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

The area minimizing problem in conformal cones: II

Qiang Gao, Hengyu Zhou Department of Mathematics, Sun Yat-Sen University, 510275, Guangzhou, P. R. China [email protected] College of Mathematics and Statistics, Chongqing University, Huxi Campus, Chongqing, 401331, P. R. China Chongqing Key Laboratory of Analytic Mathematics and Applications, Chongqing University, Huxi Campus, Chongqing, 401331, P. R. China [email protected]
Abstract.

In this paper we continue to study the connection among the area minimizing problem, certain area functional and the Dirichlet problem of minimal surface equations in a class of conformal cones with a similar motivation from [14]. These cones are certain generalizations of hyperbolic spaces. We describe the structure of area minimizing nn-integer multiplicity currents in bounded C2C^{2} conformal cones with prescribed C1C^{1} graphical boundary via a minimizing problem of these area functionals. As an application we solve the corresponding Dirichlet problem of minimal surface equations under a mean convex type assumption. We also extend the existence and uniqueness of a local area minimizing integer multiplicity current with star-shaped infinity boundary in hyperbolic spaces into a large class of complete conformal manifolds.

2010 Mathematics Subject Classification:
Primary 49Q20: Secondary 53C21. 53A10,. 35A01. 35J25

1. Introduction

In this paper we continue to study the area minimizing problem with prescribed boundary in a class of conformal cones similar to [14]. A conformal cone in this paper is defined as follows.

Definition 1.1.

Let NN be a nn-dimensional Riemannian manifold with a metric σ\sigma, let \mathbb{R} be the real line with the metric dr2dr^{2} and ϕ(x)\phi(x) be a C3C^{3} positive function on NN. In this paper we call

(1.1) Mϕ:={N×,ϕ2(x)(σ+dr2)}M_{\phi}:=\{N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R},\phi^{2}(x)(\sigma+dr^{2})\}

as a conformal product manifold. Let Ω\Omega be a C2C^{2} bounded domain in NN. We refer Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} in MϕM_{\phi} as a conformal cone, denoted by QϕQ_{\phi}.

Let ψ(x)\psi(x) be a C1C^{1} function on Ω\partial\Omega and Γ\Gamma be its graph in Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. The area minimizing problem in a conformal cone QϕQ_{\phi} is to find a nn-integer multiplicity current in Q¯ϕ\bar{Q}_{\phi}, the closure of QϕQ_{\phi}, to realize

(1.2) min{𝕄(T)T𝒢,T=Γ}\min\{\mathbb{M}(T)\mid T\in\mathcal{G},\partial T=\Gamma\}

where 𝕄\mathbb{M} is the mass of integer multiplicity currents in MϕM_{\phi}, and 𝒢\mathcal{G} denotes the set of n-integer multiplicity currents with compact support in Q¯ϕ\bar{Q}_{\phi}, i.e. for any T𝒢T\in\mathcal{G}, its support spt(T)spt(T) is contained in Ω¯×[a,b]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b] for some finite numbers a<ba<b. See subsection 4.1 for more details.
The main reason to study the conformal product manifold MϕM_{\phi} in Definition 1.1 is that hyperbolic space is a special case of MϕM_{\phi} (see remark 6.5). With this model in [22, Theorem 4.1], Hardt-Lin showed that there is a unique local area minimizing nn-integer multiplicity current for any prescribed infinity C1C^{1} star-shaped boundary. Moreover it is a radial minimal graph over S+nS^{n}_{+} in hyperbolic spaces. On the other hand in [25, Section 4.1], Lin described the solution to the area minimizing problem (1.2) with C1C^{1} graphical boundary in a bounded C2C^{2} cylinder via a minimizing problem of an area functional of bounded variation (BV) functions. Motivated by Lin’s idea in [14] we studied the area minimizing problem (1.2) in a conformal product manifold Mh={N×,h2(r)(σ+dr2)}M_{h}=\{N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R},h^{2}(r)(\sigma+dr^{2})\}. Note that there h(r)h(r) only depends on rr\in\mathbb{R}. Based on these preceding results, it is natural to consider the corresponding area minimizing problem (1.2) in conformal cones of MϕM_{\phi}. In particular it is desirable to know how many phenomena in hyperbolic sapce does not really rely on the hyperbolic structures. We refer readers to [2] for some history remarks on such kinds of area minimizing problems.
For any open set WW we denote the set of all bounded variation functions on WW by BV(W)BV(W). A key concept for our study of the problem (1.2) is an area functional in BV(W)BV(W) defined as follows.

(1.3) 𝔉ϕ(u,W)=sup{Ωϕn(x)h+udiv(ϕn(x)X)dvolhC0(W),XT0(W) and h2+X,X1}\begin{split}\mathfrak{F}_{\phi}(u,W)&=\sup\{\int_{\Omega}\phi^{n}(x)h+u{{div}}(\phi^{n}(x)X)dvol\mid h\in C_{0}(W),\\ &X\in T_{0}(W)\text{ and }h^{2}+{\left<X,X\right>}\leq 1\}\end{split}

where dvoldvol and divdiv are the volume form and the divergence of NN respectively, C0(W)C_{0}(W) and T0(W)T_{0}(W) denote the set of smooth functions and vector fields with compact support in WW respectively. Note that when uC1(W)u\in C^{1}(W), 𝔉ϕ(u,W)\mathfrak{F}_{\phi}(u,W) is the area of the graph of u(x)u(x) in MϕM_{\phi}.
Let Ω\Omega be the C2C^{2} domain in Definition 1.1. Suppose Ω\Omega^{\prime} is a C2C^{2} domain in NN such that ΩΩ\Omega\subset\subset\Omega^{\prime}. That is, the closure of Ω\Omega is a compact set in Ω\Omega^{\prime}. Suppose ψ(x)C1(Ω\Ω)\psi(x)\in C^{1}(\Omega^{\prime}\backslash\Omega). The following minimizing problem

(1.4) min{𝔉ϕ(v,Ω)v(x)BV(Ω),v(x)=ψ(x) on Ω\Ω}\min\{\mathfrak{F}_{\phi}(v,\Omega^{\prime})\mid v(x)\in BV(\Omega^{\prime}),v(x)=\psi(x)\text{ on }\Omega^{\prime}\backslash\Omega\}

also plays an important role to solve (1.2).
In fact the key idea to solve (1.2) is to establish the connection among the problem (1.2), the area functional minimizing problem (1.4) and the Dirichlet problem of minimal surface equations in MϕM_{\phi}. This can be easily seen when Σ\Sigma is a minimal graph of u(x)u(x) in MϕM_{\phi} over Ω\Omega with C1C^{1} boundary ψ(x)\psi(x) on Ω\partial\Omega. From Theorem 2.3 one has (1) u(x)u(x) should satisfy

(1.5) div(Duω)+nDlogϕ,Duω=0 on Ωdiv(\frac{Du}{\omega})+n\langle D\log\phi,\frac{Du}{\omega}\rangle=0\text{ on }\Omega

with u(x)=ψ(x)u(x)=\psi(x) on Ω\partial\Omega; (2) the area of Σ\Sigma is less than that of any compact C2C^{2} surface SS with S\partial S as the graph of ψ(x)\psi(x) containing in Ω¯×\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}; (3) u(x)u(x) realizes the minimum of (1.4) if requiring all v(x)C2(Ω)BV(Ω)C(Ω¯)v(x)\in C^{2}(\Omega)\cap BV(\Omega^{\prime})\cap C(\bar{\Omega}).
In MϕM_{\phi} a similar connection between the problem (1.2) and the problem (1.4) is obtained in Theorem 4.1. It says that if u(x)u(x) is the solution to the problem (1.4), then T=[[U]]|Q¯ϕT=\partial[[U]]|_{\bar{Q}_{\phi}} solves the problem (1.2) in MϕM_{\phi} where UU is the subgraph of u(x)u(x) and [[U]][[U]] is the corresponding integer multiplicity current. This generalizes Lin’s result [25] into conformal cones in MϕM_{\phi}. Its proof is based on the following three observations: (1) by Theorem 3.9, 𝔉ϕ(u,.)\mathfrak{F}_{\phi}(u,.) is just the perimeter of its subgraph UU; (2) if TT solves (1.2) and Γ=(x,ψ(x))\Gamma=(x,\psi(x)) for ψ(x)C1(Ω)\psi(x)\in C^{1}(\partial\Omega), then TT is a boundary of a Caccioppoli set (see Lemma 4.6); (3) by Theorem 3.10, for any Caccioppoli set FF in QMϕQ\subset M_{\phi} with FΩ×[a,b]\partial F\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}[a,b], there is a u(x)BV(Ω)u(x)\in BV(\Omega) such that 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) is less than the perimeter of FF. Note that the observation (3) is a very general phenomenon in area-type functionals [29], [26], [30], [18],[15] and [7] (see remark 3.11). We refer it as the Miranda’s observation.
A direct application of Theorem 4.1 is the Dirichlet problem of minimal surface equations in MϕM_{\phi}. In Theorem 5.3 we show that if Ω\Omega is ϕ\phi-mean convex, then the Dirichlet problem (1.5) with continuous boundary data has a unique solution in C2(Ω)C(Ω¯)C^{2}(\Omega)\cap C(\bar{\Omega}). Here Ω\Omega is ϕ\phi-mean convex if the mean curvature of Ω\partial\Omega satisfies that HΩ+nγ,Dlogϕ0 on ΩH_{\partial\Omega}+n\langle\vec{\gamma},D\log\phi\rangle\geq 0\text{ on }\partial\Omega where γ\vec{\gamma} is the outward normal vector of Ω\partial\Omega and HΩ=div(γ)H_{\partial\Omega}=div(\vec{\gamma}). A more general form of the Dirichlet problem (1.5) was already considered by Casteras-Heinonen-Holopainen [8]. To obtain the C0C^{0} esimate they relied on a lower bound of the Ricci curvature Ω\Omega and the positive mean curvature of Ω\partial\Omega (see remark 5.4) . But out result in Theorem 5.3 is independent of the curvature of Ω\Omega and also can not be obtained by the classical continuous method in [16, Chapter 11, 17.2] (see remark 5.5). A consequence of Theorem 5.3 is that if Ω\Omega is ϕ\phi-mean convex, the area minimizing integer multiplicity current in (1.2) is unique as the graph of a C2C^{2} function to solve the Dirichlet problem (see Theorem 5.6).
At last we consider the existence and unqiueness of local area minimizing integer multiplicity current in MϕM_{\phi} with infinity boundary Γ\Gamma when ϕ(x)\phi(x) can be written as ϕ(d(x,N))\phi(d(x,\partial N)) which goes to ++\infty as d(x,N)d(x,\partial N) goes to zero in NN. Here NN is a compact Riemannian manifold with C2C^{2} boundary and dd is the distance function in NN. Let NrN_{r} be the set {xN:d(x,N)>r}\{x\in N:d(x,\partial N)>r\}. In Theorem 6.4 we obtain that if there is r1r_{1} such that for any r(0,r1)r\in(0,r_{1}) NrN_{r} is ϕ\phi-mean convex, then for any ψ(x)C(N)\psi(x)\in C(\partial N) and Γ=(x,ψ(x))\Gamma=(x,\psi(x)) there is a unique local area minimizing integer multiplicity current TT with infinity boundary Γ\Gamma. Moreover TT is a minimal graph in MϕM_{\phi} over NN. This illustrates that Theorem 4.1 in [22] does not depend on the hyperbolic structure (see remark 6.5 ). Unlike the case in hyperbolic spaces our existence in Theorem 6.4 is from the Dirichlet problem of minimal surface equations in Theorem 5.3, not from the results in [5, 6] by Anderson via geometric measure theory.
For the idea of radial graphs in hyperbolic space we refer to [19, 9, 27, 30, 36, 31, 20, 21, 26] etc. For the Dirichlet problem of minimal surface equations in Riemannian manifolds, we refer to [8], [4], [35], [1], [3] and references therein. As for the variational method to study the Dirichlet problem of minimal surface equations, we refer to [18, 17, 7], [41] and references therein.
Our paper is organized as follows. In section 2 we show three properties of C2C^{2} minimal graphs in MϕM_{\phi} and collect preliminary facts on BV functions. In section 3 we show the CC^{\infty} approximation theorem of 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) (Theorem 3.8) and the Miranda’s observation (Theorem 3.9 and Theorem 3.10). In section 4 we consider the connection between the problem (1.2) and the problem (1.4) (Theorem 4.1). In section 5 we discuss the Dirichlet problem of minimal surface equations in MϕM_{\phi} on a ϕ\phi-mean convex domain (Theorem 5.3). As an application we obtain the uniqueness of the problem (1.2) under the ϕ\phi-mean convex assumption (Theorem 5.6). In section 6 we discuss the existence and uniqueness of local area minimizing nn-integer multiplicity current with infinity graphical boundary (Theorem 6.4). In Appendix A we give a proof of the interior estimate of minimal surface equations in (1.5) (Theorem A.3) following from an idea of Eichmair [11].
The first author is supported by the National Natural Science Foundation of China, No. 11771456. The second author is supported by the National Natural Science Foundation of China, No. 11801046 and by the Fundamental Research Funds for the Central Universities, China, No.2019CDXYST0015.

