ &=:\frac{C(n)}{\varepsilon}\max_{z\in\overline{B_{1}(0)}}|u(x-\varepsilon z)|.\end{split}

Using the boundedness of uu in Ω3Ω¯\overline{\Omega^{3}\setminus\Omega}, we obtain for x0Ω2εx_{0}\in\partial\Omega^{2\varepsilon},

|uiε(x0)|C(n)εmaxΩ3Ω¯|u|.|u^{\varepsilon}_{i}(x_{0})|\leq\frac{C(n)}{\varepsilon}\max_{\overline{\Omega^{3}\setminus\Omega}}|u|.

That is,

maxΩ2εε|Duε|C(n)maxΩ3Ω¯|u|.\max_{\partial\Omega^{2\varepsilon}}\varepsilon|Du^{\varepsilon}|\leq C(n)\max_{\overline{\Omega^{3}\setminus\Omega}}|u|. (2.3)

Combining (2.1) with (2.2) and (2.3), we obtain that there exists a constant K(ξ)>0K(\xi)>0 independent of ε\varepsilon such that

D2ψε(x)KIn1forxBδ(ξ)(0),D^{2}\psi^{\varepsilon}(x^{\prime})\geq-KI_{n-1}\quad\text{for}~{}x^{\prime}\in B^{\prime}_{\delta(\xi)}(0^{\prime}),

where In1I_{n-1} is the (n1)×(n1)(n-1)\times(n-1) identity matrix. That is ψε(x)+K(ξ)2|x|2\psi^{\varepsilon}(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2} is convex in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}). Then for x,yBδ(ξ)(0)x^{\prime},y^{\prime}\in B^{\prime}_{\delta(\xi)}(0^{\prime}) and 0<t<10<t<1,

ψε(tx+(1t)y)+K(ξ)2|tx+(1t)y|2t(ψε(x)+K(ξ)2|x|2)+(1t)(ψε(y)+K(ξ)2|y|2).\begin{split}&\,\psi^{\varepsilon}(tx^{\prime}+(1-t)y^{\prime})+\frac{K(\xi)}{2}|tx^{\prime}+(1-t)y^{\prime}|^{2}\\ \leq&\,t\bigg{(}\psi^{\varepsilon}(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\psi^{\varepsilon}(y^{\prime})+\frac{K(\xi)}{2}|y^{\prime}|^{2}\bigg{)}.\end{split} (2.4)

From the uniform continuity of uu in Ω3Ω¯\overline{\Omega^{3}\setminus\Omega}, it follows that for ψ(x):=φ(x,ρ(x))\psi(x^{\prime}):=\varphi(x^{\prime},\rho(x^{\prime})),

|ψε(x)ψ(x)|=|uε(x2εν(x))u(x)|B1(0)η(z)|u(x2εν(x)εz)u(x)|dzmaxzB1(0)¯|u(x2εν(x)εz)u(x)| 0asε0\begin{split}|\psi^{\varepsilon}(x^{\prime})-\psi(x^{\prime})|=&\,|u^{\varepsilon}(x-2\varepsilon\nu(x))-u(x)|\\ \leq&\,\int_{B_{1}(0)}\eta(z)\left|u(x-2\varepsilon\nu(x)-\varepsilon z)-u(x)\right|\mathrm{d}z\\ \leq&\,\max_{z\in\overline{B_{1}(0)}}\left|u(x-2\varepsilon\nu(x)-\varepsilon z)-u(x)\right|\\ \to&\,0\quad\quad\text{as}~{}\varepsilon\to 0\end{split}

holds uniformly for x=(x,ρ(x))x=(x^{\prime},\rho(x^{\prime})) with |x|<δ(ξ)|x^{\prime}|<\delta(\xi). By sending ε0\varepsilon\to 0 in (2.4), we obtain for x,yBδ(ξ)(0)x^{\prime},y^{\prime}\in B^{\prime}_{\delta(\xi)}(0^{\prime}) and 0<t<10<t<1,

ψ(tx+(1t)y)+K(ξ)2|tx+(1t)y|2t(ψ(x)+K(ξ)2|x|2)+(1t)(ψ(y)+K(ξ)2|y|2).\begin{split}&\,\psi(tx^{\prime}+(1-t)y^{\prime})+\frac{K(\xi)}{2}|tx^{\prime}+(1-t)y^{\prime}|^{2}\\ \leq&\,t\bigg{(}\psi(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\psi(y^{\prime})+\frac{K(\xi)}{2}|y^{\prime}|^{2}\bigg{)}.\end{split}

Hence, ψ\psi is semi-convex in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}). Since ξΩ\xi\in\partial\Omega is arbitrary, the proof is completed. ∎

3. Perron’s solution of equation (1.3) in the exterior domain

In this section, we give the expression of the Perron’s solution of the exterior Dirichlet problem (1.4), under fairly general assumptions that the domain Ω\Omega and the boundary value φ\varphi are bounded. We will verify that the Perron’s solution is indeed a viscosity solution of equation (1.3) in the exterior domain nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega}. The proof is similar in spirit to that of Laplace equation.

For the reader’s convenience, we recall the definition of viscosity solutions. Let DD be an open set in n\mathbb{R}^{n} and let uC0(D)u\in C^{0}(D) be a locally convex function. We say that uu is a viscosity subsolution of (1.3) in DD, if for any function vC2(D)v\in C^{2}(D) and any local maximum point x0Dx_{0}\in D of uvu-v, we have

det(D2v(x0))1.\det(D^{2}v(x_{0}))\geq 1.

We say that uu is a viscosity supersolution of (1.3) in DD, if for any local convex function vC2(D)v\in C^{2}(D) and any local minimum point x0Dx_{0}\in D of uvu-v, we have

det(D2v(x0))1.\det(D^{2}v(x_{0}))\leq 1.

uu is a viscosity solution of (1.3), if it is both a viscosity subsolution and a viscosity supersolution of (1.3). It is clear from the definition that if uu is a viscosity solution of (1.3) in each ball BDB\subset\subset D, then uu is a viscosity solution of (1.3) in DD.

To begin with, we introduce the definition of the lifting function with respect to the Monge–Ampère equation.

Definition 3.1.

Let BDB\subset\subset D be an open ball, and uC0(D)u\in C^{0}(D). We define the lifting function of uu with respect to the Monge–Ampère equation (1.3) in BB by

U(x)={u¯(x)ifxB,u(x)ifxDB,U(x)=\begin{cases}\overline{u}(x)&\quad\text{if}~{}x\in B,\\ u(x)&\quad\text{if}~{}x\in D\setminus B,\end{cases}

where u¯C0(B¯)\overline{u}\in C^{0}(\overline{B}) is convex in BB solving

{det(D2u¯)=1in B,u¯=uon B.\begin{cases}\det(D^{2}\overline{u})=1&\text{in }B,\\ \overline{u}=u&\text{on }\partial B.\end{cases}

The existence of u¯\overline{u} is guaranteed by [29, Theorem 4.1] (see also [6, Lemma A.3] for a direct proof).

Proposition 3.2.

Let uu be a viscosity subsolution of (1.3) in DD, and let UU be the lifting function of uu with respect to (1.3) in BB. Then UuU\geq u in DD, and UU is also a viscosity subsolution of (1.3) in DD.

Proof.

The comparison principle gives u¯u\overline{u}\geq u in BB, and so UuU\geq u in DD.

We claim that UU is locally convex in DD. Indeed, it suffices to prove that UU is convex in an open ball B0DB_{0}\subset D centered at x0Bx_{0}\in\partial B. For any fixed x,yB0x,y\in B_{0}, we only need to consider the case of xB0Bx\in B_{0}\cap B and yB0By\in B_{0}\setminus B, as otherwise the conclusion follows immediately.

Case 1. For any 0<t<10<t<1 such that tx+(1t)yB0Btx+(1-t)y\in B_{0}\setminus B, by the convexity of uu in B0B_{0}, we have

U(tx+(1t)y)=u(tx+(1t)y)tu(x)+(1t)u(y)tU(x)+(1t)U(y).\begin{split}U(tx+(1-t)y)&=u(tx+(1-t)y)\\ &\leq tu(x)+(1-t)u(y)\\ &\leq tU(x)+(1-t)U(y).\end{split}

Case 2. For any 0<t<10<t<1 such that z:=tx+(1t)yB0Bz:=tx+(1-t)y\in B_{0}\cap B, let z^=LxyB\hat{z}=L_{xy}\cap\partial B, where LxyL_{xy} is the line segment connecting xx and yy. Then z=t^x+(1t^)z^z=\hat{t}x+(1-\hat{t})\hat{z} for some 0<t^<10<\hat{t}<1. By |xz^||zz^||x-\hat{z}|\geq|z-\hat{z}|, we have

t=|zy||xy|=|zz^|+|z^y||xz^|+|z^y||zz^||xz^|=t^.t=\frac{|z-y|}{|x-y|}=\frac{|z-\hat{z}|+|\hat{z}-y|}{|x-\hat{z}|+|\hat{z}-y|}\geq\frac{|z-\hat{z}|}{|x-\hat{z}|}=\hat{t}.

Since u¯\overline{u} is convex in BB and u¯u\overline{u}\geq u in BB, it follows that

U(tx+(1t)y)=u¯(z)t^u¯(x)+(1t^)u¯(z^)=t^u¯(x)+(1t^)u(tt^1t^x+1t1t^y)t^u¯(x)+(tt^)u(x)+(1t)u(y)tu¯(x)+(1t)u(y)tU(x)+(1t)U(y).\begin{split}U(tx+(1-t)y)&=\overline{u}(z)\\ &\leq\hat{t}\overline{u}(x)+(1-\hat{t})\overline{u}(\hat{z})\\ &=\hat{t}\overline{u}(x)+(1-\hat{t})u\bigg{(}\frac{t-\hat{t}}{1-\hat{t}}x+\frac{1-t}{1-\hat{t}}y\bigg{)}\\ &\leq\hat{t}\overline{u}(x)+(t-\hat{t})u(x)+(1-t)u(y)\\ &\leq t\overline{u}(x)+(1-t)u(y)\\ &\leq tU(x)+(1-t)U(y).\end{split}

By Lemma 3.3 below, UU is a viscosity subsolution of (1.3) in DD, and the proof is finished. ∎

We collect some preliminary lemmas for the viscosity solutions, which will be used in the proof of this section.

Lemma 3.3.

Let DD and D1D_{1} are two domains such that DD1D\subset\subset D_{1} and DD is bounded. Suppose that uC(D1)u\in C(D_{1}) is a viscosity subsolution of (1.3) in D1D_{1}, and vC(D¯)v\in C(\overline{D}) is a viscosity subsolution of (1.3) in DD. Assume that vuv\leq u on D\partial D. If

w={max{u,v}in D¯,uin D1D,w=\begin{cases}\max\{u,v\}&\mbox{in }\overline{D},\\ u&\mbox{in }D_{1}\setminus D,\\ \end{cases}

is locally convex in D1D_{1}, then ww is a viscosity subsolution of (1.3) in D1D_{1}.

Proof.

This lemma can be deduced by following the arguments in the proof of [5, Proposition 2.8], where the property is proved for viscosity subsolutions of uniformly elliptic fully nonlinear equations. ∎

Lemma 3.4.

Let 𝒮\mathcal{S} be a nonempty family of viscosity subsolution of (1.3) in DD. Define

u(x)=sup{v(x)|v𝒮}forxD.u(x)=\sup\{v(x)|~{}v\in\mathcal{S}\}\quad\text{for}~{}x\in D.

If u(x)<u(x)<\infty for xDx\in D. Then uu is a viscosity subsolution of (1.3) in DD.

Proof.

Since v𝒮v\in\mathcal{S} is locally convex in DD, uC0(D)u\in C^{0}(D) is locally convex. The rest of the proof follows from that of [11, Lemma 4.2]. ∎

Lemma 3.5.

Assume that {uk}\{u_{k}\} is a family of viscosity solution of (1.3) in DD and ukuu_{k}\to u locally uniformly in DD as kk\to\infty. Then uu is a viscosity solution of (1.3) in DD.

Proof.

This property was also mentioned by [5, Proposition 2.9] and [15, Proposition 2.1], while the proofs therein are omitted. Also, the definition of viscosity solution of (1.3) is a bit different from that of [5] and [15]. Hence we give the proof for complements.

Clearly, uu is locally convex in DD. Let vC2(D)v\in C^{2}(D) and x0Dx_{0}\in D be a local maximum point of uvu-v. Then x0x_{0} is a strictly maximum point of uvεu-v_{\varepsilon}, where vε(x):=v(x)+ε|xx0|2v_{\varepsilon}(x):=v(x)+\varepsilon|x-x_{0}|^{2}, ε>0\varepsilon>0. Namely, for some δ>0\delta>0 with Bδ(x0)¯D\overline{B_{\delta}(x_{0})}\subset D,

u(x)vε(x)<u(x0)vε(x0),0<|xx0|δ.u(x)-v_{\varepsilon}(x)<u(x_{0})-v_{\varepsilon}(x_{0}),\quad\forall~{}0<|x-x_{0}|\leq\delta.

Let xkx_{k} be the maximum point of ukvεu_{k}-v_{\varepsilon} in Bδ(x0)¯D\overline{B_{\delta}(x_{0})}\subset D. We claim that xkx0x_{k}\to x_{0} as kk\to\infty. Indeed, we need to prove that for kk sufficiently large, |xkx0|1k|x_{k}-x_{0}|\leq\frac{1}{k}. For xBδ(x0)¯B1k(x0)x\in\overline{B_{\delta}(x_{0})}\setminus B_{\frac{1}{k}}(x_{0}), the locally uniform convergence of {uk}\{u_{k}\} implies that, for kk sufficiently large,

uk(x)vε(x)<u(x)vε(x)+σ<u(x0)vε(x0)σ<uk(x0)vε(x0)uk(xk)vε(xk),\begin{split}u_{k}(x)-v_{\varepsilon}(x)&<u(x)-v_{\varepsilon}(x)+\sigma<u(x_{0})-v_{\varepsilon}(x_{0})-\sigma\\ &<u_{k}(x_{0})-v_{\varepsilon}(x_{0})\leq u_{k}(x_{k})-v_{\varepsilon}(x_{k}),\end{split}

where

σ=14minBδ(x0)¯B1k(x0)(u(x0)vε(x0)(uvε))>0.\sigma=\frac{1}{4}\min_{\overline{B_{\delta}(x_{0})}\setminus B_{\frac{1}{k}}(x_{0})}(u(x_{0})-v_{\varepsilon}(x_{0})-(u-v_{\varepsilon}))>0.

This gives xkB1k(x0)¯x_{k}\in\overline{B_{\frac{1}{k}}(x_{0})}. Thus,

det(D2vε(xk))1.\det(D^{2}v_{\varepsilon}(x_{k}))\geq 1.

Letting ε0\varepsilon\to 0 and kk\to\infty, we have

det(D2v(x0))1.\det(D^{2}v(x_{0}))\geq 1.

Hence, uu is a viscosity subsolution of (1.3) in DD. Similar arguments leads to uu being a supersolution. ∎

With these preliminaries, we are now ready to deal with the Perron’s solution. For a bounded open set Ω\Omega of n\mathbb{R}^{n}, a function φ\varphi defined on Ω\partial\Omega and a constant cc, we begin with the special case of A=IA=I and b=0b=0, and denote the set of subfunctions by

𝒮cφ={vC0(nΩ)|v is a viscosity subsolution of det(D2v)=1in nΩ¯,vφ on Ω,lim sup|x|(v(x)12|x|2)c}.\begin{split}\mathcal{S}_{c}^{\varphi}=\bigg{\{}&v\in C^{0}(\mathbb{R}^{n}\setminus\Omega)|\ v\text{ is a viscosity subsolution of }\det(D^{2}v)=1~{}\text{in }\mathbb{R}^{n}\setminus\overline{\Omega},\\ &\ v\leq\varphi\text{ on }\partial\Omega,\ \limsup_{|x|\to\infty}\bigg{(}v(x)-\frac{1}{2}|x|^{2}\bigg{)}\leq c\bigg{\}}.\end{split}

One basic result is that the Perron’s solution satisfies the Monge–Ampère equation in the exterior domain in the viscosity sense.

Proposition 3.6.

Let Ω\Omega be a bounded open set of n\mathbb{R}^{n}, n3n\geq 3, and let φ\varphi be a bounded function on Ω\partial\Omega. Then for every constant csupxΩ(φ(x)12|x|2)c\geq\sup_{x\in\partial\Omega}\left(\varphi(x)-\frac{1}{2}|x|^{2}\right), the function

u(x)=sup{v(x)|v𝒮cφ}for xnΩ¯u(x)=\sup\{v(x)|\ v\in\mathcal{S}_{c}^{\varphi}\}\quad\text{for }x\in\mathbb{R}^{n}\setminus\overline{\Omega} (3.1)

is a viscosity solution of (1.3) in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega}.

Proof.

The proof will be divided into three steps.

Step 1. We first verify that uu is well defined in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega}. Namely, the supremum is meaningful. It is clear from cC0c\geq-C_{0} that

v(x):=12|x|2C0𝒮cφfor xnΩ,v^{-}(x):=\frac{1}{2}|x|^{2}-C_{0}\in\mathcal{S}_{c}^{\varphi}\quad\text{for }x\in\mathbb{R}^{n}\setminus\Omega,

where C0=supxΩ(12|x|2φ(x))C_{0}=\sup_{x\in\partial\Omega}\left(\frac{1}{2}|x|^{2}-\varphi(x)\right). Thus, 𝒮cφ\mathcal{S}_{c}^{\varphi} is nonempty. Moreover, set

v+(x)=12|x|2+cfor xnΩ.v^{+}(x)=\frac{1}{2}|x|^{2}+c\quad\text{for }x\in\mathbb{R}^{n}\setminus\Omega.

It is easily seen that v+φv^{+}\geq\varphi on Ω\partial\Omega. It follows from the comparison principle that, for each v𝒮cφv\in\mathcal{S}_{c}^{\varphi},

vv+in nΩ.v\leq v^{+}\quad\text{in }\mathbb{R}^{n}\setminus\Omega.

Hence, the supremum is well defined, and uv+u\leq v^{+} in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega}.

Step 2. We next prove that uC0(nΩ¯)u\in C^{0}(\mathbb{R}^{n}\setminus\overline{\Omega}) is a viscosity subsolution of (1.3). Indeed, this follows from the definition of uu and Lemma 3.4 immediately.

Step 3. We proceed to prove that uu is a viscosity solution of (1.3). It suffices to prove that, for any Br(x0)nΩ¯B_{r}(x_{0})\subset\subset\mathbb{R}^{n}\setminus\overline{\Omega},

detD2u=1in Br(x0)\det{D^{2}u}=1\quad\text{in }B_{r}(x_{0})

in the viscosity sense.

Fix Br(x0)nΩ¯B_{r}(x_{0})\subset\subset\mathbb{R}^{n}\setminus\overline{\Omega}. By the definition of uu, there exists a sequence {vk}𝒮cφ\{v_{k}\}\subset\mathcal{S}_{c}^{\varphi} such that u(x0)=limkvk(x0)u(x_{0})=\lim_{k\to\infty}v_{k}(x_{0}). Let v¯k=max{vk,v}\overline{v}_{k}=\max\{v_{k},v^{-}\}. Since vkv_{k}, v𝒮cφv^{-}\in\mathcal{S}_{c}^{\varphi}, v¯kC0(nΩ)\overline{v}_{k}\in C^{0}(\mathbb{R}^{n}\setminus\Omega) is a viscosity subsolution of (1.3) in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega} satisfying

v¯kφon Ωandlim sup|x|(v¯k(x)12|x|2)c.\overline{v}_{k}\leq\varphi\ \text{on }\partial\Omega\quad\text{and}\quad\limsup_{|x|\to\infty}\bigg{(}\overline{v}_{k}(x)-\frac{1}{2}|x|^{2}\bigg{)}\leq c.

We thus get v¯k𝒮cφ\overline{v}_{k}\in\mathcal{S}_{c}^{\varphi} and

u(x0)=limkv¯k(x0).u(x_{0})=\lim_{k\to\infty}\overline{v}_{k}(x_{0}).

Consider the lifting function w¯k\overline{w}_{k} of v¯k\overline{v}_{k} with respect to (1.3) in Br(x0)B_{r}(x_{0}). It follows from Proposition 3.2 that w¯k𝒮cφ\overline{w}_{k}\in\mathcal{S}_{c}^{\varphi} and

v¯kw¯kuin Br(x0).\overline{v}_{k}\leq\overline{w}_{k}\leq u\quad\text{in }B_{r}(x_{0}).

Hence,

u(x0)=limkw¯k(x0).u(x_{0})=\lim_{k\to\infty}\overline{w}_{k}(x_{0}).

Note that vw¯kv+v^{-}\leq\overline{w}_{k}\leq v^{+} in Br(x0)B_{r}(x_{0}). Hence, {w¯k}\{\overline{w}_{k}\} is uniformly bounded in Br(x0)B_{r}(x_{0}). By the local Lipschitz estimate for convex functions (see [13, Theorem 1 in Chapter 6.3]), we further have {|Dw¯k|}\{|D\overline{w}_{k}|\} is locally uniformly bounded in Br(x0)B_{r}(x_{0}). So, up to a subsequence, {w¯k}\{\overline{w}_{k}\} locally uniformly converges to some convex function w¯\overline{w} in Br(x0)B_{r}(x_{0}). By Lemma 3.5, w¯\overline{w} is a viscosity solution of (1.3) in Br(x0)B_{r}(x_{0}). It follows that

w¯uin Br(x0)andw¯(x0)=u(x0).\overline{w}\leq u\ \text{in }B_{r}(x_{0})\quad\text{and}\quad\overline{w}(x_{0})=u(x_{0}).

It remains to prove w¯u\overline{w}\geq u in Br(x0)B_{r}(x_{0}). Take yBr(x0)y\in B_{r}(x_{0}). There exists a sequence {Vk}𝒮cφ\{V_{k}\}\subset\mathcal{S}_{c}^{\varphi} such that u(y)=limkVk(y)u(y)=\lim_{k\to\infty}V_{k}(y). Let

V¯k=max{Vk,v¯k}.\overline{V}_{k}=\max\{V_{k},\overline{v}_{k}\}.

Then V¯k𝒮cφ\overline{V}_{k}\in\mathcal{S}_{c}^{\varphi}. Consider the lifting function W¯k\overline{W}_{k} of V¯k\overline{V}_{k} with respect to (1.3) in Br(x0)B_{r}(x_{0}). It follows from Proposition 3.2 that W¯k𝒮cφ\overline{W}_{k}\in\mathcal{S}_{c}^{\varphi} and

VkV¯kW¯kuin Br(x0).V_{k}\leq\overline{V}_{k}\leq\overline{W}_{k}\leq u\quad\text{in }B_{r}(x_{0}).

Hence,

u(y)=limkW¯k(y).u(y)=\lim_{k\to\infty}\overline{W}_{k}(y).

Also note that vW¯kv+v^{-}\leq\overline{W}_{k}\leq v^{+} in Br(x0)B_{r}(x_{0}). Similar arguments for {w¯k}\{\overline{w}_{k}\} apply to {W¯k}\{\overline{W}_{k}\} gives that {W¯k}\{\overline{W}_{k}\} locally uniformly converges to some convex function W¯\overline{W} in Br(x0)B_{r}(x_{0}), and W¯\overline{W} is a viscosity solution of (1.3) in Br(x0)B_{r}(x_{0}) satisfying

W¯(y)=u(y).\overline{W}(y)=u(y).

It follows that

v¯kV¯k,w¯kW¯k,w¯W¯in Br(x0).\overline{v}_{k}\leq\overline{V}_{k},\quad\overline{w}_{k}\leq\overline{W}_{k},\quad\overline{w}\leq\overline{W}\quad\text{in }B_{r}(x_{0}).

Then

u(x0)=w¯(x0)W¯(x0)u(x0).u(x_{0})=\overline{w}(x_{0})\leq\overline{W}(x_{0})\leq u(x_{0}).

We conclude that

{det(D2w¯)=det(D2W¯)=1in Br(x0),w¯W¯in Br(x0),w¯(x0)=W¯(x0).\begin{cases}\det(D^{2}\overline{w})=\det(D^{2}\overline{W})=1&\text{in }B_{r}(x_{0}),\\ \overline{w}\leq\overline{W}&\text{in }B_{r}(x_{0}),\\ \overline{w}(x_{0})=\overline{W}(x_{0}).\end{cases}

By the interior regularity of viscosity solution of (1.3) (see for instance [4]), one has w¯\overline{w}, W¯C(Br2(x0)¯)\overline{W}\in C^{\infty}(\overline{B_{\frac{r}{2}}(x_{0})}). This yields that

i,j=1naij(x)Dij(W¯w¯)(x)=0for xBr2(x0),\sum_{i,j=1}^{n}a_{ij}(x)D_{ij}(\overline{W}-\overline{w})(x)=0\quad\text{for }x\in B_{\frac{r}{2}}(x_{0}),

where

aij(x)=01(det)1nξij(D2w¯+t(D2W¯D2w¯))dt.a_{ij}(x)=\int_{0}^{1}\frac{\partial(\det)^{\frac{1}{n}}}{\partial\xi_{ij}}(D^{2}\overline{w}+t(D^{2}\overline{W}-D^{2}\overline{w}))\mathrm{d}t.

Since D2w¯D^{2}\overline{w} and D2W¯D^{2}\overline{W} are bounded in Br2(x0)¯\overline{B_{\frac{r}{2}}(x_{0})}, we have (aij(x))δI(a_{ij}(x))\geq\delta I in Br2(x0)B_{\frac{r}{2}}(x_{0}) for some constant δ>0\delta>0. The strong maximum principle for uniformly elliptic linear equation implies that

W¯=w¯in Br(x0).\overline{W}=\overline{w}\quad\text{in }B_{r}(x_{0}).

This gives w¯(y)=W¯(y)=u(y)\overline{w}(y)=\overline{W}(y)=u(y). By the arbitrariness of yy, we have

u=w¯in Br(x0).u=\overline{w}\quad\text{in }B_{r}(x_{0}).

This completes the proof. ∎

Remark 3.7.

We see from the proof that if φ(x)12|x|2\varphi(x)-\frac{1}{2}|x|^{2} is bounded on Ω\partial\Omega, then Proposition 3.6 also holds for unbounded Ω\Omega.

4. Asymptotic behavior of the Perron’s solution near infinity

In this section, we will demonstrate that the Perron’s solution achieves the asymptotic behavior 12|x|2+c\frac{1}{2}|x|^{2}+c near infinity, provided cc large.

Proposition 4.1.

Let Ω\Omega be a bounded open set of n\mathbb{R}^{n} satisfying Br1(0)ΩBr2(0)B_{r_{1}}(0)\subset\Omega\subset\subset B_{r_{2}}(0) for some positive constants r1r_{1} and r2r_{2}, n3n\geq 3, and let φ\varphi be a bounded function on Ω\partial\Omega. Then the function uu defined by (3.1) satisfies

lim|x|(u(x)(12|x|2+c))=0,\lim_{|x|\to\infty}\left(u(x)-\left(\frac{1}{2}|x|^{2}+c\right)\right)=0,

where

cmax{μ(r1n),supxΩ(φ(x)12|x|2)},c\geq\max\bigg{\{}\mu(-r_{1}^{n}),\sup_{x\in\partial\Omega}\bigg{(}\varphi(x)-\frac{1}{2}|x|^{2}\bigg{)}\bigg{\}}, (4.1)

and

μ(α)=infΩφ12r22+r2s((1+αsn)1n1)dsforαr1n.\mu(\alpha)=\inf_{\partial\Omega}\varphi-\frac{1}{2}r_{2}^{2}+\int_{r_{2}}^{\infty}s\bigg{(}\bigg{(}1+\frac{\alpha}{s^{n}}\bigg{)}^{\frac{1}{n}}-1\bigg{)}\mathrm{d}s\quad\text{for}~{}\alpha\geq-r_{1}^{n}.
Proof.