2. Preliminaries

Throughout this section we assume NN is a manifold with a metric σ\sigma, ϕ(x)>0\phi(x)>0 is a positive C3C^{3} function defined on NN and MϕM_{\phi} is the conformal product manifold N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} with the metric ϕ2(x)(g+dr2)\phi^{2}(x)(g+dr^{2}). We show three properties of C2C^{2} minimal graphs in MϕM_{\phi} and collect some results on BV functions in Riemannian manifolds for later use.

2.1. Three properties

Definition 2.1.

Let SS be a C2C^{2} orientable hypersurface in a Riemannian manifold MM with a normal vector v\vec{v}. We call div(v)div(\vec{v}) as the mean curvature of SS with respect to v\vec{v}.

Let Ω\Omega be a bounded domain in NN and uC2(Ω)u\in C^{2}(\Omega). Let Σ\Sigma be the graph of u(x)u(x). Then its upward normal vector in the product manifold N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is vΣ:=rDuω\vec{v}_{\Sigma}:=\frac{\partial_{r}-Du}{\omega} where DD denotes the gradient of NN and ω=1+|Du|2\omega=\sqrt{1+|Du|^{2}}. Then we have

Lemma 2.2.

The mean curvature of Σ\Sigma with respect to its upward normal vector in MϕM_{\phi} is

(2.1) HΣ=1ϕ(x)(div(Duω)nDlogϕ,Duω)H_{\Sigma}=\frac{1}{\phi(x)}(-div(\frac{Du}{\omega})-n\langle D\log\phi,\frac{Du}{\omega}\rangle)

where ω=1+|Du|2\omega=\sqrt{1+|Du|^{2}}, nn is the dimension of NN and divdiv is the divergence of NN.

Proof.

We write the metric of MϕM_{\phi} as e2f(σ+dr2)e^{2f}(\sigma+dr^{2}) where f=logϕf=\log\phi. The upward normal vector of Σ\Sigma in MϕM_{\phi} is efvΣe^{-f}\vec{v}_{\Sigma}. By [40, Lemma 3.1], the mean curvature of Σ\Sigma in MϕM_{\phi} with respect to efvΣe^{-f}\vec{v}_{\Sigma} is

(2.2) HΣ=ef(H+ndf(vΣ))H_{\Sigma}=e^{-f}(H+ndf(\vec{v}_{\Sigma}))

where HH is the mean curvature of SS with respect to vΣ\vec{v}_{\Sigma} in the product manifold N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} and nn is the dimension of NN. With a straightforward computation HH is div(Duω)-div(\frac{Du}{\omega}) where divdiv is the divergence of NN. Putting vΣ\vec{v}_{\Sigma} and f=logϕf=\log\phi into (2.2), we obtain the conclusion. ∎

Now we summarize three properties of minimal graphs in MϕM_{\phi} as follows.

Theorem 2.3.

Suppose Σ\Sigma is a minimal graph of u(x)u(x) over Ω\Omega with C1C^{1} boundary Γ:={(x,ψ(x)):xΩ,ψ(x)C1(Ω)}\Gamma:=\{(x,\psi(x)):x\in\partial\Omega,\psi(x)\in C^{1}(\partial\Omega)\}. Then

  1. (1)

    u(x)u(x) satisfies that

    (2.3) div(Duω)nDlogϕ(x),Duω=0-div(\frac{Du}{\omega})-n\langle D\log\phi(x),\frac{Du}{\omega}\rangle=0
  2. (2)

    Let Ω\Omega^{\prime} be a domain such that ΩΩ\Omega\subset\subset\Omega^{\prime} and ψ(x)C1(Ω\Ω)\psi(x)\in C^{1}(\Omega^{\prime}\backslash\Omega). Further assuming u(x)=ψ(x)u(x)=\psi(x) on Ω\Ω\Omega^{\prime}\backslash\Omega, then u(x)u(x) realizes

    (2.4) min{𝔉ϕ(v,Ω)v(x)C2(Ω),v(x)=ψ(x) on Ω¯\Ω}\min\{\mathfrak{F}_{\phi}(v,\Omega^{\prime})\mid v(x)\in C^{2}(\Omega),v(x)=\psi(x)\text{ on }\bar{\Omega}^{\prime}\backslash\Omega\}

    where 𝔉ϕ(v,Ω)=Ωϕn(x)1+|Dv|2𝑑vol\mathfrak{F}_{\phi}(v,\Omega^{\prime})=\int_{\Omega^{\prime}}\phi^{n}(x)\sqrt{1+|Dv|^{2}}dvol and dvoldvol is the volume form of NN.

  3. (3)

    The area of Σ\Sigma achieves the minimum of the area of all C2C^{2} compact orientable hypersurface SS in Q¯ϕ:=Ω¯×\bar{Q}_{\phi}:=\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} with S=Γ\partial S=\Gamma.

Proof.

Note that the property (1) is from Lemma 2.2.
For any C2C^{2} graph S=(x,v(x))S=(x,v(x)) over Ω\Omega, its area in MϕM_{\phi} is

(2.5) Ωϕn(x)1+|Dv|2𝑑vol\int_{\Omega}\phi^{n}(x)\sqrt{1+|Dv|^{2}}dvol

where dvoldvol is the volume form on Ω\Omega. Thus the property (2) follows from the property (3). It is sufficient to show the property (3).
Recall that vΣ:=rDuω\vec{v}_{\Sigma}:=\frac{\partial_{r}-Du}{\omega}. For any tt\in\mathbb{R}, define a map Tt:N×N×T_{t}:N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}\rightarrow N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} as Tt(x,r)=(x,rt)T_{t}(x,r)=(x,r-t). We define a new vector field in Q¯ϕ\bar{Q}_{\phi} as

(2.6) X(x,u(x)t)=1ϕ(x)Tt(vΣ)X(x,u(x)-t)=\frac{1}{\phi(x)}T_{t*}(\vec{v}_{\Sigma})

where TtT_{t*} is the pushforword of TtT_{t}. Thus XX is a smooth unit vector field in the tangent bundle of Q¯ϕ\bar{Q}_{\phi}. By the definition of divergence, (see Page 423 in [23]) we define an nn-form as

(2.7) ξ=X(ϕn+1dvoldr)=XdvolMϕ\xi=X\lrcorner(\phi^{n+1}dvol\wedge dr)=X\lrcorner dvol_{M_{\phi}}

A key fact is

(2.8) dξ=divMϕ(X)dvolMϕ=HTt(Σ)dvolMϕd\xi=div_{M_{\phi}}(X)dvol_{M_{\phi}}=H_{T_{t}(\Sigma)}dvol_{M_{\phi}}

where divMϕdiv_{M_{\phi}} and dvolMϕdvol_{M_{\phi}} is the divergence and volume form of MϕM_{\phi}. Note that TtT_{t} is an isometry of MϕM_{\phi} for every tt\in\mathbb{R}. Because Σ\Sigma is minimal in MϕM_{\phi}, so is Tt(Σ)T_{t}(\Sigma). Then dξ=0d\xi=0 on the whole QϕQ_{\phi}.
Now let SS be any compact C2C^{2} orientable hypersurface in Q¯ϕ\bar{Q}_{\phi} satisfying S=Γ\partial S=\Gamma. Without loss of generality we can assume SS and Σ\Sigma bounds a domain WW in MϕM_{\phi}. Suppose that the outward normal vector of WW on Σ\Sigma and SS are denoted by nΣ\vec{n}_{\Sigma} and nS\vec{n}_{S}, the volume form of W\partial W on Σ\Sigma and SS are denoted by dvolΣdvol_{\Sigma} and dvolSdvol_{S} respectively. It’s clear that the upward normal vector of Σ\Sigma in MϕM_{\phi}, nΣ\vec{n}_{\Sigma}, is 1ϕvΣ\frac{1}{\phi}\vec{v}_{\Sigma}. Applying the divergence theorem in the domain WW, we obtain

0\displaystyle 0 =W𝑑ξ=ΣX,nΣMϕ𝑑volΣSX,nSMϕ𝑑volS\displaystyle=\int_{W}d\xi=\int_{\Sigma}{\left<X,\vec{n}_{\Sigma}\right>}_{M_{\phi}}dvol_{\Sigma}-\int_{S}{\left<X,\vec{n}_{S}\right>}_{M_{\phi}}dvol_{S}
Area(Σ)Area(S)\displaystyle\geq Area(\Sigma)-Area(S)

with equality if and only if S=ΣS=\Sigma. Thus we obtain the property (3). The proof is complete. ∎

2.2. BV functions

Now we collect some preliminary facts for BV functions in Riemannian manifolds. We refer readers to the books of Giusti[18], Evans-Gariepy[12] and the papers of McGonagle-Xiao [28], [41] etc. Recall that Ω\Omega is a bounded domain in NN and dvoldvol is the volume form of NN. Let T0ΩT_{0}\Omega be the set of smooth vector fields with compact support in Ω\Omega.
Now we define BV functions and Caccioppoli sets as follows:

Definition 2.4.

Let uL1(Ω)u\in L^{1}(\Omega). Define

(2.9) Du(Ω)=sup{Ωu𝑑iv(X)𝑑volXT0Ω and X,X1}{\left|\left|Du\right|\right|}(\Omega)=\sup\{\int_{\Omega}u{{div}}(X)dvol\mid X\in T_{0}\Omega\text{ and }{\left<X,X\right>}\leq 1\}

where dvoldvol is the volume form of Ω\Omega. If Du(Ω)<{\left|\left|Du\right|\right|}(\Omega)<\infty, we say that uu has bounded variation in Ω\Omega. The set of all functions with bounded variation is denoted by BV(Ω)BV(\Omega). If uu belongs to BV(W)BV(W) for any bounded domain WΩW\subset\subset\Omega, we say uBVloc(Ω)u\in BV_{loc}(\Omega).
Let EE be a Borel set in Ω\Omega and λE\lambda_{E} be the characteristic function of EE. If λEBVloc(Ω)\lambda_{E}\in BV_{loc}(\Omega), then EE is called as a Caccioppoli set and DλE(Ω)||D\lambda_{E}||(\Omega) is called as the perimeter of EE. In the reminder of this paper, we also write it as P(E,Ω)P(E,\Omega).