For αr1n\alpha\geq-r_{1}^{n}, let

wα(x)=infΩφ+r2|x|(sn+α)1ndsfor xnΩ.w_{\alpha}(x)=\inf_{\partial\Omega}\varphi+\int_{r_{2}}^{|x|}(s^{n}+\alpha)^{\frac{1}{n}}\mathrm{d}s\quad\text{for }x\in\mathbb{R}^{n}\setminus\Omega. (4.2)

Then wαC0(nΩ)w_{\alpha}\in C^{0}(\mathbb{R}^{n}\setminus\Omega) is locally convex in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega}, and

{detD2wα=1in nΩ¯,wαφon Ω.\begin{cases}\det{D^{2}w_{\alpha}}=1&\mbox{in }\mathbb{R}^{n}\setminus\overline{\Omega},\\ w_{\alpha}\leq\varphi&\mbox{on }\partial\Omega.\end{cases}

Direct calculation shows

lim|x|(wα(x)12|x|2)=infΩφ12r22+r2s((1+αsn)1n1)ds=:μ(α).\begin{split}&\lim_{|x|\to\infty}\bigg{(}w_{\alpha}(x)-\frac{1}{2}|x|^{2}\bigg{)}\\ =&\inf_{\partial\Omega}\varphi-\frac{1}{2}r_{2}^{2}+\int_{r_{2}}^{\infty}s\bigg{(}\bigg{(}1+\frac{\alpha}{s^{n}}\bigg{)}^{\frac{1}{n}}-1\bigg{)}\mathrm{d}s\\ =&:\mu(\alpha).\end{split}

Clearly, μ(α)\mu(\alpha) is smooth and strictly increasing with respect to α[r1n,)\alpha\in[-r_{1}^{n},\infty) and

limαμ(α)=.\lim_{\alpha\to\infty}\mu(\alpha)=\infty.

Then for cμ(r1n)c\geq\mu(-r_{1}^{n}), we have wμ1(c)𝒮cφw_{\mu^{-1}(c)}\in\mathcal{S}_{c}^{\varphi}. It follows that

lim inf|x|(u(x)12|x|2)lim|x|(wμ1(c)(x)12|x|2)=c.\liminf_{|x|\to\infty}\bigg{(}u(x)-\frac{1}{2}|x|^{2}\bigg{)}\geq\lim_{|x|\to\infty}\bigg{(}w_{\mu^{-1}(c)}(x)-\frac{1}{2}|x|^{2}\bigg{)}=c.

Recall that for csupxΩ(φ(x)12|x|2)c\geq\sup_{x\in\partial\Omega}\left(\varphi(x)-\frac{1}{2}|x|^{2}\right),

lim sup|x|(u(x)12|x|2)lim|x|(v+(x)12|x|2)=c,\limsup_{|x|\to\infty}\bigg{(}u(x)-\frac{1}{2}|x|^{2}\bigg{)}\leq\lim_{|x|\to\infty}\bigg{(}v^{+}(x)-\frac{1}{2}|x|^{2}\bigg{)}=c,

where v+(x)=12|x|2+cv^{+}(x)=\frac{1}{2}|x|^{2}+c is as in the proof of Proposition 3.6. This finishes the proof. ∎

5. Boundary behavior of the Perron’s solution

In this section, we deal with the boundary behavior of the Perron’s solution. The key lies in proving a barrier lemma. We shall overcome the difficulty of non C2C^{2} regularity of boundary values and domains.

Let φ\varphi be semi-convex with respect to Ω\partial\Omega at ξ\xi. Namely, ψ(x):=φ(x,ρ(x))\psi(x^{\prime}):=\varphi(x^{\prime},\rho(x^{\prime})) is semi-convex in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}) under the local coordinate system at ξΩ\xi\in\partial\Omega, where δ(ξ)\delta(\xi) is as in (1.5). Then there exist K(ξ)>0K(\xi)>0 and p(ξ)n1p(\xi)\in\mathbb{R}^{n-1} such that

ψ(x)+K(ξ)2|x|2ψ(0)+p(ξ)xfor |x|<δ(ξ).\psi(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2}\geq\psi(0^{\prime})+p(\xi)\cdot x^{\prime}\quad\text{for }|x^{\prime}|<\delta(\xi). (5.1)

By the semi-convexity of ψ\psi at ξ\xi, ψ\psi is continuous in Bδ(ξ)(0)B^{\prime}_{\delta(\xi)}(0^{\prime}), and so φ\varphi is continuous on ΩBδ(ξ)(0)\partial\Omega\cap B_{\delta(\xi)}(0).

Now, we are able to prove the following barrier lemma.

Lemma 5.1.

Let Ω\Omega be a convex domain, and let φ\varphi be a bounded function on Ω\partial\Omega. Suppose that Ω\Omega satisfies an enclosing sphere condition at ξΩ\xi\in\partial\Omega, and φ\varphi is semi-convex with respect to Ω\partial\Omega at ξ\xi. Then there exists x¯(ξ)n\bar{x}(\xi)\in\mathbb{R}^{n} such that

|x¯(ξ)|C(ξ)andwξ<φon Ω{ξ},|\bar{x}(\xi)|\leq C(\xi)\quad\text{and}\quad w_{\xi}<\varphi\quad\text{on }\partial\Omega\setminus\{\xi\},

where

wξ(x)=φ(ξ)+12(|xx¯(ξ)|2|ξx¯(ξ)|2)forxn,w_{\xi}(x)=\varphi(\xi)+\frac{1}{2}\left(|x-\bar{x}(\xi)|^{2}-|\xi-\bar{x}(\xi)|^{2}\right)\quad\text{for}\ x\in\mathbb{R}^{n},

and C(ξ)>0C(\xi)>0 is a constant depending only on δ(ξ)\delta(\xi), r(ξ)r(\xi), K(ξ)K(\xi), |p(ξ)||p(\xi)|, supΩ|φ|\sup_{\partial\Omega}|\varphi| and the C0,1C^{0,1} norm of Ω\partial\Omega. Here δ(ξ)\delta(\xi) and r(ξ)r(\xi) are as in (1.5) and Definition 1.4 respectively, K(ξ)K(\xi) and p(ξ)p(\xi) are as in (5.1).

Proof.

By a translation and a rotation, we may assume without losing the generality that the coordinate system in n\mathbb{R}^{n} is just the local coordinate system at ξ\xi, and xnx_{n}-axis is along the direction of y(ξ)ξy(\xi)-\xi, where y(ξ)y(\xi) is the center of the enclosing sphere at ξ\xi. Then ξ=0\xi=0, y(ξ)=(0,r(ξ))y(\xi)=(0^{\prime},r(\xi)) and Ω\partial\Omega can be locally represented by the graph of xn=ρ(x)x_{n}=\rho(x^{\prime}), |x|<δ(ξ)|x^{\prime}|<\delta(\xi). Let

x¯=(p(ξ),R),\bar{x}=(-p(\xi),R),

where R>0R>0 will be chosen later. Let

w(x)=φ(0)+12(|xx¯|2|x¯|2)forxn.w(x)=\varphi(0)+\frac{1}{2}\left(|x-\bar{x}|^{2}-|\bar{x}|^{2}\right)\quad\text{for}\ x\in\mathbb{R}^{n}.

Denote δ=δ(ξ)\delta=\delta(\xi), r=r(ξ)r=r(\xi), K=K(ξ)K=K(\xi) and p=p(ξ)p=p(\xi). It is clear from the enclosing sphere condition ΩBr((0,r))¯\partial\Omega\subset\overline{B_{r}((0^{\prime},r))} that

|x|2+(ρ(x)r)2r2for|x|<δ,|x^{\prime}|^{2}+(\rho(x^{\prime})-r)^{2}\leq r^{2}\quad\text{for}~{}|x^{\prime}|<\delta,

and so

ρ(x)12r|x|2for|x|<δ.\rho(x^{\prime})\geq\frac{1}{2r}|x^{\prime}|^{2}\quad\text{for}~{}|x^{\prime}|<\delta. (5.2)

Since Ω\Omega is bounded and convex, Ω\partial\Omega is Lipschitz continuous and

ρ(x)C0|x|for |x|<δ,\rho(x^{\prime})\leq C_{0}|x^{\prime}|\quad\text{for }|x^{\prime}|<\delta, (5.3)

where C0>0C_{0}>0 is the C0,1C^{0,1} norm of Ω\partial\Omega.

Case 1. x=(x,ρ(x))Ω{0}x=(x^{\prime},\rho(x^{\prime}))\in\partial\Omega\setminus\{0\} and |x|<δ|x^{\prime}|<\delta. It follows from (5.1)-(5.3) that

(wφ)(x)=12|x|2xx¯+φ(0)φ(x)=12|x|2+12ρ(x)2Rρ(x)+φ(0)φ(x,ρ(x))+px12(1+C02Rr)|x|2+ψ(0)ψ(x)+px12(1+C02Rr+K)|x|2< 0,\begin{split}(w-\varphi)(x)=&\ \frac{1}{2}|x|^{2}-x\cdot\bar{x}+\varphi(0)-\varphi(x)\\ =&\ \frac{1}{2}|x^{\prime}|^{2}+\frac{1}{2}\rho(x^{\prime})^{2}-R\rho(x^{\prime})+\varphi(0)-\varphi(x^{\prime},\rho(x^{\prime}))+p\cdot x^{\prime}\\ \leq&\ \frac{1}{2}\bigg{(}1+C_{0}^{2}-\frac{R}{r}\bigg{)}|x^{\prime}|^{2}+\psi(0^{\prime})-\psi(x^{\prime})+p\cdot x^{\prime}\\ \leq&\ \frac{1}{2}\bigg{(}1+C_{0}^{2}-\frac{R}{r}+K\bigg{)}|x^{\prime}|^{2}\\ <&\ 0,\end{split}

provided

R>R1:=r(1+C02+K).R>R_{1}:=r(1+C_{0}^{2}+K).

Case 2. xΩ{(x,ρ(x))||x|<δ}x\in\partial\Omega\setminus\{(x^{\prime},\rho(x^{\prime}))|\,|x^{\prime}|<\delta\}. By the convexity of Ω\Omega and (5.2), we have xn12rδ2x_{n}\geq\frac{1}{2r}\delta^{2}. It follows that

(wφ)(x)=12|x|2xx¯+φ(0)φ(x)=12|x|2+pxRxn+φ(0)φ(x)12(2r)2+r|p|R2rδ2+2supΩ|φ|< 0,\begin{split}(w-\varphi)(x)=&\ \frac{1}{2}|x|^{2}-x\cdot\bar{x}+\varphi(0)-\varphi(x)\\ =&\ \frac{1}{2}|x|^{2}+p\cdot x^{\prime}-Rx_{n}+\varphi(0)-\varphi(x)\\ \leq&\ \frac{1}{2}(2r)^{2}+r|p|-\frac{R}{2r}\delta^{2}+2\sup_{\partial\Omega}|\varphi|\\ <&\ 0,\end{split}

provided

R>R2:=2rδ2(2r2+r|p|+2supΩ|φ|).R>R_{2}:=\frac{2r}{\delta^{2}}(2r^{2}+r|p|+2\sup_{\partial\Omega}|\varphi|).

Take R>max{R1,R2}R>\max\{R_{1},R_{2}\}. In both cases, we conclude w<φw<\varphi on Ω{ξ}\partial\Omega\setminus\{\xi\}. We finish the proof by taking C(ξ)=|p|2+R2C(\xi)=\sqrt{|p|^{2}+R^{2}}. ∎

Benefiting from Lemma 5.1, we obtain that the Perron’s solution can be continuously extended to the boundary Ω\partial\Omega in a pointwise way.

Proposition 5.2.

Let uu be the function defined by (3.1). Suppose that Ω\Omega, φ\varphi and ξΩ\xi\in\partial\Omega satisfy the assumptions in Lemma 5.1. Then there exists a constant c0(ξ)c_{0}(\xi), depending only on nn, δ(ξ)\delta(\xi), r(ξ)r(\xi), K(ξ)K(\xi), |p(ξ)||p(\xi)|, supΩ|φ|\sup_{\partial\Omega}|\varphi| and the C0,1C^{0,1} norm of Ω\partial\Omega, such that for every c>c0(ξ)c>c_{0}(\xi), u(x)φ(ξ)u(x)\to\varphi(\xi) as nΩ¯xξ\mathbb{R}^{n}\setminus\overline{\Omega}\ni x\to\xi.

Proof.

Since Ω\Omega satisfies an enclosing sphere condition at ξ\xi, Ω\Omega is bounded. By a translation, we may assume Br1(0)ΩBr2(0)B_{r_{1}}(0)\subset\Omega\subset\subset B_{r_{2}}(0) for some constants r1,r2>0r_{1},r_{2}>0. By Lemma 5.1, we have

wξ(x)=φ(ξ)+12|x|212|ξ|2+(ξx)x¯(ξ)φ(ξ)12r222r2C(ξ)φ(ξ)C1forxBr2(0)Ω,\begin{split}w_{\xi}(x)&=\varphi(\xi)+\frac{1}{2}|x|^{2}-\frac{1}{2}|\xi|^{2}+(\xi-x)\cdot\bar{x}(\xi)\\ &\geq\varphi(\xi)-\frac{1}{2}r_{2}^{2}-2r_{2}C(\xi)\\ &\geq\varphi(\xi)-C_{1}\quad\text{for}~{}x\in B_{r_{2}}(0)\setminus\Omega,\end{split}

and

wξ(x)=φ(ξ)+12|x|212|ξ|2+(ξx)x¯(ξ)φ(ξ)+12(r2+1)2+(2r2+1)C(ξ)φ(ξ)+C1forxBr2+1(0)Br2(0),\begin{split}w_{\xi}(x)&=\varphi(\xi)+\frac{1}{2}|x|^{2}-\frac{1}{2}|\xi|^{2}+(\xi-x)\cdot\bar{x}(\xi)\\ &\leq\varphi(\xi)+\frac{1}{2}(r_{2}+1)^{2}+(2r_{2}+1)C(\xi)\\ &\leq\varphi(\xi)+C_{1}\quad\text{for}~{}x\in B_{r_{2}+1}(0)\setminus B_{r_{2}}(0),\end{split}

where C(ξ)C(\xi) is as in Lemma 5.1 and C1C_{1} depends only on C(ξ)C(\xi) and r2r_{2}. Take α0r1n\alpha_{0}\geq-r_{1}^{n} such that

infΩφ+r2r2+1(sn+α0)1nds>supΩφ+2C1,\inf_{\partial\Omega}\varphi+\int_{r_{2}}^{r_{2}+1}\left(s^{n}+\alpha_{0}\right)^{\frac{1}{n}}\mathrm{d}s>\sup_{\partial\Omega}\varphi+2C_{1},

and let

w¯α(x)=wα(x)C1,\underline{w}_{\alpha}(x)=w_{\alpha}(x)-C_{1},

where wαw_{\alpha} is given by (4.2). Then for any αα0\alpha\geq\alpha_{0},

wξinfΩφC1w¯αinBr2(0)Ω,w_{\xi}\geq\inf_{\partial\Omega}\varphi-C_{1}\geq\underline{w}_{\alpha}\quad\text{in}~{}B_{r_{2}}(0)\setminus\Omega,
wξ<w¯α0w¯αonBr2+1(0).w_{\xi}<\underline{w}_{\alpha_{0}}\leq\underline{w}_{\alpha}\quad\text{on}~{}\partial B_{r_{2}+1}(0).

Let

u¯(x)={wξ(x),xBr2(0)Ω,max{w¯α(x),wξ(x)},xBr2+1(0)Br2(0),w¯α(x),xnBr2+1(0).\underline{u}(x)=\begin{cases}w_{\xi}(x),&x\in B_{r_{2}}(0)\setminus\Omega,\\ \max\{\underline{w}_{\alpha}(x),w_{\xi}(x)\},&x\in B_{r_{2}+1}(0)\setminus B_{r_{2}}(0),\\ \underline{w}_{\alpha}(x),&x\in\mathbb{R}^{n}\setminus B_{r_{2}+1}(0).\end{cases}

In view of Lemma 3.3, u¯𝒮cφ\underline{u}\in\mathcal{S}_{c}^{\varphi} provided cμ(α0)c\geq\mu(\alpha_{0}). Hence,

lim infnΩ¯xξu(x)lim infnΩ¯xξu¯(x)=wξ(ξ)=φ(ξ).\liminf_{\mathbb{R}^{n}\setminus\overline{\Omega}\ni x\to\xi}u(x)\geq\liminf_{\mathbb{R}^{n}\setminus\overline{\Omega}\ni x\to\xi}\underline{u}(x)=w_{\xi}(\xi)=\varphi(\xi).

Recall that φ\varphi is continuous in a neighborhood of ξ\xi on Ω\partial\Omega. Since φ\varphi is bounded, the extension theorem implies that there exists φ¯C0(Ω)\overline{\varphi}\in C^{0}(\partial\Omega) satisfying φ¯=φ\overline{\varphi}=\varphi near ξ\xi and

φ¯φon Ω.\overline{\varphi}\geq\varphi\quad\text{on }\partial\Omega.

Since Ω\Omega is bounded and convex, Ω\partial\Omega is Lipschitz continuous. Then Br2(0)Ω¯B_{r_{2}}(0)\setminus\overline{\Omega} satisfies the exterior cone condition. By Proposition B, there exists w+C0(Br2(0)Ω¯)w^{+}\in C^{0}(\overline{B_{r_{2}}(0)\setminus\Omega}) satisfying

{Δw+=0in Br2(0)Ω¯,w+=φ¯on Ω,w+=maxBr2(0)uon Br2(0).\begin{cases}\Delta w^{+}=0&\mbox{in }B_{r_{2}}(0)\setminus\overline{\Omega},\\ w^{+}=\overline{\varphi}&\mbox{on }\partial\Omega,\\ w^{+}=\max_{\partial B_{r_{2}}(0)}u&\mbox{on }\partial B_{r_{2}}(0).\end{cases}

Then comparison principal gives that for any v𝒮cφv\in\mathcal{S}_{c}^{\varphi},

vw+in Br2(0)Ω¯.v\leq w^{+}\quad\text{in }B_{r_{2}}(0)\setminus\overline{\Omega}.

It follows that

uw+in Br2(0)Ω¯,u\leq w^{+}\quad\text{in }B_{r_{2}}(0)\setminus\overline{\Omega},

and so

lim supnΩ¯xξu(x)limnΩ¯xξw+(x)=φ¯(ξ)=φ(ξ).\limsup_{\mathbb{R}^{n}\setminus\overline{\Omega}\ni x\to\xi}u(x)\leq\lim_{\mathbb{R}^{n}\setminus\overline{\Omega}\ni x\to\xi}w^{+}(x)=\overline{\varphi}(\xi)=\varphi(\xi).

Therefore, uu satisfies the boundary condition at ξ\xi. ∎

6. Proof of Theorem 1.5

In the previous section, we proved that the Perron’s solution is continuous up to a boundary point ξ\xi when c>c0(ξ)c>c_{0}(\xi). In this section, we will give the uniform estimates for c0(ξ)c_{0}(\xi) with respect to ξΩ\xi\in\partial\Omega, and thus obtain that the Perron’s solution is continuous up to the whole boundary when cc is sufficiently large. Combining with the conclusions in Sections 3-5, we can complete the proof of the existence part of Theorem 1.5. The nonexistence part can be deduced as in [22].

To establish the uniform estimate for c0(ξ)c_{0}(\xi), we need the following uniform estimates for δ(ξ)\delta(\xi), K(ξ)K(\xi) and p(ξ)p(\xi), where δ(ξ)\delta(\xi) is as in (1.5), and K(ξ)K(\xi) and p(ξ)p(\xi) are as in (5.1). These estimates are also of independent interest.

Lemma 6.1.

Let Ω\Omega be a bounded open set of n\mathbb{R}^{n}, ΩC1\partial\Omega\in C^{1}. Then Ω\Omega satisfies

δ:=infΩδ(ξ)>0.\delta:=\inf_{\partial\Omega}\delta(\xi)>0. (H)
Proof.

By the finite covering theorem, there exists {ξ(1),,ξ(N)}Ω\{\xi^{(1)},\cdots,\xi^{(N)}\}\subset\partial\Omega such that

Ω=i=1N{((x(i)),ρ(i)((x(i))))||(x(i))|<δi4},\partial\Omega=\bigcup_{i=1}^{N}\bigg{\{}((x^{(i)})^{\prime},\rho^{(i)}((x^{(i)})^{\prime}))|\,|(x^{(i)})^{\prime}|<\frac{\delta_{i}}{4}\bigg{\}},

where x(i)x^{(i)} is the coordinate under the local coordinate system at ξ(i)\xi^{(i)} and δi=δ(ξ(i))\delta_{i}=\delta(\xi^{(i)}). We may assume ω(δi)<12\omega(\delta_{i})<\frac{1}{2}, where ω\omega is the modulus of the continuity of DρD^{\prime}\rho. Let δ=18min{δ1,,δN}\delta=\frac{1}{8}\min\{\delta_{1},\cdots,\delta_{N}\}.

For any fixed ξ~Ω\widetilde{\xi}\in\partial\Omega, there exists i0{1,,N}i_{0}\in\{1,\cdots,N\} such that

ξ~{(x,ρ(x))||x|<δi04}.\widetilde{\xi}\in\bigg{\{}(x^{\prime},\rho(x^{\prime}))|\,|x^{\prime}|<\frac{\delta_{i_{0}}}{4}\bigg{\}}.

Here we write the local coordinate ((x(i0)),ρ(i0)((x(i0))))((x^{(i_{0})})^{\prime},\rho^{(i_{0})}((x^{(i_{0})})^{\prime})) as (x,ρ(x))(x^{\prime},\rho(x^{\prime})) for simplicity. We may assume the coordinate system in n\mathbb{R}^{n} is just the local coordinate system at ξ(i0)\xi^{(i_{0})}. Since Ω\partial\Omega is C1C^{1}, we have

|Dρ(ξ~)|=|Dρ(ξ~)Dρ(0)|ω(|ξ~|)ω(δi04).|D^{\prime}\rho(\widetilde{\xi}^{\prime})|=|D^{\prime}\rho(\widetilde{\xi}^{\prime})-D^{\prime}\rho(0^{\prime})|\leq\omega(|\widetilde{\xi}^{\prime}|)\leq\omega\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}. (6.1)

Clearly, the unit inner normal of Ω\partial\Omega at 0 and ξ~\widetilde{\xi} are

ν(0)=(0,1)andν(ξ~)=(Dρ(ξ~),1)|Dρ(ξ~)|2+1,\nu(0)=(0^{\prime},1)\quad\text{and}\quad\nu(\widetilde{\xi})=\frac{(-D^{\prime}\rho(\widetilde{\xi^{\prime}}),1)}{\sqrt{|D^{\prime}\rho(\widetilde{\xi}^{\prime})|^{2}+1}},

respectively. It follows from (6.1) that

|ν(ξ~)ν(0)|=2(11|Dρ(ξ~)|2+1)<|Dρ(ξ~)|ω(δi04).|\nu(\widetilde{\xi})-\nu(0)|=\sqrt{2\Bigg{(}1-\frac{1}{\sqrt{|D^{\prime}\rho(\widetilde{\xi}^{\prime})|^{2}+1}}\Bigg{)}}<|D^{\prime}\rho(\widetilde{\xi}^{\prime})|\leq\omega\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}. (6.2)
[Uncaptioned image]

Denote by x~\widetilde{x} the coordinate under the local coordinate system at ξ~=0~\widetilde{\xi}=\widetilde{0}. Then there exists a n×nn\times n orthogonal matrix Q=(qij)Q=(q_{ij}) depending only on ξ~\widetilde{\xi} such that

x~=(xξ~)Q.\widetilde{x}=(x-\widetilde{\xi})Q. (6.3)

Note that the coordinate of ν(ξ~)\nu(\widetilde{\xi}) under the local coordinate system at ξ~\widetilde{\xi} is (0,1)(0^{\prime},1). This yields that

ν~(ξ~)=ν(ξ~)Q,\widetilde{\nu}(\widetilde{\xi})=\nu(\widetilde{\xi})Q,

and so

ν(ξ~)=ν(0)Q1=ν(0)QT=(0,1)QT=(q1n,,q(n1)n,qnn).\nu(\widetilde{\xi})=\nu(0)Q^{-1}=\nu(0)Q^{T}=(0^{\prime},1)Q^{T}=(q_{1n},\cdots,q_{(n-1)n},q_{nn}).

Combining with (6.2), we get

|(q1n,,q(n1)n)|<w(δi04)and1w(δi04)<qnn1.|(q_{1n},\cdots,q_{(n-1)n})|<w\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}\quad\text{and}\quad 1-w\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}<q_{nn}\leq 1.

It follows that

(x~Q1)=(x~QT)=x~(QT)+x~n(q1n,,q(n1)n),(\widetilde{x}Q^{-1})^{\prime}=(\widetilde{x}Q^{T})^{\prime}=\widetilde{x}^{\prime}(Q^{T})^{\prime}+\widetilde{x}_{n}(q_{1n},\cdots,q_{(n-1)n}), (6.4)

where (QT)(Q^{T})^{\prime} denotes the matrix composed of the first (n1)(n-1) rows and columns of QTQ^{T}. We see for (x~,x~n)E:={(x~,x~n)||x~|<δ,|x~n|<δw(δi04)}(\widetilde{x}^{\prime},\widetilde{x}_{n})\in E:=\Big{\{}(\widetilde{x}^{\prime},\widetilde{x}_{n})|\,|\widetilde{x}^{\prime}|<\delta,~{}|\widetilde{x}_{n}|<\frac{\delta}{w\big{(}\frac{\delta_{i_{0}}}{4}\big{)}}\Big{\}},

|x|=|(x~Q1)+ξ~||x~(QT)|+|x~n(q1n,,q(n1)n)|+|ξ~||x~|+w(δi04)|x~n|+|ξ~|<δi02.\begin{split}|x^{\prime}|=|(\widetilde{x}Q^{-1})^{\prime}+\widetilde{\xi}^{\prime}|&\leq|\widetilde{x}^{\prime}(Q^{T})^{\prime}|+|\widetilde{x}_{n}(q_{1n},\cdots,q_{(n-1)n})|+|\widetilde{\xi}^{\prime}|\\ &\leq|\widetilde{x}^{\prime}|+w\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}|\widetilde{x}_{n}|+|\widetilde{\xi}^{\prime}|\\ &<\frac{\delta_{i_{0}}}{2}.\end{split}

For (x~,x~n)E(\widetilde{x}^{\prime},\widetilde{x}_{n})\in E, set

F(x~,x~n)=ρ((x~Q1+ξ~))(x~Q1+ξ~)n.F(\widetilde{x}^{\prime},\widetilde{x}_{n})=\rho((\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime})-(\widetilde{x}Q^{-1}+\widetilde{\xi})_{n}.

From ρC1\rho\in C^{1}, we have FC1(E)F\in C^{1}(E). Since for (x~,x~n)E(\widetilde{x}^{\prime},\widetilde{x}_{n})\in E, |x|<δi02|x^{\prime}|<\frac{\delta_{i_{0}}}{2}, and so |Dρ(x)|w(δi02)|D^{\prime}\rho(x^{\prime})|\leq w\Big{(}\frac{\delta_{i_{0}}}{2}\Big{)}. It follows that

Fx~n=Dρ((x~Q1+ξ~))(q1n,,q(n1)n)qnnw(δi02)w(δi04)(1w(δi04))<0,\begin{split}\frac{\partial F}{\partial\widetilde{x}_{n}}&=D^{\prime}\rho((\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime})\cdot(q_{1n},\cdots,q_{(n-1)n})-q_{nn}\\ &\leq w\bigg{(}\frac{\delta_{i_{0}}}{2}\bigg{)}w\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}-\bigg{(}1-w\bigg{(}\frac{\delta_{i_{0}}}{4}\bigg{)}\bigg{)}\\ &<0,\end{split}

due to w(δi02)<12w(\frac{\delta_{i_{0}}}{2})<\frac{1}{2}. In view of (6.3), we have

(x~Q1+ξ~)n=xn=ρ(x)=ρ((x~Q1+ξ~)).(\widetilde{x}Q^{-1}+\widetilde{\xi})_{n}=x_{n}=\rho(x^{\prime})=\rho((\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}).