Remark 2.5.

In some settings, in order to emphasize the ambient manifold MM, we use the notation DuM(Ω){\left|\left|Du\right|\right|}_{M}(\Omega) instead of Du(Ω){\left|\left|Du\right|\right|}(\Omega).
For a Caccioppoli set EE all properties are unchanged if we make alterations of any Hausdorff measure zero set. Arguing exactly as Proposition 3.1 in [18], we can always choose a set EE^{\prime} differing by a Hausdorff measure zero set with EE and satisfying for any xEx\in\partial E^{\prime}

(2.10) 0<vol(EB(x,ρ))<vol(B(x,ρ))0<vol(E^{\prime}\cap B(x,\rho))<vol(B(x,\rho))

where ρ\rho is sufficiently small. From now on, we always assume that condition (2.10) holds for any Caccioppoli set EE.

Definition 2.6.

We say a sequence of measurable functions {uk}k=1\{u_{k}\}_{k=1}^{\infty} locally converges to uu in L1(Ω)L^{1}(\Omega) if for any open set WΩW\subset\subset\Omega we have

(2.11) limk=1W|uku|𝑑vol=0\lim_{k=1}^{\infty}\int_{W}|u_{k}-u|dvol=0
Remark 2.7.

If {uk}k=1\{u_{k}\}_{k=1}^{\infty} converges to u(x)u(x) a.e, then this sequence locally converges to u(x)u(x) in L1(Ω)L^{1}(\Omega).

By the definition in (2.9) it is easy to see that

Theorem 2.8 (Lower-semicontinuity).

Let {ui}\{u_{i}\} be a sequence of functions in BV(Ω)BV(\Omega) locally converging to a function uu in the L1(Ω)L^{1}(\Omega). Then

(2.12) Du(Ω)lim infi+Dui(Ω){\left|\left|Du\right|\right|}(\Omega)\leq\mathop{\liminf}\limits_{i\to+\infty}{\left|\left|Du_{i}\right|\right|}(\Omega)

The following CC^{\infty} approximation result is not trivial when Ω\Omega is not contained in a simply-connected domain of a Riemannian manifold. Because in this case no global symmetric mollifies exist for this domain as that in [18, page 15] for Euclidean spaces. For a complete proof, we refer to [41, section 3].

Theorem 2.9 (Theorem 3.6, [41]).

Suppose uBV(Ω)u\in BV(\Omega). Then there exists a sequence of functions {uk}k=1\{u_{k}\}_{k=1}^{\infty} in C(Ω)C^{\infty}(\Omega) such that

(2.13) limk+Ω|uuk|𝑑vol=0,and\displaystyle\mathop{\lim}\limits_{k\to+\infty}\int_{\Omega}|u-u_{k}|dvol=0,and
(2.14) limk+Duk(Ω)=Du(Ω)\displaystyle\mathop{\lim}\limits_{k\to+\infty}{\left|\left|Du_{k}\right|\right|}(\Omega)={\left|\left|Du\right|\right|}(\Omega)

As a conclusion, we can obtain the following well-known compactness result.

Theorem 2.10 (Compactness theorem).

Let ΩM\Omega\subset M be a bounded domain with Lipschitz boundary. Suppose there is a constant C>0C>0 such that for {uk}k=1BV(Ω)\{u_{k}\}_{k=1}^{\infty}\subset BV(\Omega)

(2.15) ukL1+Duk(Ω)C for any k{\left|\left|u_{k}\right|\right|}_{L^{1}}+{\left|\left|Du_{k}\right|\right|}(\Omega)\leq C\text{ for any k}

Then there is a function uBV(Ω)u\in BV(\Omega) such that there is a subsequence of {uk}k=1\{u_{k}\}_{k=1}^{\infty} converging to u(x)u(x) in L1(Ω)L^{1}(\Omega).

We can also view a BVlocBV_{loc} function as a Radon measure.

Theorem 2.11 (Theorem 2.6, [41]).

Let ΩM\Omega\subset M be a bounded domain. Suppose that uBVloc(Ω)u\in BV_{loc}(\Omega).

  1. (1)

    There exists a Radon measure |Du||Du| on Ω\Omega such that

    (2.16) Ωfd|Du|=sup{Ωu𝑑iv(X)𝑑volXT0Ω and X,Xf2}\int_{\Omega^{\prime}}fd|Du|=\sup\{\int_{\Omega^{\prime}}u{{div}}(X)dvol\mid X\in T_{0}\Omega^{\prime}\text{ and }\langle X,X\rangle\leq f^{2}\}

    for any open set ΩΩ\Omega^{\prime}\subset\subset\Omega and any nonnegative function fL1(|Du|)f\in L^{1}(|Du|).

  2. (2)

    There exists a vector field ν\vec{\nu} on Ω\Omega satisfying

    (2.17) Ωu𝑑iv(X)𝑑vol=ΩX,νd|Du|\int_{\Omega}u{{div}}(X)dvol=-\int_{\Omega}{\left<X,\vec{\nu}\right>}d|Du|

    where ν,ν=1\langle\vec{\nu},\vec{\nu}\rangle=1 |Du||Du|-a.e. for any XT0ΩX\in T_{0}\Omega.

3. The area functional

In this section we define a new area functional 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) for BV functions to generalize the area of the graph of C1C^{1} functions in MϕM_{\phi} (see (2.5)). Suppose NN is a fixed Riemannian manifold with metric σ\sigma and ϕ(x)\phi(x) is a C3C^{3} positive function on NN. Here MϕM_{\phi} is the set N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} equipped with ϕ2(x)(σ+dr2)\phi^{2}(x)(\sigma+dr^{2}). We shall establish the CC^{\infty} approximation of 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) and the Miranda’s observation mentioned in the introduction.
Recall that Ω\Omega is a bounded domain in NN, T0(Ω)T_{0}(\Omega) and C0(Ω)C_{0}(\Omega) denote the set of all smooth vector fields and smooth functions with compact supports in Ω\Omega respectively. Let dvoldvol denote the volume form of NN.

Definition 3.1.

Let uu be a measurable function in Ω\Omega. Define

(3.1) 𝔉ϕ(u,Ω)=sup{Ωϕnh+udiv(ϕnX)dvolhC0(Ω),XT0(Ω) and h2+X,X1}\begin{split}\mathfrak{F}_{\phi}(u,\Omega)&=\sup\{\int_{\Omega}\phi^{n}h+u{{div}}(\phi^{n}X)dvol\mid h\in C_{0}(\Omega),\\ &X\in T_{0}(\Omega)\ \text{ and }h^{2}+{\left<X,X\right>}\leq 1\}\end{split}
Remark 3.2.

If uC1(Ω)u\in C^{1}(\Omega), 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) is the area of the graph of uu in the conformal product manifold MϕM_{\phi}.

3.1. Properties of the area functional

By the definition we have

Lemma 3.3 (Lower-semicontinuity).

Let {uk}k=1\{u_{k}\}_{k=1}^{\infty} be a sequence converging to u(x)u(x) in L1(Ω)L^{1}(\Omega). Then

(3.2) 𝔉ϕ(u,Ω)lim infk+𝔉ϕ(uk,Ω)\mathfrak{F}_{\phi}(u,\Omega)\leq\liminf_{k\to+\infty}\mathfrak{F}_{\phi}(u_{k},\Omega)

The following lemma establishes a relation between 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) and Du(Ω)||Du||(\Omega).

Lemma 3.4.

There is a positive constant μ0:=μ0(Ω,ϕ)\mu_{0}:=\mu_{0}(\Omega,\phi) such that for any uL1(Ω)u\in L^{1}(\Omega)

(3.3) 1μ0max{Du(Ω),vol(Ω)}𝔉ϕ(u,Ω)μ0(Du(Ω)+vol(Ω))\frac{1}{\mu_{0}}\max\{{\left|\left|Du\right|\right|}(\Omega),vol(\Omega)\}\leq\mathfrak{F}_{\phi}(u,\Omega)\leq\mu_{0}({\left|\left|Du\right|\right|}(\Omega)+vol(\Omega))
Proof.

Since Ω¯\bar{\Omega} is compact, there is a positive constant μ0\mu_{0} such that 1μ0ϕn(x)μ0\frac{1}{\mu_{0}}\leq\phi^{n}(x)\leq\mu_{0} on Ω¯\bar{\Omega}. By the conclusion (1) in Theorem 2.11 we have

(3.4) supX,X1,XT0ΩΩu𝑑iv(ϕnX)𝑑vol=Ωϕn(x)d|Du|\sup_{\langle X,X\rangle\leq 1,X\in T_{0}\Omega}\int_{\Omega}u{{div}}(\phi^{n}X)dvol=\int_{\Omega}\phi^{n}(x)d|Du|

We choose hC0(Ω)h\in C_{0}(\Omega) and XT0(Ω)X\in T_{0}(\Omega) satisfying h2+X,X1h^{2}+\langle X,X\rangle\leq 1. By the definition in (3.1)

(3.5) 𝔉ϕ(u,Ω)μ0(vol(Ω))+supX21Ωu𝑑iv(ϕnX)𝑑volμ0(vol(Ω)+Du(Ω)) by (3.4\begin{split}\mathfrak{F}_{\phi}(u,\Omega)&\leq\mu_{0}(vol(\Omega))+\sup_{{\left|\left|X\right|\right|}^{2}\leq 1}\int_{\Omega}u{{div}}(\phi^{n}X)dvol\\ &\leq\mu_{0}(vol(\Omega)+||Du||(\Omega))\text{ by \eqref{eq:af:mid} }\end{split}

Let X=0X=0 and h=1h=1, the definition (3.1) implies that 𝔉ϕ(u,Ω)1μ0vol(Ω)\mathfrak{F}_{\phi}(u,\Omega)\geq\frac{1}{\mu_{0}}vol(\Omega). Now let h=0h=0, (3.4) implies that 𝔉ϕ(u,Ω)1μ0Du(Ω)\mathfrak{F}_{\phi}(u,\Omega)\geq\frac{1}{\mu_{0}}||Du||(\Omega). The proof is complete. ∎

The area functional 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) induces a Radon measure on Ω\Omega as follows.

Theorem 3.5.

Suppose uBV(Ω)u\in BV(\Omega). Then there is a Radon measure ν\nu on Ω\Omega such that for any open set WW in Ω\Omega,

(3.6) ν(W)=𝔉ϕ(u,W)\nu(W)=\mathfrak{F}_{\phi}(u,W)
Proof.

For any nonnegative function fC0(Ω)f\in C_{0}(\Omega), define

(3.7) λ(f)=sup{Ωϕnh+udiv(ϕnX)dvolhC0(Ω),XT0(Ω) and h2+X,Xf2}\begin{split}\lambda(f)&=\sup\{\int_{\Omega}\phi^{n}h+u{{div}}(\phi^{n}X)dvol\mid h\in C_{0}(\Omega),X\in T_{0}(\Omega)\\ &\text{ and }h^{2}+{\left<X,X\right>}\leq f^{2}\}\end{split}

It is easy to see that λ(cf)=cλ(f)\lambda(cf)=c\lambda(f) for any positive constant c>0c>0 and λ(f+g)=λ(f)+λ(g)\lambda(f+g)=\lambda(f)+\lambda(g). Thus λ\lambda is a positive linear functional on C0(Ω)C_{0}(\Omega). By [34, remark 4.3] (see also [41, Theorem 2.5]), then there is a Radon measure ν\nu on Ω\Omega such that

(3.8) ν(W)=sup{λ(f)spt(f)W,0f1}\nu(W)=\sup\{\lambda(f)\mid spt(f)\subset\subset W,0\leq f\leq 1\}

for any open set WΩW\subset\Omega. Here spt(f)spt(f) denotes the closure of {f(x)0:xΩ}\{f(x)\neq 0:x\in\Omega\}. From the definition it is clear that 𝔉ϕ(u,W)=sup{λ(f)spt(f)W,0f1}\mathfrak{F}_{\phi}(u,W)=\sup\{\lambda(f)\mid spt(f)\subset\subset W,0\leq f\leq 1\}. The proof is complete. ∎

Fix any pNp\in N. Let exppexp_{p} be the exponential map at pp. Let injΩ¯inj_{\bar{\Omega}} be the finite number given by

(3.9) min{ the injective radius of x in N:xΩ¯,1}\min\{\text{ the injective radius of }x\text{ in }N:x\in\bar{\Omega},1\}

For any r<injΩ¯r<inj_{\bar{\Omega}}, let Br(0)B_{r}(0) be the Euclidean ball centering at 0 with radius rr. Thus expp:Br(0)Br(p)exp_{p}:B_{r}(0)\rightarrow B_{r}(p) is a diffeomorphism. Via this exponential map we identify Br(p)B_{r}(p) with Br(0)B_{r}(0) equipped with the metric g=gijdxidxjg=g_{ij}dx^{i}dx^{j}. Then Br(p)B_{r}(p) is called a normal ball. A vector field XX along Br(p)B_{r}(p) can be represented by

(3.10) X=XixiX=X^{i}\frac{\partial}{\partial x_{i}}

Let φ(x)\varphi(x) be a symmetric mollifier in n\mathbb{R}^{n}, i.e. a function satisfying φ(x)=φ(x)\varphi(x)=\varphi(-x), spt(φ)B1(0)spt(\varphi)\subset B_{1}(0) and nφ(x)𝑑voln=1\int_{\mathbb{R}^{n}}\varphi(x)dvol_{\mathbb{R}^{n}}=1. Here dvolndvol_{\mathbb{R}^{n}} denotes dx1dxndx^{1}\wedge\cdots\wedge dx^{n}. If uL1(n)u\in L^{1}(\mathbb{R}^{n}), let uφεu*\varphi_{\varepsilon} denote the convolution of uu where φε=1εnφ(xε)\varphi_{\varepsilon}=\frac{1}{\varepsilon^{n}}\varphi(\frac{x}{\varepsilon}) namely

(3.11) uφε(y)=nφε(xy)u(x)𝑑voln.u*\varphi_{\varepsilon}(y)=\int_{\mathbb{R}^{n}}\varphi_{\varepsilon}(x-y)u(x)dvol_{\mathbb{R}^{n}}.

The convolution of a vector field X=XixiX=X^{i}\frac{\partial}{\partial x_{i}} and φε\varphi_{\varepsilon} on Br(p)B_{r}(p) is defined by

(3.12) Xφε:=XiφεxiX*\varphi_{\varepsilon}:=X^{i}*\varphi_{\varepsilon}\frac{\partial}{\partial x_{i}}

Note that dvol=det(g)dvolndvol=\sqrt{\det(g)}dvol_{\mathbb{R}^{n}}. Suppose f,hC0(Br(p))f,h\in C_{0}(B_{r}(p)). By a direct computation we have

(3.13) Br(p)(φεf)h𝑑vol=Br(p)fdetgφε(hdetg)𝑑vol\int_{B_{r}(p)}(\varphi_{\varepsilon}*f)hdvol=\int_{B_{r}(p)}\frac{f}{\sqrt{\det g}}\varphi_{\varepsilon}*(h\sqrt{\det g})dvol
(3.14) Br(p)(φεh)𝑑iv(X)𝑑vol=Br(p)h𝑑iv(1detgφε(detgX))𝑑vol\int_{B_{r}(p)}(\varphi_{\varepsilon}*h){{div}}(X)dvol=\int_{B_{r}(p)}h{{div}}(\frac{1}{\sqrt{\det g}}\varphi_{\varepsilon}*(\sqrt{\det g}X))dvol

To prove the CC^{\infty} approximation property of the area functional 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega), we need the following two techinique lemmas from [41]. Both of their proofs follow from those of [41] with minor modifications. So we skip them here.

Lemma 3.6 (Theorem 2.12, [41]).

Let BB be a normal ball in Ω\Omega with a metric g=gijdxidxjg=g_{ij}dx^{i}dx^{j}. Let ff be a nonnegative continuous function and KBK\subset\subset B be a compact subset. Then for any δ>0\delta>0, there exists a ε0=ε0(Ω,K,g,f)\varepsilon_{0}=\varepsilon_{0}(\Omega,K,g,f) such that for all ε<ε0\varepsilon<\varepsilon_{0}, any continuous function hh and any vector field XX satisfying

(3.15) h2+X,Xf2, in Ωh^{2}+\langle X,X\rangle\leq f^{2},\text{ in }\Omega

we have

(3.16) h2+X,X(f+δ)2 in Kh^{2}+\langle X^{\prime},X^{\prime}\rangle\leq(f+\delta)^{2}\text{ in }K

where XX^{\prime} is defined by

(3.17) X=1det(g)ϕn(x)φε(det(g)ϕn(x)X)X^{\prime}=\frac{1}{\sqrt{\det(g)}\phi^{n}(x)}\varphi_{\varepsilon}*(\sqrt{\det(g)}\phi^{n}(x)X)
Lemma 3.7 (Lemma 2.13, [41]).

Let BB be a normal ball in Ω\Omega with a metric g=gijdxidxjg=g_{ij}dx^{i}dx^{j}. Suppose uBV(B)u\in BV(B) and q(x)q(x) be a smooth function with compact support in BB. Then for any ε>0\varepsilon>0 there is a σ0=σ0(u,g,q,K)\sigma_{0}=\sigma_{0}(u,g,q,K) such that for all σ(0,σ0)\sigma\in(0,\sigma_{0}) and any smooth vector field XX on BB satisfying X,X1\langle X,X\rangle\leq 1.

(3.18) Bφσ(qu)𝑑iv(ϕnX)𝑑volBu𝑑iv(qϕnYσ)𝑑volBϕnX,q𝑑vol+ε\begin{split}\int_{B}\varphi_{\sigma}*(qu)div(\phi^{n}X)dvol&\leq\int_{B}udiv(q\phi^{n}Y_{\sigma})dvol\\ &-\int_{B}\langle\phi^{n}X,\nabla q\rangle dvol+\varepsilon\end{split}

where \nabla is the covariant derivative of NN, Yσ=1det(g)ϕnφσ(det(g)ϕnX)Y_{\sigma}=\frac{1}{\sqrt{det(g)}\phi^{n}}\varphi_{\sigma}*(\sqrt{det(g)}\phi^{n}X) and assume X=0X=0 outside BB.

The CC^{\infty} approximation property of 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) is stated as follows.

Theorem 3.8.

Let Ω\Omega be a bounded domain in NN and uBV(Ω)u\in BV(\Omega). Then there is a sequence {uk}k=1C(Ω)\{u_{k}\}_{k=1}^{\infty}\in C^{\infty}(\Omega) such that

(3.19) limk+Ω|uuk|𝑑vol=0\displaystyle\mathop{\lim}\limits_{k\to+\infty}\int_{\Omega}|u-u_{k}|dvol=0
(3.20) limk+𝔉ϕ(uk,Ω)=𝔉ϕ(u,Ω)\displaystyle\mathop{\lim}\limits_{k\to+\infty}\mathfrak{F}_{\phi}(u_{k},\Omega)=\mathfrak{F}_{\phi}(u,\Omega)
Proof.

By Lemma 3.4 uBV(Ω)u\in BV(\Omega) and that Ω\Omega is bounded imply that 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) is finite.
By Theorem 3.5, there is a Radon measure ν\nu on Ω\Omega such that ν(W)=𝔉ϕ(u,W)\nu(W)=\mathfrak{F}_{\phi}(u,W) for any open set WΩW\subset\Omega. Fix any ε>0\varepsilon>0 and r0>0r_{0}>0. By [41, Theorem A.4], there exists a countable open cover {Bk}k=1\{B_{k}\}_{k=1}^{\infty} of Ω\Omega and a positive integer κ0\kappa_{0} such that

  1. (1)

    Each BkB_{k} is a normal ball such that ν(Bk)=0\nu(\partial B_{k})=0 and diam(Bk)2r0\text{diam}(B_{k})\leq 2r_{0}.

  2. (2)

    {B1,,Bκ0}\{B_{1},...,B_{\kappa_{0}}\} is a pairwise disjoint subcollection with

    (3.21) 𝔉ϕ(u,Ω)εk=1κ0𝔉ϕ(u,Bk)𝔉ϕ(u,Ω).\mathfrak{F}_{\phi}(u,\Omega)-\varepsilon\leq\sum_{k=1}^{\kappa_{0}}\mathfrak{F}_{\phi}(u,B_{k})\leq\mathfrak{F}_{\phi}(u,\Omega).
  3. (3)

    The subcollection {Bk}k=κ0+1\{B_{k}\}_{k=\kappa_{0}+1}^{\infty} satisfies

    (3.22) k=κ0+1𝔉ϕ(u,Bk)κ1ε\sum_{k=\kappa_{0}+1}^{\infty}\mathfrak{F}_{\phi}(u,B_{k})\leq\kappa_{1}\varepsilon

    where κ1\kappa_{1} is a positive integer only depending on Ω\Omega and nn.

Take a partition of unity {ζk}\{\zeta_{k}\} subordinate to the open cover {Bk}\{B_{k}\}, that is

(3.23) ζkC0(Bk),0ζk1 and k=1ζk=1 on Ω.\zeta_{k}\in C^{\infty}_{0}(B_{k}),0\leq\zeta_{k}\leq 1\text{ and }\sum_{k=1}^{\infty}\zeta_{k}=1\text{ on }\Omega.

and for each xΩx\in\Omega. And for any xx in Ω\Omega there is a compact neighborhood VV of xx such that on VV the summation above is finite.
Now fix any hC0(Ω)h\in C_{0}(\Omega) and any XT0ΩX\in T_{0}\Omega satisfying h2+X,X21h^{2}+\langle X,X\rangle^{2}\leq 1. For each kk by Lemma 3.6 and Lemma 3.7 we can choose εk\varepsilon_{k} so small independent of XX and hh such that spt(φεk(uζk))Bk{\text{spt}}(\varphi_{\varepsilon_{k}}*(u\zeta_{k}))\subset B_{k},

(3.24) Bk|φεk(uζk)uζk|𝑑vol<ε2k\displaystyle\int_{B_{k}}|\varphi_{\varepsilon_{k}}*(u\zeta_{k})-u\zeta_{k}|dvol<\frac{\varepsilon}{2^{k}}
(3.25) (ζkh)2+ζk2Yεk,Yεk1+ε\displaystyle(\zeta_{k}h)^{2}+\zeta_{k}^{2}\langle Y_{\varepsilon_{k}},Y_{\varepsilon_{k}}\rangle\leq 1+\varepsilon
(3.26) Bkφεk(uζk)div(ϕn(x)X)dvolBku𝑑iv(ϕnζkYεk)𝑑volBkuϕnX,ζk𝑑vol+ε2k\begin{split}\int_{B_{k}}\varphi_{\varepsilon_{k}}*(u\zeta_{k})&div(\phi^{n}(x)X)dvol\leq\int_{B_{k}}udiv(\phi^{n}\zeta_{k}Y_{\varepsilon_{k}})dvol\\ &-\int_{B_{k}}u\langle\phi^{n}X,\nabla\zeta_{k}\rangle dvol+\frac{\varepsilon}{2^{k}}\end{split}