That is F=0F=0 in E{(x~,x~n)|x~=(xξ~)Q}E\cap\{(\widetilde{x}^{\prime},\widetilde{x}_{n})|\,\widetilde{x}=(x-\widetilde{\xi})Q\}. Therefore, the implicit function theorem yields that there exists a function ρ~\widetilde{\rho} such that

x~n=ρ~(x~)for|x~|<δ2.\widetilde{x}_{n}=\widetilde{\rho}(\widetilde{x}^{\prime})\quad\text{for}~{}|\widetilde{x}^{\prime}|<\frac{\delta}{2}.

This finishes the proof. ∎

Lemma 6.2.

Let Ω\Omega be a bounded open set of n\mathbb{R}^{n} and ΩC1\partial\Omega\in C^{1}. If φ\varphi is semi-convex with respect to Ω\partial\Omega, then KK and |p||p| are bounded on Ω\partial\Omega.

Proof.

For any fixed ξΩ\xi\in\partial\Omega, Ω\partial\Omega can be locally represented by the graph of

xn=ρ(x)for|x|<δ.x_{n}=\rho(x^{\prime})\quad\text{for}~{}|x^{\prime}|<\delta.

Here δ\delta is independent of ξ\xi due to Lemma 6.1. Take ξ~Ω{(x,ρ(x))||x|<δ2}\widetilde{\xi}\in\partial\Omega\cap\{(x^{\prime},\rho(x^{\prime}))|\,|x^{\prime}|<\frac{\delta}{2}\}. Denote by x~\widetilde{x} the coordinate under the local coordinate system at ξ~\widetilde{\xi}. Then there exists a n×nn\times n orthogonal matrix QQ depending only on the local coordinate systems at ξ\xi and ξ~\widetilde{\xi} such that

x~=(xξ~)Q.\widetilde{x}=(x-\widetilde{\xi})Q.

It is easily seen that for x~Ω~\widetilde{x}\in\partial\widetilde{\Omega} with |x~|<δ~|\widetilde{x}^{\prime}|<\widetilde{\delta},

|x|=|(x~Q1+ξ~)||x~|+|ξ~|<δ2+δ2=δfor|x~|<δ~,|x^{\prime}|=|(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}|\leq|\widetilde{x}|+|\widetilde{\xi}^{\prime}|<\frac{\delta}{2}+\frac{\delta}{2}=\delta\quad\text{for}\ |\widetilde{x}^{\prime}|<\widetilde{\delta}, (6.5)

where δ~:=δ21+C02\widetilde{\delta}:=\frac{\delta}{2\sqrt{1+C_{0}^{2}}} and C0C_{0} is the C0,1C^{0,1} norm of Ω\partial\Omega. Correspondingly, we write

ψ~(x~)=ψ(x)=ψ((x~Q1+ξ~))for|x~|<δ~.\widetilde{\psi}(\widetilde{x}^{\prime})=\psi(x^{\prime})=\psi((\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime})\quad\text{for}~{}|\widetilde{x}^{\prime}|<\widetilde{\delta}. (6.6)

By (5.1), there exist positive constants K(ξ)K(\xi) and K(ξ~)K(\widetilde{\xi}) such that

ψ(x)+K(ξ)2|x|2andψ~(x~)+K(ξ~)2|x~|2\psi(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2}\quad\text{and}\quad\widetilde{\psi}(\widetilde{x}^{\prime})+\frac{K(\widetilde{\xi})}{2}|\widetilde{x}^{\prime}|^{2}

are convex in Bδ(0)B^{\prime}_{\delta}(0^{\prime}) and Bδ~(0~)B^{\prime}_{\widetilde{\delta}}(\widetilde{0}^{\prime}), respectively.

We first prove that KK is bounded on Ω{(x,ρ(x))||x|<δ2}\partial\Omega\cap\{(x^{\prime},\rho(x^{\prime}))|\,|x^{\prime}|<\frac{\delta}{2}\}. For x~,y~Ω\widetilde{x},\widetilde{y}\in\partial\Omega with |x~|,|y~|<δ~|\widetilde{x}^{\prime}|,|\widetilde{y}^{\prime}|<\widetilde{\delta} and 0<t<10<t<1, we have by (6.5),

|x|=|(x~Q1+ξ~)|<δand|y|=|(y~Q1+ξ~)|<δ.|x^{\prime}|=|(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}|<\delta\quad\text{and}\quad|y^{\prime}|=|(\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime}|<\delta. (6.7)

It follows from (6.6), (6.7) and the semi-convexity of ψ\psi in Bδ(0)B^{\prime}_{\delta}(0^{\prime}) that

ψ~(tx~+(1t)y~)=ψ(t(x~Q1+ξ~)+(1t)(y~Q1+ξ~))+K(ξ)2|t(x~Q1+ξ~)+(1t)(y~Q1+ξ~)|2K(ξ)2|t(x~Q1+ξ~)+(1t)(y~Q1+ξ~)|2\begin{split}\widetilde{\psi}({t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}})=&\,\psi(t(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}+(1-t)(\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime})\\ &\,+\frac{K(\xi)}{2}|t(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}+(1-t)(\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime}|^{2}\\ &\,-\frac{K(\xi)}{2}|t(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}+(1-t)(\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime}|^{2}\\ \end{split}
t(ψ((x~Q1+ξ~))+K(ξ)2|(x~Q1+ξ~)|2)+(1t)(ψ((y~Q1+ξ~))+K(ξ)2|(y~Q1+ξ~)|2)K(ξ)2|t(x~Q1+ξ~)+(1t)(y~Q1+ξ~)|2=tψ~(x~)+(1t)ψ~(y~)+t(1t)K(ξ)2|((x~y~)Q1)|2.\begin{split}\leq&\,t\bigg{(}\psi((\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime})+\frac{K(\xi)}{2}|(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}|^{2}\bigg{)}\\ &\,+(1-t)\bigg{(}\psi((\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime})+\frac{K(\xi)}{2}|(\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime}|^{2}\bigg{)}\\ &\,-\frac{K(\xi)}{2}|t(\widetilde{x}Q^{-1}+\widetilde{\xi})^{\prime}+(1-t)(\widetilde{y}Q^{-1}+\widetilde{\xi})^{\prime}|^{2}\\ =&\,t\widetilde{\psi}(\widetilde{x}^{\prime})+(1-t)\widetilde{\psi}(\widetilde{y}^{\prime})+\frac{t(1-t)K(\xi)}{2}|((\widetilde{x}-\widetilde{y})Q^{-1})^{\prime}|^{2}.\\ \end{split}

Taking K(ξ~)=(1+C02)K(ξ)K(\widetilde{\xi})=(1+C_{0}^{2})K(\xi), we obtain for x~,y~Bδ~(0~)\widetilde{x}^{\prime},\widetilde{y}^{\prime}\in B^{\prime}_{\widetilde{\delta}}(\widetilde{0}),

ψ~(tx~+(1t)y~)+K(ξ~)2|tx~+(1t)y~|2tψ~(x~)+(1t)ψ~(y~)+t(1t)K(ξ~)2|x~y~|2+K(ξ~)2|tx~+(1t)y~|2=t(ψ~(x~)+K(ξ~)2|x~|2)+(1t)(ψ~(y~)+K(ξ~)2|y~|2).\begin{split}&\widetilde{\psi}({t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}})+\frac{K(\widetilde{\xi})}{2}|{t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}}|^{2}\\ \leq&\,t\widetilde{\psi}(\widetilde{x}^{\prime})+(1-t)\widetilde{\psi}(\widetilde{y}^{\prime})+\frac{t(1-t)K(\widetilde{\xi})}{2}|\widetilde{x}^{\prime}-\widetilde{y}^{\prime}|^{2}+\frac{K(\widetilde{\xi})}{2}|{t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}}|^{2}\\ =&\,t\bigg{(}\widetilde{\psi}(\widetilde{x}^{\prime})+\frac{K(\widetilde{\xi})}{2}|\widetilde{x}^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\widetilde{\psi}(\widetilde{y}^{\prime})+\frac{K(\widetilde{\xi})}{2}|\widetilde{y}^{\prime}|^{2}\bigg{)}.\end{split}

Hence, KK is uniformly bounded on Ω{(x,ρ(x))||x|<δ2}\partial\Omega\cap\{(x^{\prime},\rho(x^{\prime}))|\,|x^{\prime}|<\frac{\delta}{2}\} and thus is bounded on Ω\partial\Omega by a finite cover argument.

We proceed to prove that pp is bounded on Ω\partial\Omega. Since ψ(x)+K(ξ)2|x|2\psi(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2} is convex in Bδ(0)B^{\prime}_{\delta}(0^{\prime}), we have for x=δp(ξ)2|p(ξ)|Bδ(0)x^{\prime}=\frac{\delta p(\xi)}{2|p(\xi)|}\in B^{\prime}_{\delta}(0^{\prime}),

|p(ξ)|2δ(ψ(x)ψ(0)+K(ξ)2|x|2)2δ(2supΩ|φ|+δ2K(ξ)8)C,|p(\xi)|\leq\frac{2}{\delta}\bigg{(}\psi(x^{\prime})-\psi(0^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2}\bigg{)}\leq\frac{2}{\delta}\bigg{(}2\sup_{\partial\Omega}|\varphi|+\frac{\delta^{2}K(\xi)}{8}\bigg{)}\leq C,

where CC is a constant independent of ξ\xi. This completes the proof. ∎

Summing up, we now have all ingredients to present the proof of Theorem 1.5. We start the proof by proving a special and simple case of Theorem 1.5 where A=IA=I and b=0b=0.

Proposition 6.3.

Let Ω\Omega be a domain of n\mathbb{R}^{n} satisfying a uniform enclosing sphere condition, n3n\geq 3, ΩC1\partial\Omega\in C^{1}. Let φ\varphi be semi-convex with respect to Ω\partial\Omega. Then there exists some constant cc_{*}, such that

{det(D2u)=1innΩ¯,u=φonΩ,lim|x|(u(x)(12|x|2+c))=0\begin{cases}\det(D^{2}u)=1\quad\text{in}~{}\mathbb{R}^{n}\setminus\overline{\Omega},\\ u=\varphi\quad\text{on}~{}\partial\Omega,\\ \lim_{|x|\to\infty}\Big{(}u(x)-\Big{(}\frac{1}{2}|x|^{2}+c\Big{)}\Big{)}=0\end{cases} (6.8)

has a viscosity solution in C0(nΩ)C^{0}(\mathbb{R}^{n}\setminus\Omega) if and only if ccc\geq c_{*}, where cc_{*} depends only on nn, Ω\Omega and φ\varphi.

Proof.

We divide the proof into four steps.

Step 1. There is a constant c1c_{1} such that for cc1c\geq c_{1}, (1.4) has a viscosity solution. Let

u(x)=sup{v(x)|v𝒮cφ}for xnΩ¯.u(x)=\sup\{v(x)|\ v\in\mathcal{S}_{c}^{\varphi}\}\quad\text{for }x\in\mathbb{R}^{n}\setminus\overline{\Omega}. (6.9)

By Propositions 3.6 and 4.1, uu is a viscosity solution of (1.3) in nD¯\mathbb{R}^{n}\setminus\overline{D} and approaches 12|x|2+c\frac{1}{2}|x|^{2}+c at infinity when cc satisfies (4.1), where we used the fact that φ\varphi is bounded on Ω\partial\Omega. By the uniform enclosing sphere condition and Lemma 6.2, r(ξ)r(\xi), K(ξ)K(\xi) and |p(ξ)||p(\xi)| are bounded on Ω\partial\Omega. By Lemma 6.1, δ(ξ)\delta(\xi) has a positive lower bound on Ω\partial\Omega. Note that Ω\Omega satisfies a enclosing sphere condition on Ω\partial\Omega and thus is convex. Therefore, Proposition 5.2 yields that there exists a constant c0c_{0}, such that for every c>c0c>c_{0}, uC0(nΩ)u\in C^{0}(\mathbb{R}^{n}\setminus\Omega) and u=φu=\varphi on Ω\partial\Omega. Consequently, (1.4) has a viscosity solution for cc1c\geq c_{1} with c1c_{1} sufficiently large.

We can find the sharp cc_{*} almost without change as in [22, Theorem 1.2], so we briefly sketch the rest of the proof and omit the details.

Step 2. There is a constant c2c_{2} such that for c<c2c<c_{2}, there is no viscosity subsolution u¯\underline{u} of (1.3) in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega} satisfying

{u¯=φon Ω,lim|x|(u¯(x)(12|x|2+c))=0.\begin{cases}\underline{u}=\varphi\quad\mbox{on }\partial\Omega,\\ \lim_{|x|\to\infty}\left(\underline{u}(x)-\left(\frac{1}{2}|x|^{2}+c\right)\right)=0.\end{cases} (6.10)

Precisely, here c2c_{2} depends only on nn, the diameter of Ω\Omega and infΩφ\inf_{\partial\Omega}\varphi, but not depends on the C2C^{2} regularity of Ω\Omega and φ\varphi.

Step 3. If (1.4) has a viscosity solution uc3u_{c_{3}} with c=c3c=c_{3} (c3<c1)(c_{3}<c_{1}), then (1.4) has a viscosity solution for all c4(c3,c1)c_{4}\in(c_{3},c_{1}). When proving that there is a viscosity solution of

{det(D2u)=1in nΩ¯,u=φon Ω\begin{cases}\det(D^{2}u)=1&\mbox{in }\mathbb{R}^{n}\setminus\overline{\Omega},\\ u=\varphi&\mbox{on }\partial\Omega\\ \end{cases}

provided that there is a viscosity subsolution u¯c4\underline{u}_{c_{4}} of (1.3) in nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega} satisfying (6.10) with c=c4c=c_{4}, the proof is slightly different from [22] due to the weaker condition on Ω\Omega. Indeed, let uu be as in (6.9) with c=c4c=c_{4}. Then 𝒮c4φ\mathcal{S}_{c_{4}}^{\varphi} is nonempty and uuc3u\leq u_{c_{3}}. Then uu satisfies the equation due to Steps 2-3 of the proof of Proposition 3.6. Fix R>0R>0 such that ΩBR(0)\Omega\subset\subset B_{R}(0). Recalling that φC0(Ω)\varphi\in C^{0}(\partial\Omega), Ω\Omega is convex and BR(0)Ω¯B_{R}(0)\setminus\overline{\Omega} satisfies the exterior cone condition. By Proposition B, there exists hC0(BR(0)Ω¯)h\in C^{0}(\overline{B_{R}(0)\setminus\Omega}) satisfying

{Δh=0in BR(0)Ω¯,h=φon Ω,h=maxBR(0)uon BR(0).\begin{cases}\Delta h=0&\mbox{in }B_{R}(0)\setminus\overline{\Omega},\\ h=\varphi&\mbox{on }\partial\Omega,\\ h=\max_{\partial B_{R}(0)}u&\mbox{on }\partial B_{R}(0).\end{cases}

The comparison principle implies that uhu\leq h in BR(0)Ω¯\overline{B_{R}(0)\setminus\Omega}. It follows that

φ(x0)=u¯c4(x0)limnΩ¯xx0u(x)h(x0)=φ(x0).\varphi(x_{0})=\underline{u}_{c_{4}}(x_{0})\leq\lim_{\mathbb{R}^{n}\setminus\overline{\Omega}\ni x\to x_{0}}u(x)\leq h(x_{0})=\varphi(x_{0}).