Here Yεk=1det(g)ϕnφεk(det(g)ϕnX)Y_{\varepsilon_{k}}=\frac{1}{\sqrt{\det(g)}\phi^{n}}\varphi_{\varepsilon_{k}}*(\sqrt{\det(g)}\phi^{n}X). Now we define uεu_{\varepsilon} by

(3.27) uε=kφεk(uζk)u_{\varepsilon}=\sum_{k}^{\infty}\varphi_{\varepsilon_{k}}*(u\zeta_{k})

Thus uεC(Ω)u_{\varepsilon}\in C^{\infty}(\Omega). By (3.26) we obtain

Ω{ϕnh+uεdiv(ϕnX)}𝑑vol=k=1Bk{ϕnζkh+φεk(uζk)div(ϕn(x)X}dvolk=1Bk{ϕnζkh+udiv(ϕnζkYεk)uϕnX,ζk}𝑑vol+ε by (3.26)=k=1Bk{ϕnζkh+udiv(ϕnζkYεk)}𝑑vol+ε\displaystyle\begin{split}&\quad\int_{\Omega}\{\phi^{n}h+u_{\varepsilon}{{div}}(\phi^{n}X)\}dvol\\ &=\sum_{k=1}^{\infty}\int_{B_{k}}\{\phi^{n}\zeta_{k}h+\varphi_{\varepsilon_{k}}*(u\zeta_{k}){{div}}(\phi^{n}(x)X\}dvol\\ &\leq\sum_{k=1}^{\infty}\int_{B_{k}}\{\phi^{n}\zeta_{k}h+udiv(\phi^{n}\zeta_{k}Y_{\varepsilon_{k}})-u\langle\phi^{n}X,\nabla\zeta_{k}\rangle\}dvol+\varepsilon\text{ by \eqref{eq:mid:result}}\\ &=\sum_{k=1}^{\infty}\int_{B_{k}}\{\phi^{n}\zeta_{k}h+udiv(\phi^{n}\zeta_{k}Y_{\varepsilon_{k}})\}dvol+\varepsilon\end{split}

The last line we use the fact that k=1BkuϕnX,ζkdvol=0\sum_{k=1}^{\infty}\int_{B_{k}}-u\langle\phi^{n}X,\nabla\zeta_{k}\rangle dvol=0. By definition we have

(3.28) Bk{ϕnζkh+udiv(ϕnζkYεk)}𝑑vol(1+ε)𝔉ϕ(u,Bk)\int_{B_{k}}\{\phi^{n}\zeta_{k}h+udiv(\phi^{n}\zeta_{k}Y_{\varepsilon_{k}})\}dvol\leq(1+\varepsilon)\mathfrak{F}_{\phi}(u,B_{k})

Combining this with (3.21), (3.22) together we obtain

(3.29) Ω{ϕnh+uεdiv(ϕnX)}𝑑vol(1+ε)k=1𝔉ϕ(u,Bk)(1+ε){𝔉ϕ(u,Ω)+κε}\begin{split}\int_{\Omega}\{\phi^{n}h+u_{\varepsilon}{{div}}(\phi^{n}X)\}dvol&\leq(1+\varepsilon)\sum_{k=1}^{\infty}\mathfrak{F}_{\phi}(u,B_{k})\\ &\leq(1+\varepsilon)\{\mathfrak{F}_{\phi}(u,\Omega)+\kappa\varepsilon\}\end{split}

Since h,Xh,X are chosen arbitrarily satisfying hC0(Ω),XT0(Ω)h\in C_{0}(\Omega),X\in T_{0}(\Omega) and h2+X,X1h^{2}+\langle X,X\rangle\leq 1, we conclude that

(3.30) 𝔉ϕ(uε,Ω)(1+ε)(𝔉ϕ(u,Ω)+κε)\mathfrak{F}_{\phi}(u_{\varepsilon},\Omega)\leq(1+\varepsilon)(\mathfrak{F}_{\phi}(u,\Omega)+\kappa\varepsilon)

As a result we can choose εi0\varepsilon_{i}\rightarrow 0 such that limεi0sup𝔉ϕ(uεi,Ω)𝔉ϕ(u,Ω)\lim_{\varepsilon_{i}\rightarrow 0}\sup\mathfrak{F}_{\phi}(u_{\varepsilon_{i}},\Omega)\leq\mathfrak{F}_{\phi}(u,\Omega). On the other hand by (3.24) uεiu_{\varepsilon_{i}} converges to uu in L1(Ω)L^{1}(\Omega). On the other hand by Lemma 3.3 liminf𝔉ϕ(uεi,Ω)𝔉ϕ(u,Ω)\lim\inf\mathfrak{F}_{\phi}(u_{\varepsilon_{i}},\Omega)\geq\mathfrak{F}_{\phi}(u,\Omega). Thus we obtain the conclusion. The proof is complete. ∎

3.2. The Miranda’s observation

In this subsection we show the Miranda’s observation for 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) mentioned in the introduction.
Let uu be a measurable function on Ω\Omega. The set {(x,t)|xΩ,t<u(x)}\{(x,t)|x\in\Omega,t<u(x)\} is called as the subgraph of uu, written as UU. For any Borel set EE we denote by λE\lambda_{E} its characterization function.
Let PϕP_{\phi} denote the perimeter of Caccioppoli sets in MϕM_{\phi}. Its connection with the area functional 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) can be summarized as follows.

Theorem 3.9.

Suppose uBV(Ω)u\in BV(\Omega), UU is its subgraph and Ω\Omega is a bounded domain. Then

(3.31) Pϕ(U,Ω×)=𝔉ϕ(u,Ω)P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})=\mathfrak{F}_{\phi}(u,\Omega)
Proof.

By Lemma 3.4 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) is finite. If uu is a smooth function, the equality is obvious since both sides equal to the area of the graph of uu in MϕM_{\phi}.
Suppose uBV(Ω)u\in BV(\Omega). By Theorem 3.8 there is a sequence of smooth functions {uk}k=1\{u_{k}\}_{k=1}^{\infty} such that {uk}\{u_{k}\} converges to uu in L1(Ω)L^{1}(\Omega) and 𝔉ϕ(uk,Ω)\mathfrak{F}_{\phi}(u_{k},\Omega) converges to 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega) as kk\rightarrow\infty. Let UkU_{k} be the subgraph of uku_{k}. Note that {λUk}\{\lambda_{U_{k}}\} converges to λU\lambda_{U} in a.e. By Lemma 2.8 we have

(3.32) Pϕ(U,Ω×)limkinfPϕ(Uk,Ω×)=limkinf𝔉ϕ(uk,Ω)=𝔉ϕ(u,Ω)\begin{split}P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})&\leq\lim_{k\rightarrow\infty}\inf P_{\phi}(U_{k},\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\\ &=\lim_{k\rightarrow\infty}\inf\mathfrak{F}_{\phi}(u_{k},\Omega)=\mathfrak{F}_{\phi}(u,\Omega)\end{split}

To prove the converse first suppose that |u|μ|u|\leq\mu for some positive constant μ>1\mu>1. Observe that we can add a constant to uu without changing Pϕ(U,Ω×)P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}) and 𝔉ϕ(u,Ω)\mathfrak{F}_{\phi}(u,\Omega). Thus without loss of generality we assume that u1u\geq 1. Let XT0ΩX\in T_{0}\Omega and h(x)C0(Ω)h(x)\in C_{0}(\Omega) satisfying X,X+h2(x)1\langle X,X\rangle+h^{2}(x)\leq 1. Let η(t)\eta(t) be a smooth nonnegative function with compact support in (0,μ+1)(0,\mu+1) such that η1\eta\equiv 1 in [1,μ][1,\mu] and η1\eta\leq 1 on (0,μ+1)(0,\mu+1). Define

(3.33) X=1ϕ(x)η(r)(X+h(x)r)X^{\prime}=\frac{1}{\phi(x)}\eta(r)(X+h(x)\partial_{r})

Then XT0(Ω×)X^{\prime}\in T_{0}(\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}) and X,X1\langle X^{\prime},X^{\prime}\rangle\leq 1 in MϕM_{\phi}. Observe that for each xΩx\in\Omega, we have

(3.34) λU(x,r)η𝑑r=u(x)+C,λU(x,r)η𝑑r=1\int_{\mathbb{R}}\lambda_{U}(x,r)\eta dr=u(x)+C,\quad\int_{\mathbb{R}}\lambda_{U}(x,r)\eta^{\prime}dr=1

where CC is a constant depending only on η\eta. We denote the product manifold N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} with metric σ+dr2\sigma+dr^{2} by MM. Let divMϕdiv_{M_{\phi}} and divMdiv_{M} be the divergence on MϕM_{\phi} and MM respectively. Note that divMϕ(X)=1ϕn+1divM(ϕn+1X)div_{M_{\phi}}(X^{\prime})=\frac{1}{\phi^{n+1}}div_{M}{(\phi^{n+1}X^{\prime})} (For example see [41, (2.11)]. Then by definition of perimeter, we have

Pϕ(U,Ω×)Ω×λU𝑑ivMϕ(X)𝑑volMϕ=Ω{div(ϕnX)λUη𝑑r+hϕn(λUη𝑑r)}𝑑vol=Ω{udiv(ϕnX)+hϕn}𝑑vol\begin{split}P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})&\geq\int_{\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}}\lambda_{U}{{div}}_{M_{\phi}}(X^{\prime})dvol_{M_{\phi}}\\ &=\int_{\Omega}\{{{div}}(\phi^{n}X)\int_{\mathbb{R}}\lambda_{U}\eta dr+h\phi^{n}(\int_{\mathbb{R}}\lambda_{U}\eta^{\prime}dr)\}dvol\\ &=\int_{\Omega}\{u{{div}}(\phi^{n}X)+h\phi^{n}\}dvol\end{split}

Taking supremum over both sides for all hC0(Ω),XT0Ωh\in C_{0}(\Omega),X\in T_{0}\Omega with h2+X,X1h^{2}+\langle X,X\rangle\leq 1 yields that Pϕ(U,Ω×)𝔉ϕ(u,Ω)P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\geq\mathfrak{F}_{\phi}(u,\Omega). With (3.32) this yields (3.31) in the case of |u|μ|u|\leq\mu.
As for the general case we consider an approximation procedure based on the finite case. Let uTu_{T} be the truncation of uu by TT, i.e. uT=max{min{u,T},T}u_{T}=\max\{\min\{u,T\},-T\}. Let UTU_{T} be the subgraph of uTu_{T}. Thus λUT\lambda_{U_{T}} converges to λU\lambda_{U} a.e. in Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. By (3.32) and the lower semicontinuity in Lemma 3.3, Pϕ(U,Ω×)P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}) is finite. Note that

(3.35) limT+Pϕ(U,Ω×(T,T))=Pϕ(U,Ω×)Pϕ(UT,Ω×(T,T))=Pϕ(U,Ω×(T,T))\begin{split}\lim_{T\rightarrow+\infty}&P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T))=P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\\ &P_{\phi}(U_{T},\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T))=P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T))\end{split}

Combining this with Lemma 3.3 we obtain

(3.36) limT+Pϕ(UT,Ω×)=Pϕ(U,Ω×(T,T))\begin{split}\lim_{T\rightarrow+\infty}P_{\phi}(U_{T},\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})=P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T))\end{split}

As a result we conclude

(3.37) Pϕ(U,Ω×)=limT+Pϕ(UT,Ω×)=limT+𝔉ϕ(uT,Ω)𝔉ϕ(u,Ω)\begin{split}P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})&=\lim_{T\rightarrow+\infty}P_{\phi}(U_{T},\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\\ &=\lim_{T\rightarrow+\infty}\mathfrak{F}_{\phi}(u_{T},\Omega)\geq\mathfrak{F}_{\phi}(u,\Omega)\end{split}

Combining this with (3.32) yields the conclusion. The proof is complete. ∎

As an application we obtain a decreasing property of certain Caccioppoli sets in MϕM_{\phi}.

Theorem 3.10.

Let EΩ×E\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} be a Caccioppoli set in MϕM_{\phi} with the following assumption: for almost every xΩx\in\Omega, there exists a Tx>0T_{x}>0 such that λE(x,t)=0\lambda_{E}(x,t)=0 for all t>Txt>T_{x} and λE(x,t)=1\lambda_{E}(x,t)=1 for all t<Txt<-T_{x}. Then the function

(3.38) w(x)=limk+(kkλE(x,t)𝑑tk)w(x)=\lim_{k\to+\infty}(\int^{k}_{-k}\lambda_{E}(x,t)dt-k)

is well-defined and

(3.39) 𝔉ϕ(w,Ω)Pϕ(E,Ω×)\mathfrak{F}_{\phi}(w,\Omega)\leq P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})
Remark 3.11.

This result should be firstly observed by Miranda [29] in the case of ϕ(x)1\phi(x)\equiv 1 and N=nN=\mathbb{R}^{n} (see also [2]). It is a generalization of [18, Theorem 14.8] and the symmetrization of hyperbolic spaces in [25, remark 2.3] by Lin. It is also similar to the arrangement of constant mean curvature functional in [9, Theorem 3.1] by De Silva-Spruck, the decreasing perimeter property of singular area functional in [7, Lemma 9] by Bemelmans-Dierkes, [15, Lemma 3.3] by Gerhardt and [41, Lemma 5.8].

Proof.

By the assumption on EE it is not hard to see that ω(x)\omega(x) is well defined. Moreover we assume that EE has finite perimeter in MϕM_{\phi}. Otherwise nothing needs to prove.
First we assume that there is a T>0T>0 such that EΩ×(T,T)\partial E\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T). Let η(r)\eta(r) be a compactly supported smooth function on \mathbb{R} such that η(t)1\eta(t)\equiv 1 on (1,1)(-1,1) and η(r)1\eta(r)\leq 1. For each kN+k\in N^{+}, define

(3.40) ηk(r)={η(r+k)r(,k)1r[k,k]η(rk)r(k,+)\eta_{k}(r)=\left\{\begin{aligned} &\eta(r+k)\quad r\in(-\infty,-k)\\ &1\quad r\in[-k,k]\\ &\eta(r-k)\quad r\in(k,+\infty)\end{aligned}\right.

Choose hC0(Ω)h\in C_{0}(\Omega) and XT0(Ω)X\in T_{0}(\Omega) satisfying h2+X,X1h^{2}+\langle X,X\rangle\leq 1. Define

(3.41) X=1ϕ(x)ηk(r)(X+h(x)r)X^{\prime}=\frac{1}{\phi(x)}\eta_{k}(r)(X+h(x)\partial_{r})

Thus X,X1\langle X^{\prime},X^{\prime}\rangle\leq 1. Then we have

(3.42) Pϕ(E,Ω×)Ω×λE𝑑ivMϕ(X)𝑑volMϕ=Ω×λE1ϕn+1𝑑ivM(ϕnηk(r)(X+hr))ϕn+1𝑑r𝑑vol=Ω{div(ϕnX)λEηk(r)𝑑r+ϕnhλEηk(r)𝑑r}𝑑vol\begin{split}&P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\geq\int_{\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}}\lambda_{E}{{div}}_{M_{\phi}}(X^{\prime})dvol_{M_{\phi}}\\ &=\int_{\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}}\lambda_{E}\frac{1}{\phi^{n+1}}div_{M}(\phi^{n}\eta_{k}(r)(X+h\partial_{r}))\phi^{n+1}drdvol\\ &=\int_{\Omega}\{{{div}}(\phi^{n}X)\int_{\mathbb{R}}\lambda_{E}\eta_{k}(r)dr+\phi^{n}h\int_{\mathbb{R}}\lambda_{E}\eta_{k}^{\prime}(r)dr\}dvol\end{split}

where dvolMϕdvol_{M_{\phi}} and divMϕdiv_{M_{\phi}} denote the volume form and the divergence of MϕM_{\phi} respectively, divMdiv_{M} is the divergence of the product manifold M:=(N×,σ+dr2)M:=(N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R},\sigma+dr^{2}). By the assumption on EE and EΩ×[T,T]\partial E\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}[-T,T], for almost every xΩx\in\Omega we have

(3.43) limk+λE(x,r)ηk(r)𝑑r=1\displaystyle\lim_{k\to+\infty}\int_{\mathbb{R}}\lambda_{E}(x,r)\eta^{\prime}_{k}(r)dr=1
(3.44) limk+ηk(r)λE(x,r)𝑑r=w(x)+C\displaystyle\lim_{k\to+\infty}\int_{\mathbb{R}}\eta_{k}(r)\lambda_{E}(x,r)dr=w(x)+C

where CC is a constant depending only on η\eta and TT.

Let kk\to\infty, we get

(3.45) Pϕ(E,Ω×)Ω𝑑iv(ϕnX)w+ϕnhdvolP_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\geq\int_{\Omega}{{div}}(\phi^{n}X)w+\phi^{n}hdvol

Because h(x)C0(Ω)h(x)\in C_{0}(\Omega) and XT0ΩX\in T_{0}\Omega are arbitrarily choosen satisfying h2+X,X1h^{2}+\langle X,X\rangle\leq 1, we obtain (3.39) in the case of EΩ×(T,T)\partial E\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T).
As for a general Caccioppoli set EE with finite perimeter, we set ET=EΩ×(,T)\Ω×[T,+)E_{T}=E\cup\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,-T)\backslash\Omega{\mkern-1.0mu\times\mkern-1.0mu}[T,+\infty). Note that limT+Pϕ(E,Ω×(T,T))=Pϕ(E,Ω×)\lim_{T\rightarrow+\infty}P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T))=P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}) and Pϕ(ET,Ω×(T,T))=Pϕ(E,Ω×(T,T))P_{\phi}(E_{T},\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T))=P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}(-T,T)). By the lower semicontinuity in Lemma 3.3, we obtain

(3.46) limT+Pϕ(ET,Ω×)=Pϕ(E,Ω×)\lim_{T\rightarrow+\infty}P_{\phi}(E_{T},\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})=P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})

Define

wT(x)=limk+(kkλET(x,t)𝑑tk)w_{T}(x)=\lim_{k\to+\infty}(\int^{k}_{-k}\lambda_{E_{T}}(x,t)dt-k)

By the assumption of EE, wT(x)w_{T}(x) converges to w(x)w(x) a.e. xΩx\in\Omega as T+T\rightarrow+\infty. Again by Lemma 3.3

(3.47) 𝔉ϕ(w(x),Ω)liminf𝔉ϕ(wT(x),Ω)liminfPϕ(ET,Ω×)Pϕ(E,Ω×)\begin{split}\mathfrak{F}_{\phi}(w(x),\Omega)&\leq\lim\inf\mathfrak{F}_{\phi}(w_{T}(x),\Omega)\\ &\leq\lim\inf P_{\phi}(E_{T},\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\leq P_{\phi}(E,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\end{split}

The proof is complete. ∎

Theorem 3.10 and the following result together are referred as the Miranda’s observation.

Theorem 3.12.

Let uBV(Ω)u\in BV(\Omega) with 𝔉ϕ(u,Ω)<\mathfrak{F}_{\phi}(u,\Omega)<\infty. Suppose u(x)u(x) locally minimizes 𝔉ϕ(.,Ω)\mathfrak{F}_{\phi}(.,\Omega), i.e. 𝔉ϕ(u,Ω)𝔉ϕ(v,Ω)\mathfrak{F}_{\phi}(u,\Omega)\leq\mathfrak{F}_{\phi}(v,\Omega) if UΔVΩ×U\Delta V\subset\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} where U,VU,V are the subgraphs of u(x),v(x)u(x),v(x) respectively. Then UU locally minimizes the perimeter in MϕM_{\phi}, i.e.

Pϕ(U,Ω×)Pϕ(F,Ω×)P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\leq P_{\phi}(F,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})

for any Caccioppoli set FF satisfying FΔUΩ×F\Delta U\subset\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}.

Proof.

Let FΩ×F\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} be a Caccioppoli set satisfying FΔUΩ×F\Delta U\subset\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. Since UU is a subgraph of u(x)u(x), it is easy to see FF satisfying the condition in Theorem 3.10. Let w(x)w(x) be defined as in Theorem 3.10. Then

(3.48) Pϕ(U,Ω×)=𝔉ϕ(u,Ω)𝔉ϕ(w(x),Ω)Pϕ(F,Ω×)P_{\phi}(U,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})=\mathfrak{F}_{\phi}(u,\Omega)\leq\mathfrak{F}_{\phi}(w(x),\Omega)\leq P_{\phi}(F,\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})

The poof is complete. ∎

4. The area minimizing problem

In this section we consider the area minimizing problem in the conformal product manifold MϕM_{\phi} (see Definition 1.1) similar to that in [14].
Throughout this section suppose Ω\Omega and Ω\Omega^{\prime} are two bounded C2C^{2} domains in NN satisfying ΩΩ\Omega\subset\subset\Omega^{\prime}. Let QϕQ_{\phi} be the set Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} in MϕM_{\phi} and Q¯ϕ\bar{Q}_{\phi} be its closure Ω¯×\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. Suppose ψ(x)C1(Ω\Ω)\psi(x)\in C^{1}(\Omega^{\prime}\backslash\Omega). We denote by nn the dimension of NN. Let 𝒢\mathcal{G} denote the set of nn-integer multiplicity currents with compact support in Q¯ϕ\bar{Q}_{\phi}, i.e. for any T𝒢T\in\mathcal{G}, its support spt(T)spt(T) is contained in Ω¯×[a,b]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b] for some finite numbers a<ba<b. Let Γ\Gamma be the graph of ψ(x)\psi(x) on Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. The area minimizing problem in this section is to find a solution to attain the value

(4.1) min{𝕄(T):T𝒢,T=Γ}\min\{\mathbb{M}(T):T\in\mathcal{G},\partial T=\Gamma\}

where 𝕄\mathbb{M} is the mass of TT in MϕM_{\phi} (see subsection 4.1).
The main result of this section is stated as follows.

Theorem 4.1.

Suppose ϕ(x)\phi(x) is a C3C^{3} positive function on NN. Then there is a u(x)BV(Ω)u(x)\in BV(\Omega^{\prime}) satisfying u(x)=ψ(x)u(x)=\psi(x) outside Ω\Omega with the following three properties:

  1. (1)

    u(x)u(x) realizes the minimum

    (4.2) min{𝔉ϕ(v(x),Ω):v(x)BV(Ω),v(x)=ψ(x) outside Ω}\min\{\mathfrak{F}_{\phi}(v(x),\Omega^{\prime}):v(x)\in BV(\Omega^{\prime}),v(x)=\psi(x)\text{ outside $\Omega$}\}
  2. (2)

    T=[[U]]|Q¯ϕT=\partial[[U]]|_{\bar{Q}_{\phi}} solves the area minimizing problem (4.1) where UU is the subgraph of u(x)u(x) in Ω×\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R},

  3. (3)

    u(x)C2,β(Ω)u(x)\in C^{2,\beta}(\Omega) satisfies (2.3) and minΩψ(x)u(x)maxΩψ(x)\min_{\partial\Omega}\psi(x)\leq u(x)\leq\max_{\partial\Omega}\psi(x) for any xΩx\in\Omega and any β(0,1)\beta\in(0,1).