Step 4. The sharp constant cc_{*} is determined by

c=inf{c|(1.4)has a viscosity solution}.c_{*}=\inf\{c\in\mathbb{R}|\,\eqref{eq:DiriProb}~{}\text{has~{}a~{}viscosity~{}solution}\}.

Now, we give the proof of Theorem 1.5.

Proof of Theorem 1.5.

For A𝒜A\in\mathcal{A}, there exists a n×nn\times n orthogonal matrix P1P_{1} such that A=P1P2P2TP1TA=P_{1}P_{2}P_{2}^{T}P_{1}^{T}, where P2=diag(λ1(A),,λn(A))P_{2}=\text{diag}(\sqrt{\lambda_{1}(A)},\cdots,\sqrt{\lambda_{n}(A)}) and λ1(A),,λn(A)\lambda_{1}(A),\cdots,\lambda_{n}(A) are the eigenvalues of AA. Denote Q=P1P2Q=P_{1}P_{2}. Let

x~=xQ,Ω~={xQ|xΩ},\widetilde{x}=xQ,\quad\widetilde{\Omega}=\{xQ|\,x\in\Omega\},

and

φ~(x~)=φ(x)bx=φ(x~Q1)b(x~Q1).\widetilde{\varphi}(\widetilde{x})=\varphi(x)-b\cdot x=\varphi(\widetilde{x}Q^{-1})-b\cdot(\widetilde{x}Q^{-1}).

If Ω~\widetilde{\Omega} and φ~\widetilde{\varphi} satisfy the assumptions in Proposition 6.3, then we conclude that there is cc_{*} depending only on nn, Ω~\widetilde{\Omega}, φ~\widetilde{\varphi} such that there a viscosity solution u~C0(nΩ~)\widetilde{u}\in C^{0}(\mathbb{R}^{n}\setminus\widetilde{\Omega}) of

{det(D2u~)=1innΩ~¯,u~=φ~onΩ~,lim|x~|(u~(x~)(12|x~|2+c))=0\begin{cases}\det(D^{2}\widetilde{u})=1\quad\text{in}~{}\mathbb{R}^{n}\setminus\overline{\widetilde{\Omega}},\\ \widetilde{u}=\widetilde{\varphi}\quad\text{on}~{}\partial\widetilde{\Omega},\\ \lim_{|\widetilde{x}|\to\infty}\Big{(}\widetilde{u}(\widetilde{x})-\Big{(}\frac{1}{2}|\widetilde{x}|^{2}+c\Big{)}\Big{)}=0\end{cases} (6.11)

if and only if ccc\geq c_{*}. Let

u(x)=u~(x~)+bx=u~(xQ)+bx.u(x)=\widetilde{u}(\widetilde{x})+b\cdot x=\widetilde{u}(xQ)+b\cdot x.

Then direct calculation shows that uu is a viscosity solution of (1.4) in Theorem 1.5 when ccc\geq c_{*}. While for c<cc<c_{*}, if (1.4) has a viscosity solution uu, then

u~(x~)=u(x)bx\widetilde{u}(\widetilde{x})=u(x)-b\cdot x

is a viscosity solution of (6.11), which is a contradiction! Hence, we establish Theorem 1.5.

It remains to prove that Ω~\widetilde{\Omega} satisfies a uniform enclosing sphere condition, Ω~C1\partial\widetilde{\Omega}\in C^{1}, and φ~\widetilde{\varphi} is semi-convex with respect to Ω~\partial\widetilde{\Omega}. Since Ω\partial\Omega is C1C^{1}, so is Ω~\partial\widetilde{\Omega}. Denote by Br(ξ)(y(ξ))\partial B_{r(\xi)}(y(\xi)) an enclosing sphere of Ω\Omega at ξ\xi and r=maxξΩr(ξ)r=\max_{\xi\in\partial\Omega}r(\xi). Then Br(y(ξ))\partial B_{r}(y(\xi)) is also an enclosing sphere of Ω\Omega at ξ\xi. Denote

Eξ={xQ|xBr(y(ξ))}.E_{\xi}=\{xQ|\,x\in B_{r}(y(\xi))\}.

Then EξE_{\xi} is an ellipsoid and

ξQΩ~EξandΩ~Eξ.\xi Q\in\partial\widetilde{\Omega}\cap\partial E_{\xi}\quad\text{and}\quad\widetilde{\Omega}\subset E_{\xi}.

Thus Ω~\widetilde{\Omega} satisfies an enclosing sphere condition with a uniform radius. Hence, Ω~\widetilde{\Omega} satisfies a uniform enclosing sphere condition.

We continue to prove that φ~\widetilde{\varphi} is semi-convex with respect to Ω~\partial\widetilde{\Omega}. Fix ξΩ\xi\in\partial\Omega and denote ξ~=ξQ\widetilde{\xi}=\xi Q. Without losing the generality, we may assume the coordinate system in n\mathbb{R}^{n} is just the local coordinate system at ξΩ\xi\in\partial\Omega. Ω\partial\Omega can be locally represented by the graph of

xn=ρ(x)for|x|<δ(ξ),x_{n}=\rho(x^{\prime})\quad\text{for}~{}|x^{\prime}|<\delta(\xi),

for some δ(ξ)>0\delta(\xi)>0. Let ψ(x)=φ(x,ρ(x))\psi(x^{\prime})=\varphi(x^{\prime},\rho(x^{\prime})). Since φ\varphi is semi-convex with respect to Ω\partial\Omega at ξ\xi, there exists K(ξ)>0K(\xi)>0 such that for x,yBδ(ξ)(0)x^{\prime},y^{\prime}\in B^{\prime}_{\delta(\xi)}(0^{\prime}) and 0<t<10<t<1,

ψ(tx+(1t)y)+K(ξ)2|tx+(1t)y|2t(ψ(x)+K(ξ)2|x|2)+(1t)(ψ(y)+K(ξ)2|y|2).\begin{split}&\psi(tx^{\prime}+(1-t)y^{\prime})+\frac{K(\xi)}{2}|tx^{\prime}+(1-t)y^{\prime}|^{2}\\ \leq&t\bigg{(}\psi(x^{\prime})+\frac{K(\xi)}{2}|x^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\psi(y^{\prime})+\frac{K(\xi)}{2}|y^{\prime}|^{2}\bigg{)}.\end{split} (6.12)

Suppose that Ω~\partial\widetilde{\Omega} can be locally represented by the graph of

x~n=ρ~(x~)for|x~|<δ~(ξ~),\widetilde{x}_{n}=\widetilde{\rho}(\widetilde{x}^{\prime})\quad\text{for}~{}|\widetilde{x}^{\prime}|<\widetilde{\delta}(\widetilde{\xi}),

for some δ~(ξ~)>0\widetilde{\delta}(\widetilde{\xi})>0. We may assume

δ~(ξ~)min1inλi(A)1+C02δ(ξ),\widetilde{\delta}(\widetilde{\xi})\leq\sqrt{\frac{\min_{1\leq i\leq n}\lambda_{i}(A)}{1+C_{0}^{2}}}\delta(\xi),

where C0C_{0} is the Lipschitz norm of Ω~\partial\widetilde{\Omega} depending only on Ω\partial\Omega and AA. Then for x~=(x~,ρ~(x~))Ω~\widetilde{x}=(\widetilde{x}^{\prime},\widetilde{\rho}(\widetilde{x}^{\prime}))\in\partial\widetilde{\Omega} with |x~|<δ~(ξ~)|\widetilde{x}^{\prime}|<\widetilde{\delta}(\widetilde{\xi}), we have for x=x~Q1x=\widetilde{x}Q^{-1},

|x||x~Q1|max1inλi(Q1(Q1)T)|x~|max1inλi(A1)1+C02|x~|<δ(ξ).\begin{split}|x^{\prime}|&\leq|\widetilde{x}Q^{-1}|\leq\sqrt{\max_{1\leq i\leq n}\lambda_{i}(Q^{-1}(Q^{-1})^{T})}|\widetilde{x}|\\ &\leq\sqrt{\max_{1\leq i\leq n}\lambda_{i}(A^{-1})}\sqrt{1+C_{0}^{2}}|\widetilde{x}^{\prime}|<\delta(\xi).\end{split} (6.13)

Thus we get for x~=(x~,ρ~(x~))Ω~\widetilde{x}=(\widetilde{x}^{\prime},\widetilde{\rho}(\widetilde{x}^{\prime}))\in\partial\widetilde{\Omega} with |x~|<δ~(ξ~)|\widetilde{x}^{\prime}|<\widetilde{\delta}(\widetilde{\xi}),

ψ~(x~):=φ~(x~,ρ~(x~))=φ~(x~)=φ(x)bx=φ((x~Q1),ρ((x~Q1)))b(x~Q1)=ψ((x~Q1))b(x~Q1).\begin{split}\widetilde{\psi}(\widetilde{x}^{\prime}):&=\widetilde{\varphi}(\widetilde{x}^{\prime},\widetilde{\rho}(\widetilde{x}^{\prime}))=\widetilde{\varphi}(\widetilde{x})=\varphi(x)-b\cdot x\\ &=\varphi((\widetilde{x}Q^{-1})^{\prime},\rho((\widetilde{x}Q^{-1})^{\prime}))-b\cdot(\widetilde{x}Q^{-1})\\ &=\psi((\widetilde{x}Q^{-1})^{\prime})-b\cdot(\widetilde{x}Q^{-1}).\end{split}

Replacing xx^{\prime} and yy^{\prime} in (6.12) by (x~Q1)(\widetilde{x}Q^{-1})^{\prime} and (y~Q1)(\widetilde{y}Q^{-1})^{\prime} respectively, together with (6.13), we obtain

ψ~(tx~+(1t)y~)=ψ((t(x~Q1)+(1t)(y~Q1))b(tx~Q1+(1t)y~Q1)t(ψ((x~Q1))+K(ξ)2|(x~Q1)|2)+(1t)(ψ((y~Q1))+K(ξ)2|(y~Q1)|2)b(tx~Q1+(1t)y~Q1)K(ξ)2|t(x~Q1)+(1t)(y~Q1)|2=t(ψ((x~Q1))b(x~Q1))+(1t)(ψ(y~Q1)b(y~Q1))+t(1t)K(ξ)2|(x~Q1)(y~Q1)|2=tψ~(x~)+(1t)ψ~(y~)+t(1t)K(ξ)2|((x~y~)Q1)|2.\begin{split}&\,\widetilde{\psi}\left(t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}\right)\\ =&\,\psi((t(\widetilde{x}Q^{-1})^{\prime}+(1-t)(\widetilde{y}Q^{-1})^{\prime})-b\cdot(t\widetilde{x}Q^{-1}+(1-t)\widetilde{y}Q^{-1})\\ \leq&\,t\bigg{(}\psi((\widetilde{x}Q^{-1})^{\prime})+\frac{K(\xi)}{2}|(\widetilde{x}Q^{-1})^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\psi((\widetilde{y}Q^{-1})^{\prime})+\frac{K(\xi)}{2}|(\widetilde{y}Q^{-1})^{\prime}|^{2}\bigg{)}\\ &-b\cdot(t\widetilde{x}Q^{-1}+(1-t)\widetilde{y}Q^{-1})-\frac{K(\xi)}{2}\left|t(\widetilde{x}Q^{-1})^{\prime}+(1-t)(\widetilde{y}Q^{-1})^{\prime}\right|^{2}\\ =&\,t(\psi((\widetilde{x}Q^{-1})^{\prime})-b\cdot(\widetilde{x}Q^{-1}))+(1-t)(\psi(\widetilde{y}Q^{-1})^{\prime}-b\cdot(\widetilde{y}Q^{-1}))\\ &\,+\frac{t(1-t)K(\xi)}{2}\left|(\widetilde{x}Q^{-1})^{\prime}-(\widetilde{y}Q^{-1})^{\prime}\right|^{2}\\ =&\,t\widetilde{\psi}(\widetilde{x}^{\prime})+(1-t)\widetilde{\psi}(\widetilde{y}^{\prime})+\frac{t(1-t)K(\xi)}{2}\left|((\widetilde{x}-\widetilde{y})Q^{-1})^{\prime}\right|^{2}.\end{split}

Taking

K~(ξ~)=1+C02min1inλi(A)K(ξ),\widetilde{K}(\widetilde{\xi})=\frac{1+C_{0}^{2}}{\min_{1\leq i\leq n}\lambda_{i}(A)}K(\xi),

we get

ψ~(tx~+(1t)y~)+K~(ξ~)2|tx~+(1t)y~|2t(ψ~(x~)+K~(ξ~)2|x~|2)+(1t)(ψ~(y~)+K~(ξ~)2|y~|2)+t(1t)K(ξ)2|((x~y~)Q1)|2t(1t)K~(ξ~)2|x~y~|2t(ψ~(x~)+K~(ξ~)2|x~|2)+(1t)(ψ~(y~)+K~(ξ~)2|y~|2),\begin{split}&\,\widetilde{\psi}\left(t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}\right)+\frac{\widetilde{K}(\widetilde{\xi})}{2}\left|t\widetilde{x}^{\prime}+(1-t)\widetilde{y}^{\prime}\right|^{2}\\ \leq&\,t\bigg{(}\widetilde{\psi}(\widetilde{x}^{\prime})+\frac{\widetilde{K}(\widetilde{\xi})}{2}|\widetilde{x}^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\widetilde{\psi}(\widetilde{y}^{\prime})+\frac{\widetilde{K}(\widetilde{\xi})}{2}|\widetilde{y}^{\prime}|^{2}\bigg{)}\\ &\,+\frac{t(1-t)K(\xi)}{2}\left|((\widetilde{x}-\widetilde{y})Q^{-1})^{\prime}\right|^{2}-\frac{t(1-t)\widetilde{K}(\widetilde{\xi})}{2}\left|\widetilde{x}^{\prime}-\widetilde{y}^{\prime}\right|^{2}\\ \leq&\,t\bigg{(}\widetilde{\psi}(\widetilde{x}^{\prime})+\frac{\widetilde{K}(\widetilde{\xi})}{2}|\widetilde{x}^{\prime}|^{2}\bigg{)}+(1-t)\bigg{(}\widetilde{\psi}(\widetilde{y}^{\prime})+\frac{\widetilde{K}(\widetilde{\xi})}{2}|\widetilde{y}^{\prime}|^{2}\bigg{)},\end{split}

where we used

|((x~y~)Q1)|max1inλi(A1)1+C02|x~y~||((\widetilde{x}-\widetilde{y})Q^{-1})^{\prime}|\leq\sqrt{\max_{1\leq i\leq n}\lambda_{i}(A^{-1})}\sqrt{1+C_{0}^{2}}|\widetilde{x}^{\prime}-\widetilde{y}^{\prime}|

in the last “\leq” as in (6.13). That is, φ~\widetilde{\varphi} is semi-convex with respect to Ω~\partial\widetilde{\Omega} at ξ~\widetilde{\xi}. This finishes the proof of Theorem 1.5. ∎

Remark 6.4.