Remark 4.2.

In the case of ϕ(x)1\phi(x)\equiv 1, the above theorem is obtained by Lin in [25, section 4.1]. But we do not understand the proof of its uniqueness just assuming Ω\Omega is bounded Lipschitz.

Remark 4.3.

We show that for TT in the property (2) T=Γ\partial T=\Gamma. For r>0r>0 we denote the set {xΩ:dist(x,Ω)<r}\{x\in\Omega^{\prime}:dist(x,\Omega)<r\} by Ωr\Omega_{r}. When rr is sufficiently small, Ωr\Omega_{r} is a C2C^{2} domain. We define Tr:=[[U]]|Ω¯r×T_{r}:=\partial[[U]]|_{\bar{\Omega}_{r}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}}. It is obvious that TrT_{r} is a C2C^{2} graph near Ωr\partial\Omega_{r}. Thus Tr=graph(ψ(x)|Ωr)\partial T_{r}=graph(\psi(x)|_{\partial\Omega_{r}}). And TrT_{r} converges to TT and Tr\partial T_{r} converges to T\partial T in the sense of current as r0r\rightarrow 0. Moreover graph(ψ(x)|Ωr)graph(\psi(x)|_{\partial\Omega_{r}}) converges to Γ\Gamma in the C1C^{1} sense as r0r\rightarrow 0. By the definition of the current T=Γ\partial T=\Gamma. This is the reason that we assume ψ(x)C1(Ω\Ω)\psi(x)\in C^{1}(\Omega^{\prime}\backslash\Omega).

4.1. Integer Multiplicity Currents

Here we collect some necessary facts on integer multiplicity currents. Our main references are [34, section 27] by Simon and [24] by Lin-Yang.
Let UU be an open domain in a Riemannian manifold MM with dimension mm and HjH^{j} denotes the jj-dimensional Hausdorff measure in MM for any j>0j>0. Suppose kk is an integer in [0,m][0,m]. Let Dk(U)D^{k}(U) be the set of all kk-smooth forms with compact support in UU. A kk-current in MM is a linear functional in Dk(M)D^{k}(M).

Definition 4.4.

A set EME\subset M is said to be countable kk-rectifiable if

EE0j=1Fj(Ej)E\subset E_{0}\cup_{j=1}^{\infty}F_{j}(E_{j})

where Hk(E0)=0H^{k}(E_{0})=0 and Fj:EjkMF_{j}:E_{j}\subset\mathbb{R}^{k}\rightarrow M is a Lipschitz map for each jj.

Now we can define a kk-integer multiplicity rectifiable current.

Definition 4.5.

Let TT be a kk-current in MM. We say that TT is a kk-integer multiplicity current if

(4.3) T(ω)=Sω,ηθ(x)𝑑Hk(x)T(\omega)=\int_{S}\langle\omega,\eta\rangle\theta(x)dH^{k}(x)

where SS is a countable kk-rectifiable subset of MM, θ\theta is a positive locally HkH^{k}-integrable function which is integer-valued, and η\eta is a kk-form τ1τk\tau_{1}\wedge\cdots\wedge\tau_{k} oriented the tangent space of SS a.e. HkH^{k}. TT is also written as τ(S,θ,η)\tau(S,\theta,\eta).
The mass of TT in UU is

(4.4) 𝕄U(T):=sup{T(ω):ω,ω1,ωDk(U)}\mathbb{M}_{U}(T):=sup\{T(\omega):\langle\omega,\omega\rangle\leq 1,\omega\in D^{k}(U)\}

where <,><,> denotes the usual pairing of kk-form. The boundary of TT is defined by T(ω)=T(dω)\partial T(\omega)=T(d\omega) for any ωDk1(U)\omega\in D^{k-1}(U).

Remark 4.6.

For any kk-submanifold MM^{\prime} [[M]][[M^{\prime}]] is a k-integer multiplicity current just choosing η\eta as its orientation which is equal to τ(M,1,η)\tau(M^{\prime},1,\eta)
If the dimension of an integer multiplicity current TT, written as τ(S,θ,η)\tau(S,\theta,\eta), is equal to the dimension of MM we always choose η\eta as the volume form of MM. In this case TT is written as τ(S,θ)\tau(S,\theta).

A good property of integer multiplicity currents is their compactness obtained by Federer and Fleming in [13] (see also [34, Theorem 27.3]).

Theorem 4.7.

Suppose {Tj}j=1\{T_{j}\}_{j=1}^{\infty} is a sequence of kk-integer multiplicity currents with

sup{𝕄W(Tj)+𝕄W(Tj)}<\sup\{\mathbb{M}_{W}(T_{j})+\mathbb{M}_{W}(\partial T_{j})\}<\infty

for any open set WMW\subset\subset M. Then there is a kk-integer multiplicity current TT such that TjT_{j} converges weakly to TT and

𝕄W(T)limj+supij𝕄W(Ti)\mathbb{M}_{W}(T)\leq\lim_{j\rightarrow+\infty}\sup_{i\geq j}\mathbb{M}_{W}(T_{i})

A useful way to construct integer multiplicity currents is the pushforward of local Lipschitz maps.

Definition 4.8.

Let U,VU,V be two open sets in (different) Riemannian manifolds. Suppose f:UVf:U\rightarrow V is local Lipschitz, T=τ(S,η,θ)T=\tau(S,\eta,\theta) is a kk-integer multiplicity current and f|sptTf|sptT is proper. We define f#Tf_{\#}T by

(4.5) f#T(ω)=Sω|f(x),df#ηθ(x)𝑑Hk(x)f_{\#}T(\omega)=\int_{S}\langle\omega|_{f(x)},df_{\#}\eta\rangle\theta(x)dH^{k}(x)
Theorem 4.9.

Let QϕQ_{\phi} be the conformal cone in MϕM_{\phi} (see Definition 1.1) and n=dimNn=dimN. Let k<n+1k<n+1 be a positive integer. Let TT be a kk-integer multiplicity current in 𝒢\mathcal{G} satisfying T=0\partial T=0. Then there is a (k+1)(k+1)-integer multiplicity current RR in Q¯ϕ\bar{Q}_{\phi} such that R=T\partial R=T. Here spt(R)spt(R) may be noncompact in Q¯ϕ\bar{Q}_{\phi}.

Remark 4.10.

A similar proof also appears in [34, section 26.26].

Proof.

The proof is exactly the same as that of [14, Theorem 3.9]. Note that even with the metric ϕ2(x)(σ+dr2)\phi^{2}(x)(\sigma+dr^{2}) the map h:Mϕ×Mϕh:M_{\phi}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}\rightarrow M_{\phi} as h((x,r),t)=(x,r+t)h((x,r),t)=(x,r+t) is still proper and local Lipschitz. Thus h#([[(,0)]]×T)h_{\#}([[(-\infty,0)]]{\mkern-1.0mu\times\mkern-1.0mu}T) is well-defined. Moreover

h#([[(,0)]]×T)\displaystyle\partial h_{\#}([[(-\infty,0)]]{\mkern-1.0mu\times\mkern-1.0mu}T) =h#(([[(,0)]]×T))\displaystyle=h_{\#}(\partial([[(-\infty,0)]]{\mkern-1.0mu\times\mkern-1.0mu}T))
=h#({0}×T)h#((,0)×T)\displaystyle=h_{\#}(\{0\}{\mkern-1.0mu\times\mkern-1.0mu}T)-h_{\#}((-\infty,0){\mkern-1.0mu\times\mkern-1.0mu}\partial T)
=Th#((,0)×T)\displaystyle=T-h_{\#}((-\infty,0){\mkern-1.0mu\times\mkern-1.0mu}\partial T)

The conclusion follows from that T=0\partial T=0. ∎

Fix two numbers a<ba<b such that ΓΩ×(a,b)\Gamma\subset\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}(a,b). Now consider an auxiliary problem of (4.1) as follows.

(4.6) Aa,b:=min{𝕄(T):T𝒢,T=Γ,spt(T)Ω¯×[a,b]}A_{a,b}:=\min\{\mathbb{M}(T):T\in\mathcal{G},\partial T=\Gamma,spt(T)\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]\}

By Theorem 4.7 there is a n-integer multiplicity current Ta,bT_{a,b} contained in Ω¯×[a,b]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b] with Ta,b=Γ\partial T_{a,b}=\Gamma satisfying 𝕄(Ta,b)=Aa,b\mathbb{M}(T_{a,b})=A_{a,b}. We denote Ω\Omega^{\prime} by an open domain in NN satisfying ΩΩ\Omega\subset\subset\Omega^{\prime}. Similar to [14, Lemma 4.6] we have

Lemma 4.11.

We can choose Ta,bT_{a,b} such that there is a Caccioppoli set FF in Ω×\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} such that Ta,b=[[F]]|Q¯ϕT_{a,b}=\partial[[F]]|_{\bar{Q}_{\phi}} and FF is the subgraph of ψ(x)\psi(x) outside Q¯ϕ\bar{Q}_{\phi} .

Remark 4.12.

Because Ta,b=[[F]]|Q¯ϕT_{a,b}=\partial[[F]]|_{\bar{Q}_{\phi}} is contained in Ω¯×[a,b]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b] one sees that for each xΩx\in\Omega it holds that λF(x,t)=1\lambda_{F}(x,t)=1 for t<at<a and λF(x,t)=0\lambda_{F}(x,t)=0 for t>bt>b. If not the case we can replace FΩ×F\cap\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}, the complement of FF, with FcΩ×F^{c}\cap\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} without any change of the perimeter.

Proof.

Since ψ(x)C1(Ω\Ω)\psi(x)\in C^{1}(\Omega^{\prime}\backslash\Omega) we can extend ψ(x)\psi(x) as a C1C^{1} function on Ω\Omega^{\prime} satisfying a<ψ(x)<ba<\psi(x)<b on Ω\Omega^{\prime} and its subgraph EE is a Caccioppoli set in Ω×\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. Define S:=[[E]]|Ω¯×[a,b]S:=\partial[[E]]|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]} . Then by remark 4.3 S=Γ\partial S=\Gamma .
From Theorem 4.9 there is a (n+1)(n+1)-integer multiplicity current RR in Ω×(,b]\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b] such that Ta,bS=RT_{a,b}-S=\partial R. Then we have

(4.7) Ta,b=[[E]]|Q¯ϕ+R=[[E]]|Ω¯×[a,b]+R|Ω¯×[a,b]T_{a,b}=\partial[[E]]|_{\bar{Q}_{\phi}}+\partial R=\partial[[E]]|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]}+\partial R|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]}

Observe that [[E]]+R[[E]]+R can be represented as τ(Ω×(,b),θ)\tau(\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b),\theta) where θ\theta is some integer value measurable function on N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. Since spt(R)Ω¯×(,b]spt(R)\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b], θ=λE\theta=\lambda_{E} outside Q¯ϕ\bar{Q}_{\phi} i.e.