From the proof of Proposition 6.3 and Theorem 1.5, we see that Theorem 1.5 actually holds as long as Ω\Omega satisfies (H) and a uniform enclosing sphere condition. We note that there exist bounded and convex domains but not satisfy (H); see Example 3 in Appendix.

Appendix

Here we prove some conclusions that are involved in the main context. We also include some examples to demonstrate that the conditions in Theorem 1.5 holds for more boundary values and domains than that in Theorem 1.1. Meanwhile, we give some examples to further understand the conditions in Theorem 1.5.

Proposition A.

Let Ω\Omega be a domain of n\mathbb{R}^{n}, ΩC2\partial\Omega\in C^{2}. If Ω\Omega satisfies a uniform enclosing sphere condition, then Ω\Omega is bounded and strictly convex. The converse is also true.

Proof.

If Ω\Omega satisfies a uniform enclosing sphere condition, then Ω\Omega is clearly bounded. Under the local coordinate system at ξΩ\xi\in\partial\Omega, there exists an enclosing sphere Br(y)\partial B_{r}(y) at ξ\xi with y=(0,r)y=(0^{\prime},r). It is clear from ΩBr(y)¯\partial\Omega\subset\overline{B_{r}(y)} that

|x|2+(ρ(x)r)2r2.|x^{\prime}|^{2}+(\rho(x^{\prime})-r)^{2}\leq r^{2}.

This gives

ρ(x)12r|x|2.\rho(x^{\prime})\geq\frac{1}{2r}|x^{\prime}|^{2}.

Since ρ(0)=0\rho(0^{\prime})=0 and Dρ(0)=0D^{\prime}\rho(0^{\prime})=0^{\prime}, we have D2ρ(0)>0D^{\prime 2}\rho(0^{\prime})>0. Hence, Ω\Omega is strictly convex.

Conversely, if Ω\Omega is bounded and strictly convex at ξΩ\xi\in\partial\Omega, there exist constants 0<ε,δ<10<\varepsilon,\delta<1 independent of ξ\xi, such that

D2ρ(0)εIn1andρ(x)ε2|x|2 for |x|<δ,D^{\prime 2}\rho(0^{\prime})\geq\varepsilon I_{n-1}\quad\text{and}\quad\rho(x^{\prime})\geq\frac{\varepsilon}{2}|x^{\prime}|^{2}\text{ for }|x^{\prime}|<\delta,

under the local coordinate system at ξ\xi, where In1I_{n-1} is the (n1)×(n1)(n-1)\times(n-1) identity matrix. Take

r=max{2ε,(diamΩ+1)2εδ2+ε4δ2}andy=(0,r).r=\max\bigg{\{}\frac{2}{\varepsilon},\frac{(\text{diam}\Omega+1)^{2}}{\varepsilon\delta^{2}}+\frac{\varepsilon}{4}\delta^{2}\bigg{\}}\quad\text{and}\quad y=(0^{\prime},r).

We claim that for any zΩz\in\partial\Omega, zBr(y)¯z\in\overline{B_{r}(y)}. Indeed, if z=(z,ρ(z))z=(z^{\prime},\rho(z^{\prime})) with |z|<δ|z^{\prime}|<\delta, then

|zy|2=|z|2+(rρ(z))2|z|2+(rε2|z|2)2=r2+(1εr)|z|2+ε24|z|4r2.\begin{split}|z-y|^{2}&=|z^{\prime}|^{2}+(r-\rho(z^{\prime}))^{2}\leq|z^{\prime}|^{2}+\left(r-\frac{\varepsilon}{2}|z^{\prime}|^{2}\right)^{2}\\ &=r^{2}+(1-\varepsilon r)|z^{\prime}|^{2}+\frac{\varepsilon^{2}}{4}|z^{\prime}|^{4}\\ &\leq r^{2}.\end{split}

If zΩ{(z,ρ(z))||z|<δ}z\in\partial\Omega\setminus\{(z^{\prime},\rho(z^{\prime}))|\,|z^{\prime}|<\delta\}, then ε2δ2znr\frac{\varepsilon}{2}\delta^{2}\leq z_{n}\leq r. It follows that

|zy|2=|z|2+(rzn)2|z|2+(rε2δ2)2=r2+(diamΩ)2εδ2r+ε24δ4r2.\begin{split}|z-y|^{2}&=|z^{\prime}|^{2}+(r-z_{n})^{2}\leq|z^{\prime}|^{2}+\left(r-\frac{\varepsilon}{2}\delta^{2}\right)^{2}\\ &=r^{2}+(\text{diam}\Omega)^{2}-\varepsilon\delta^{2}r+\frac{\varepsilon^{2}}{4}\delta^{4}\\ &\leq r^{2}.\end{split}

Hence, Ω\Omega satisfies a uniform enclosing sphere condition. ∎

Proposition B.

Let Ω\Omega be a bounded domain of n\mathbb{R}^{n} satisfying an exterior cone condition, n2n\geq 2; that is for every ξΩ\xi\in\partial\Omega, there exists a finite right circular cone 𝒞\mathcal{C}, with vertex ξ\xi, such that Ω¯𝒞¯={ξ}\overline{\Omega}\cap\overline{\mathcal{C}}=\{\xi\}. Let φC0(Ω)\varphi\in C^{0}(\partial\Omega). Then there exists a solution uC0(Ω¯)u\in C^{0}(\overline{\Omega}) of

{Δu=0in Ω,u=φon Ω.\begin{cases}\Delta u=0&\mbox{in }\Omega,\\ u=\varphi&\mbox{on }\partial\Omega.\end{cases}
Proof.

The proposition was mentioned in [14, Problem 2.12], and here we give the proof for the reader’s convenience. By [14, Theorem 2.14], it suffices to prove that for every ξΩ\xi\in\partial\Omega, there is a local barrier ww at ξΩ\xi\in\partial\Omega relative to Ω\Omega. Without losing the generality, we may assume ξ=0\xi=0 and xnx_{n}-axis is in the direction of the axis of 𝒞\mathcal{C}. For x0x\neq 0, let

r(x)=|x|andθ(x)=arccosxnr.r(x)=|x|\quad\text{and}\quad\theta(x)=\arccos\frac{x_{n}}{r}.

Take 0<r0<10<r_{0}<1 and θ0(π2,π)\theta_{0}\in(\frac{\pi}{2},\pi) such that

𝒞:={0}{xn| 0<r(x)r0,θ0θ(x)π},\mathcal{C}:=\{0\}\cup\{x\in\mathbb{R}^{n}|\,0<r(x)\leq r_{0},~{}\theta_{0}\leq\theta(x)\leq\pi\},

and

ΩBr0(0){xBr0(0)| 0θ(x)<θ0}.\Omega\cap B_{r_{0}}(0)\subset\{x\in B_{r_{0}}(0)|\,0\leq\theta(x)<\theta_{0}\}.

Consider ww given by w(0)=0w(0)=0 and

w(x)=rλf(θ)forx0,w(x)=r^{\lambda}f(\theta)\quad\text{for}~{}x\neq 0,

where the constant λ\lambda and the function ff will be chosen later. For n=2n=2, we can take 0<λπ2θ00<\lambda\leq\frac{\pi}{2\theta_{0}} and f(θ)=cos(λθ)f(\theta)=\cos(\lambda\theta). Then

Δw=rλ2(λ2f(θ)+f′′(θ))=0inΩBr0(0),\Delta w=r^{\lambda-2}(\lambda^{2}f(\theta)+f^{\prime\prime}(\theta))=0\quad\text{in}~{}\Omega\cap B_{r_{0}}(0),

and w>w(0)w>w(0) in ΩBr0(0)\Omega\cap B_{r_{0}}(0). For n3n\geq 3, direct calculation gives

θi(x)=11(xnr)2(δinrxnr2xir)=r1(δinxixnr2)cscθ,\theta_{i}(x)=\frac{-1}{\sqrt{1-(\frac{x_{n}}{r})^{2}}}\bigg{(}\frac{\delta_{in}}{r}-\frac{x_{n}}{r^{2}}\frac{x_{i}}{r}\bigg{)}=-r^{-1}\bigg{(}\delta_{in}-\frac{x_{i}x_{n}}{r^{2}}\bigg{)}\csc\theta,

and

wi(x)=λrλ2xif(θ)rλ1(δinxixnr2)f(θ)cscθ,w_{i}(x)=\lambda r^{\lambda-2}x_{i}f(\theta)-r^{\lambda-1}\bigg{(}\delta_{in}-\frac{x_{i}x_{n}}{r^{2}}\bigg{)}f^{\prime}(\theta)\csc\theta,
wii(x)=(λrλ2+λ(λ2)rλ3xirxi)f(θ)λrλ3(δinxixnr2)xif(θ)cscθ((λ1)rλ2xir(δinxixnr2)rλ1(δinxi+xnr22xixnr3xir))f(θ)cscθrλ2(δinxixnr2)2f(θ)(cscθ)2cotθ+rλ2(δinxixnr2)2f′′(θ)(cscθ)2.\begin{split}w_{ii}(x)=&\bigg{(}\lambda r^{\lambda-2}+\lambda(\lambda-2)r^{\lambda-3}\frac{x_{i}}{r}x_{i}\bigg{)}f(\theta)-\lambda r^{\lambda-3}\bigg{(}\delta_{in}-\frac{x_{i}x_{n}}{r^{2}}\bigg{)}x_{i}f^{\prime}(\theta)\csc\theta\\ &-\bigg{(}(\lambda-1)r^{\lambda-2}\frac{x_{i}}{r}\bigg{(}\delta_{in}-\frac{x_{i}x_{n}}{r^{2}}\bigg{)}-r^{\lambda-1}\bigg{(}\frac{\delta_{in}x_{i}+x_{n}}{r^{2}}-2\frac{x_{i}x_{n}}{r^{3}}\frac{x_{i}}{r}\bigg{)}\bigg{)}f^{\prime}(\theta)\csc\theta\\ &-r^{\lambda-2}\bigg{(}\delta_{in}-\frac{x_{i}x_{n}}{r^{2}}\bigg{)}^{2}f^{\prime}(\theta)(\csc\theta)^{2}\cot\theta\\ &+r^{\lambda-2}\bigg{(}\delta_{in}-\frac{x_{i}x_{n}}{r^{2}}\bigg{)}^{2}f^{\prime\prime}(\theta)(\csc\theta)^{2}.\end{split}

From this, we have

Δw=rλ2(λ(λ+n2)f(θ)+(n2)cotθf(θ)+f′′(θ))=rλ2(λ(λ+n2)f(θ)+(cscθ)n2ddθ((sinθ)n2f(θ))).\begin{split}\Delta w&=r^{\lambda-2}\bigg{(}\lambda(\lambda+n-2)f(\theta)+(n-2)\cot\theta f^{\prime}(\theta)+f^{\prime\prime}(\theta)\bigg{)}\\ &=r^{\lambda-2}\bigg{(}\lambda(\lambda+n-2)f(\theta)+(\csc\theta)^{n-2}\frac{\mathrm{d}}{\mathrm{d}\theta}\bigg{(}(\sin\theta)^{n-2}f^{\prime}(\theta)\bigg{)}\bigg{)}.\end{split}

Fix θ0<Θ0<π\theta_{0}<\Theta_{0}<\pi, let

f(θ)=θΘ0(tcsct)n2dt.f(\theta)=\int_{\theta}^{\Theta_{0}}(t\csc t)^{n-2}\mathrm{d}t.

Then

Δw=rλ2(λ(λ+n2)f(θ)(n2)θn3(cscθ)n2)rλ2(λ(λ+n2)f(θ)n2θ0)inΩBr0(0).\begin{split}\Delta w&=r^{\lambda-2}\bigg{(}\lambda(\lambda+n-2)f(\theta)-(n-2)\theta^{n-3}(\csc\theta)^{n-2}\bigg{)}\\ &\leq r^{\lambda-2}\bigg{(}\lambda(\lambda+n-2)f(\theta)-\frac{n-2}{\theta_{0}}\bigg{)}\quad\text{in}~{}\Omega\cap B_{r_{0}}(0).\end{split}

Take

λ=n22((1+4(n2)θ0f(0))121)>0.\lambda=\frac{n-2}{2}\bigg{(}\bigg{(}1+\frac{4}{(n-2)\theta_{0}f(0)}\bigg{)}^{\frac{1}{2}}-1\bigg{)}>0.

Combining f(θ)f(0)f(\theta)\leq f(0), we get

Δw0inΩBr0(0).\Delta w\leq 0\quad\text{in}~{}\Omega\cap B_{r_{0}}(0).

Clearly, w>w(0)w>w(0) in ΩBr0(0)¯{0}\overline{\Omega\cap B_{r_{0}}(0)}\setminus\{0\}. Therefore, ww is a local barrier at 0. ∎

We end this section by giving a few specific examples.

Example 1.

Let Ω\Omega be a convex domain with ΩC2\partial\Omega\in C^{2} and 0Ω0\in\partial\Omega. Then φ(x)=|x|\varphi(x)=|x| is semi-convex with respect to Ω\partial\Omega but φC2(Ω)\varphi\notin C^{2}(\partial\Omega). The definition of a C2C^{2} function on ΩC2\partial\Omega\in C^{2} can refer to [14, Chapter 6.2].

We first check that φ=|x|\varphi=|x| is not C2C^{2} at 0. Indeed, under the local coordinate system at 0, we have for i=1,,n1i=1,\cdots,n-1 and x=(0,,0,xi,0,,0)n1x^{\prime}=(0,\cdots,0,x_{i},0,\cdots,0)\in\mathbb{R}^{n-1},

limxi0+φ(x,ρ(x))φ(0,ρ(0))xi=limxi0+xi2+ρ(x)2xi1,\lim_{x_{i}\to 0^{+}}\frac{\varphi(x^{\prime},\rho(x^{\prime}))-\varphi(0^{\prime},\rho(0^{\prime}))}{x_{i}}=\lim_{x_{i}\to 0^{+}}\frac{\sqrt{x_{i}^{2}+\rho(x^{\prime})^{2}}}{x_{i}}\geq 1,
limxi0φ(x,ρ(x))φ(0,ρ(0))xi=limxi0xi2+ρ(x)2xi1.\lim_{x_{i}\to 0^{-}}\frac{\varphi(x^{\prime},\rho(x^{\prime}))-\varphi(0^{\prime},\rho(0^{\prime}))}{x_{i}}=\lim_{x_{i}\to 0^{-}}\frac{\sqrt{x_{i}^{2}+\rho(x^{\prime})^{2}}}{x_{i}}\leq-1.

Hence, φ\varphi is not C1C^{1} at 0.

On the other hand, under the local coordinate system at 0, ρ0\rho\geq 0 is convex due to the convexity of Ω\Omega. Then for ψ(x):=φ(x,ρ(x))\psi(x^{\prime}):=\varphi(x^{\prime},\rho(x^{\prime})) and x,yΩx,y\in\partial\Omega near 0, we have

(ψ(tx+(1t)y))2=(φ(tx+(1t)y,ρ(tx+(1t)y)))2=|tx+(1t)y|2+(ρ(tx+(1t)y))2|tx+(1t)y|2+(tρ(x)+(1t)ρ(y))2=|tx+(1t)y|2(t|x|+(1t)|y|)2=(tψ(x)+(1t)ψ(y))2.\begin{split}(\psi(tx^{\prime}+(1-t)y^{\prime}))^{2}=&\,(\varphi(tx^{\prime}+(1-t)y^{\prime},\rho(tx^{\prime}+(1-t)y^{\prime})))^{2}\\ =&\,|tx^{\prime}+(1-t)y^{\prime}|^{2}+(\rho(tx^{\prime}+(1-t)y^{\prime}))^{2}\\ \leq&\,|tx^{\prime}+(1-t)y^{\prime}|^{2}+(t\rho(x^{\prime})+(1-t)\rho(y^{\prime}))^{2}\\ =&\,|tx+(1-t)y|^{2}\\ \leq&\,(t|x|+(1-t)|y|)^{2}\\ =&\,(t\psi(x^{\prime})+(1-t)\psi(y^{\prime}))^{2}.\end{split}

Together with ψ0\psi\geq 0, we get

ψ(tx+(1t)y)tψ(x)+(1t)ψ(y),\psi(tx^{\prime}+(1-t)y^{\prime})\leq t\psi(x^{\prime})+(1-t)\psi(y^{\prime}),

and so ψ\psi is convex near 00^{\prime}. Hence, φ\varphi is semi-convex with respect to Ω\partial\Omega at 0. For ξΩ{0}\xi\in\partial\Omega\setminus\{0\}, φ\varphi is C2C^{2} at ξ\xi. It is obvious that φ\varphi is semi-convex with respect to Ω\partial\Omega at ξ\xi.

Example 2.

There exist domains Ω\Omega satisfying that there is an enclosing sphere at every ξΩ\xi\in\partial\Omega, but r(ξ)r(\xi) is not bounded on Ω\partial\Omega. For instance,

Ω={xn|x1,,xn1>0,|x|3<xn<|x|1/3}.\Omega=\{x\in\mathbb{R}^{n}|~{}x_{1},\cdots,x_{n-1}>0,|x^{\prime}|^{3}<x_{n}<|x^{\prime}|^{1/3}\}.
[Uncaptioned image]

Indeed, an enclosing sphere at ξ=(0,,0)\xi=(0,\cdots,0) can be

|x(1,,1)|2=n.|x-(1,\cdots,1)|^{2}=n.

An enclosing sphere at ξ=(1,,1)\xi=(1,\cdots,1) can be

|x|2=n.|x|^{2}=n.

Since Ω\partial\Omega is C2C^{2} and strictly convex except for points (0,,0)(0,\cdots,0) and (1,,1)(1,\cdots,1), there is an enclosing sphere at these boundary points due to the proof of Proposition A.

On the other hand, we will prove that the radius rr of the enclosing sphere is not bounded on Ω\partial\Omega. Denote Γ=Ω{ξn| 0<|ξ|<1,ξn=|ξ|3}\Gamma=\partial\Omega\cap\{\xi\in\mathbb{R}^{n}|\,0<|\xi^{\prime}|<1,~{}\xi_{n}=|\xi^{\prime}|^{3}\}. Direct calculation gives that the unit inner normal vector ν\nu of Ω\partial\Omega at ξ=(ξ,|ξ|3)Γ\xi=(\xi^{\prime},|\xi^{\prime}|^{3})\in\Gamma is

ν=(3|ξ|ξ,1)9|ξ|4+1.\nu=\frac{(-3|\xi^{\prime}|\xi^{\prime},1)}{\sqrt{9|\xi^{\prime}|^{4}+1}}.

Denote by Br(y)\partial B_{r}(y) an enclosing sphere at ξΓ\xi\in\Gamma. Then y=ξ+rνy=\xi+r\nu, and an enclosing sphere at ξ\xi is

|xξrν|=r.|x-\xi-r\nu|=r.

Since 0ΩBr(y)¯0\in\partial\Omega\subset\overline{B_{r}(y)}, we have

|ξrν|r.|-\xi-r\nu|\leq r.

That is |ξ|2+2rνξ0|\xi|^{2}+2r\nu\cdot\xi\leq 0. It follows that

r|ξ|22νξ=(1+|ξ|4)9|ξ4|+14|ξ|asξ0.r\geq\frac{|\xi|^{2}}{-2\nu\cdot\xi}=\frac{(1+|\xi^{\prime}|^{4})\sqrt{9|\xi^{4}|+1}}{4|\xi^{\prime}|}\to\infty\quad\text{as}~{}\xi\to 0.
Example 3.

There exist many domains Ω\Omega which satisfy (H) and a uniform enclosing sphere condition but ΩC1\partial\Omega\notin C^{1}. For instance Ω=Ω1Ω2\Omega=\Omega_{1}\cap\Omega_{2}, where

Ω1={xn||x|<1}andΩ2={xn||x|2+(xna)2<1},\Omega_{1}=\{x\in\mathbb{R}^{n}|\,|x|<1\}\quad\text{and}\quad\Omega_{2}=\{x\in\mathbb{R}^{n}|\,|x^{\prime}|^{2}+(x_{n}-a)^{2}<1\},

and |a|<2|a|<\sqrt{2}. Indeed, Ω\partial\Omega is clearly not C2C^{2} on Ω1Ω2\partial\Omega_{1}\cap\partial\Omega_{2}. On the other hand, Ω\Omega clearly satisfies (H) and a uniform enclosing sphere condition.

In addition, Ω\Omega does not satisfy (H) if |a|2|a|\geq\sqrt{2}.

Acknowledgements

The author C. Wang would like to thank Professor Bo Wang for his helpful suggestions in preliminary discussions.

References

  • [1] A.D. Aleksandrov, Dirichlet’s problem for the equation Detzij=φ(z1,,zn,z,x1,,\|z_{ij}\|=\varphi(z_{1},\cdots,z_{n},z,x_{1},\cdots, xn)x_{n}). I, (Russian) Vestnik Leningrad. Univ. Ser. Mat. Meh. Astr. 13 (1958) 5–24.
  • [2] J. Bao, J. Chen, B. Guan, M. Ji, Liouville property and regularity of a Hessian quotient equation, Amer. J. Math. 125 (2003) 301–316.
  • [3] J. Bao, H. Li, Y.Y. Li, On the exterior Dirichlet problem for Hessian equations, Trans. Amer. Math. Soc. 366 (2014) 6183–6200.
  • [4] L.A. Caffarelli, Topics in PDEs: The Monge–Ampère Equation, Courant Institute, New York University, Graduate Course (1995).
  • [5] L.A. Caffarelli, X. Cabré, Fully nonlinear elliptic equations. In American Mathematical Society Colloquium Publications, vol. 43. American Mathematical Society, Providence, RI (1995).
  • [6] L.A. Caffarelli, Y.Y. Li, An extension to a theorem of Jörgens, Calabi, and Pogorelov, Comm. Pure Appl. Math. 56 (2003) 549–583.
  • [7] L.A. Caffarelli, L. Nirenberg, J. Spruck, The Dirichlet problem for nonlinear second-order elliptic equations. I. Monge–Ampère equation, Comm. Pure Appl. Math. 37 (1984) 369–402.
  • [8] E. Calabi, Improper affine hyperspheres of convex type and a generalization of a theorem by K. Jörgens, Mich. Math. J. 5 (1958) 105–126.
  • [9] S.Y. Cheng, S.T. Yau, On the regularity of the Monge–Ampère equation det(2u/xixj)=F(x,u)\det(\partial^{2}u/\partial x_{i}\partial x_{j})=F(x,u), Comm. Pure Appl. Math. 30 (1977) 41–68.
  • [10] S.Y. Cheng, S.T. Yau, Complete affine hypersurfaces. I. The completeness of affine metrics, Comm. Pure Appl. Math. 39 (1986) 839–866.
  • [11] M. Crandall, H. Ishii, P.L. Lions, User’s guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc. 27 (1992) 1–67.
  • [12] G. De Philippis, A. Figalli, The Monge–Ampère equation and its link to optimal transportation, Bull. Amer. Math. Soc. (N.S.) 51 (2014) 527–580.
  • [13] L.C. Evans, R.F. Gariepy, Measure Theory and Fine Properties of Functions. CRC Press, Boca Raton (1992).
  • [14] D. Gilbarg, N.S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics in Mathematics. Springer, Berlin (2001).
  • [15] H. Ishii, On uniqueness and existence of viscosity solutions of fully nonlinear second-order elliptic PDE’s, Comm. Pure Appl. Math. 1 (1989) 15–45.
  • [16] T. Jiang, H. Li, X. Li, On the exterior Dirichlet problem for a class of fully nonlinear elliptic equations, Calc. Var. Partial Differential Equations 60 (2021), Paper No. 17, 20 pp.
  • [17] K. Jörgens, Über die Lösungen der Differentialgleichung rts2=1rt-s^{2}=1, Math. Ann. 127 (1954) 130–134.
  • [18] N.V. Krylov, Boundedly inhomogeneous elliptic and parabolic equations in a domain, (Russian) Izv. Akad. Nauk SSSR Ser. Mat. 47 (1983) 75–108; English translation: Math. USSR-Izv. 22 (1984) 67–98.
  • [19] D. Li, Z. Li, On the exterior Dirichlet problem for Hessian quotient equations, J. Differential Equations 264 (2018) 6633–6662.
  • [20] H. Li, J. Bao, The exterior Dirichlet problem for fully nonlinear elliptic equations related to the eigenvalues of the Hessian, J. Differential Equations 256 (2014) 2480–2501.
  • [21] M. Li, C. Ren, Z. Wang, An interior estimate for convex solutions and a rigidity theorem, J. Funct. Anal. 270 (2016) 2691–2714.
  • [22] Y.Y. Li, S. Lu, Existence and nonexistence to exterior Dirichlet problem for Monge–Ampère equation, Calc. Var. Partial Differential Equations 57 (2018), Paper No. 161, 17 pp.
  • [23] Z. Li, On the exterior Dirichlet problem for special Lagrangian equations, Trans. Amer. Math. Soc. 372 (2019) 889–924.
  • [24] N. Meyers, J. Serrin, The exterior Dirichlet problem for second order elliptic partial differential equations, J. Math. Mech. 9 (1960) 513–538.
  • [25] L. Nirenberg, The Weyl and Minkowski problems in differential geometry in the large, Comm. Pure Appl. Math. 6 (1953) 337–394.
  • [26] A.V. Pogorelov, The Dirichlet problem for the multidimensional analogue of the Monge–Ampère equation, (Russian) Dokl. Akad. Nauk SSSR 201 (1971) 790–793.
  • [27] A.V. Pogorelov, On the improper convex affine hyperspheres, Geom. Dedicata 1 (1972) 33–46.
  • [28] A.V. Pogorelov, The Minkowski Multidimensional Problem, Wiley, New York (1978).
  • [29] J. Rauch, B.A. Taylor, The Dirichlet problem for the multidimensional Monge–Ampère equation, Rocky Mountain J. Math. 7 (1977) 345–364.
  • [30] N.S. Trudinger, X.-J. Wang, The Monge–Ampère equation and its applications. Handbook of geometric analysis. No. 1, 467–524, Adv. Lect. Math. (ALM), 7, Int. Press, Somerville, MA, 2008.
  • [31] J. Urbas, The equation of prescribed Gauss curvature without boundary conditions, J. Differential Geom. 20 (1984) 311–327.
  • [32] J. Urbas, Complete noncompact self-similar solutions of Gauss curvature flows. I. Positive powers, Math. Ann. 311 (1998) 251–274.
  • [33] C. Wang, J, Bao, Liouville property and existence of entire solutions of Hessian equations, Nonlinear Anal. 223 (2022), Paper No. 113020, 18 pp.
  • [34] M. Yan, Extension of convex function, J. Convex Anal. 21 (2014) 965–987.