(4.8) θ=1 or 0 outside Q¯ϕ\theta=1\text{ or }0\quad\text{ outside $\bar{Q}_{\phi}$}

Now define a function θ1=θ\theta_{1}=\theta if θ1\theta\neq 1 and θ1=0\theta_{1}=0 if θ=1\theta=1. Let θ0\theta_{0} denote the function of θθ1\theta-\theta_{1}.
Set F:={pΩ×(,b]:θ(p)=1}F:=\{p\in\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b]:\theta(p)=1\}. It is not hard to see that θ0=λF\theta_{0}=\lambda_{F} and

(4.9) [[F]]=τ(Ω×(,b),θ0)[[F]]=\tau(\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b),\theta_{0})

As a result one has the following decomposition

(4.10) [[E]]+R=[[F]]+G[[E]]+R=[[F]]+G

where GG is the integer multiplicity current τ(Ω×(,b),θ1)\tau(\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b),\theta_{1}). By (4.8) and the definition of θ1\theta_{1}, θ1=0\theta_{1}=0 outside Q¯ϕ\bar{Q}_{\phi}. As a result spt(G)Q¯ϕspt(G)\subset\bar{Q}_{\phi}.
Now define Uj={pΩ×(,b):θ(p)j}U_{j}=\{p\in\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}(-\infty,b):\theta(p)\geq j\} for any integer jj. By the definition of EE, we have spt([[Uj]])Q¯ϕspt(\partial[[U_{j}]])\subset\bar{Q}_{\phi} for any j1j\neq 1. Note that spt(Ta,b)Ω¯×[a,b]Q¯ϕspt(T_{a,b})\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]\subset\bar{Q}_{\phi}. Applying the decomposition theorem of integer multiplicity currents [34, corollary 27.8] we obtain

μTa,b\displaystyle\mu_{T_{a,b}} =j=,j1μ[[Uj]]+μ[[U1]]|Q¯ϕ\displaystyle=\sum_{j=-\infty,j\neq 1}^{\infty}\mu_{\partial[[U_{j}]]}+\mu_{\partial[[U_{1}]]}|_{\bar{Q}_{\phi}}
=j=,j1μ[[Uj]]|Ω¯×[a,b]+μ[[U1]]|Ω¯×[a,b]\displaystyle=\sum_{j=-\infty,j\neq 1}^{\infty}\mu_{\partial[[U_{j}]]}|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]}+\mu_{\partial[[U_{1}]]}|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]}

This implies that

(4.11) spt(μ[[Uj]])Ω¯×[a,b],j1,spt(μ[[U1]])|Q¯ϕΩ¯×[a,b]spt(\mu_{\partial[[U_{j}]]})\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b],j\neq 1,spt(\mu_{\partial[[U_{1}]]})|_{\bar{Q}_{\phi}}\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]

For G=τ(M,θ1)G=\tau(M,\theta_{1}) applying the decomposition theorem [34, corollary 27.8] again gives that

(4.12) μG=j=3μ[[Uj]]+2μ[[U2]]+j=0μ[[Uj]],\mu_{\partial G}=\sum_{j=3}^{\infty}\mu_{\partial[[U_{j}]]}+2\mu_{\partial[[U_{2}]]}+\sum_{j=-\infty}^{0}\mu_{\partial[[U_{j}]]},

By (4.11) spt(G)Ω¯×[a,b]spt(\partial G)\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]. As for FF the decomposition theorem gives that [[F]]=[[U1\U2]]\partial[[F]]=\partial[[U_{1}\backslash U_{2}]]. Since U2U1U_{2}\subset U_{1}, with (4.7) and (4.10), [14, Lemma 3.17] implies that

(4.13) μ[[F]]μ[[U1]]+μ[[U2]]\mu_{\partial[[F]]}\leq\mu_{\partial[[U_{1}]]}+\mu_{\partial[[U_{2}]]}

and spt([[F]])|Q¯ϕΩ¯×[a,b]spt(\partial[[F]])|_{\bar{Q}_{\phi}}\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b] from (4.11). Thus 𝕄([[F]]|Q¯ϕ)𝕄(Ta,b)<\mathbb{M}(\partial[[F]]|_{\bar{Q}_{\phi}})\leq\mathbb{M}(T_{a,b})<\infty. From (4.10),

(4.14) Ta,b=[[F]]|Q¯ϕ+G=[[F]]|Ω¯×[a,b]+GT_{a,b}=\partial[[F]]|_{\bar{Q}_{\phi}}+\partial G=\partial[[F]]|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[a,b]}+\partial G

and ([[F]]|Q¯ϕ)=Ta,b=Γ\partial(\partial[[F]]|_{\bar{Q}_{\phi}})=\partial T_{a,b}=\Gamma. Since Ta,bT_{a,b} solves (4.1), (4.13) implies that we can choose Ta,bT_{a,b} as [[F]]|Q¯ϕ\partial[[F]]|_{\bar{Q}_{\phi}}. By (4.10) and (4.9) FF coincides with the subgraph of ψ(x)\psi(x) outside Q¯ϕ\bar{Q}_{\phi}. The proof is complete. ∎

4.2. Regularity of almost minimal boundary

In this subsection we recall some results on the regularity of the almost minimal boundary from [14, section 3.2]. All of their proofs are skipped here. We refer the reader to [18], [37] and [10] for more details. Throughout this subsection let MM be a n+1n+1-dimensional Riemannian manifold.
Suppose TT is a (n+1)(n+1)-dimensional integer multiplicity current in MM, represented as τ(V,θ)\tau(V,\theta) where VV is a Ln+1L^{n+1} measurable set of MM. The relationship between the mass and BV functions can be summarized as follows:

(4.15) 𝕄W(T)=DθM(W);\mathbb{M}_{W}(\partial T)=||D\theta||_{M}(W);

for any open set WMW\subset\subset M. In particular if EE is a Caccioppoli set in WW and T=[[E]]T=[[E]] we have

(4.16) 𝕄W([[E]])=DλEM(W)=P(E,W)\mathbb{M}_{W}(\partial[[E]])=||D\lambda_{E}||_{M}(W)=P(E,W)

where PP is the perimeter in MM. For a derivation see [34, subsection 27.7].
For a point pMp\in M and r>0r>0 we denote by Br(p)B_{r}(p) the open ball centered at pp with radius rr . Now we define an almost minimal set in an open set and a closed set respectively.

Definition 4.13.

Let Ω\Omega be a domain. Fix μ0>0\mu_{0}>0 and α(0,12]\alpha\in(0,\frac{1}{2}]. Suppose EΩE\subset\Omega is a Caccioppoli set in MM.

  1. (1)

    We say that EE is an (μ0,α)(\mu_{0},\alpha)-almost minimal set in Ω\Omega if there is an r0>0r_{0}>0 and a constant μ0\mu_{0} with the property that for any r<r0r<r_{0}, xΩx\in\Omega with Br(p)ΩB_{r}(p)\subset\subset\Omega,

    (4.17) P(E,Br(x))P(F,Br(x))+μ0rn+2αP(E,B_{r}(x))\leq P(F,B_{r}(x))+\mu_{0}r^{n+2\alpha}

    where FF is any Caccioppoli set satisfying EΔFBr(p)E\Delta F\subset B_{r}(p). In particular if C=0C=0, we say FF is a minimal set in Ω\Omega.

  2. (2)

    We say that EE is an (μ0,α)(\mu_{0},\alpha)-almost minimal set in Ω¯\bar{\Omega} if there is an r0>0r_{0}>0 and a constant μ0\mu_{0} with the property that for any r<r0r<r_{0}, xΩ¯x\in\bar{\Omega}

    (4.18) P(E,Br(x))DλFΩM(Br(x))+μ0rn+2αP(E,B_{r}(x))\leq||D\lambda_{F\cap\Omega}||_{M}(B_{r}(x))+\mu_{0}r^{n+2\alpha}

    where FF is any Caccioppoli set satisfying EΔFBr(p)E\Delta F\subset B_{r}(p). In particular if C=0C=0, we say FF is a minimal set in Ω\Omega.

  3. (3)

    The regular set of E\partial E is the set {pE:E\{p\in\partial E:\partial E is a C1,αC^{1,\alpha} graph in a ball containing p}p\}. The singular set of E\partial E is the complement of the regular set in E\partial E.

Remark 4.14.

By [41, Lemma 7.6] all C2C^{2} bounded domains are almost minimal sets in an open neighborhood of their boundaries.

A good property of almost minimal sets in a domain is their boundary regularity.

Theorem 4.15 (Theorem 1 in [37], Theorem 5.6 in [10]).

Suppose a Caccioppoli set EE is a (μ0,α)(\mu_{0},\alpha)-almost minimal set in a domain Ω\Omega. Let SS be the singular set of E\partial E in Ω\Omega. Then

  1. (1)

    if m7m\leq 7, S=S=\emptyset;

  2. (2)

    if m=7m=7, SS consists of isolated points;

  3. (3)

    if m>7m>7, Ht(S)=0H^{t}(S)=0 for any t>m7t>m-7. Here HH denotes the Hausdorff measure in MM.

where m=dimMm=dimM.

Remark 4.16.

Note that in general the boundary of almost minimal sets in closed sets does not have such good regularity.

The following result is very important.

Theorem 4.17 (Theorem 3.19 in [14]).

Let Ω1\Omega_{1} and Ω2\Omega_{2} be two C2C^{2} domains in MM. Define Ω=Ω1Ω2\Omega^{\prime}=\Omega_{1}\cap\Omega_{2}. Fix a point pp in Ω1Ω2\partial\Omega_{1}\cap\partial\Omega_{2}. Suppose EΩE\subset\Omega^{\prime} is an (μ0,α0)(\mu_{0},\alpha_{0})- almost minimal set in Ω¯\bar{\Omega}^{\prime} (the closure of Ω\Omega^{\prime}) and E\partial E passes through pp, then E\partial E is a C1,αC^{1,\alpha^{\prime}} graph in an open ball containing pp for some α(0,1)\alpha^{\prime}\in(0,1).

Remark 4.18.

In general the boundary of Ω\Omega^{\prime} is not C2C^{2}.

4.3. The proof of Theorem 4.1

Proof.

Set α0:=minΩψ(x)\alpha_{0}:=\min_{\partial\Omega}\psi(x) and α1:=maxΩψ(x)\alpha_{1}:=\max_{\partial\Omega}\psi(x). Fix n>0n>0 such that n<α0α1<n-n<\alpha_{0}\leq\alpha_{1}<n. By Lemma 4.11 there is a Caccioppoli set in Ω×\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} such that Tn=[[Fn]]|Ω¯×[n,n]T_{n}=\partial[[F_{n}]]|_{\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[-n,n]} realizes the minimum in (4.6) and FnF_{n} is the subgraph of ψ(x)\psi(x) outside Q¯ϕ\bar{Q}_{\phi}.
We claim that TnT_{n} has to be contained in the closed set Ω¯×[α0,α1]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[\alpha_{0},\alpha_{1}]. If not the case, assume

(4.19) r0:=max{rΩ¯:p=(x,r)Tn=[[Fn]]|Q¯ϕ}(α1,n]r_{0}:=\max\{r\in\bar{\Omega}:p=(x,r)\in T_{n}=\partial[[F_{n}]]|_{\bar{Q}_{\phi}}\}\in(\alpha_{1},n]

Let p0p_{0} be the point on TnT_{n} achieving this maximum. Since r0>α1r_{0}>\alpha_{1}, there is an embedded ball Br(p0)B_{r}(p_{0}) such that Br(p0)Γ=B_{r}(p_{0})\cap\Gamma=\emptyset.
The first case is that p0=(x0,r0)p_{0}=(x_{0},r_{0}) for some x0Ωx_{0}\in\partial\Omega. Note that FnF_{n} is an minimal set in Ω¯×Br(p0)\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}\cap B_{r}(p_{0}) By Theorem 4.17 Fn\partial F_{n} is still a C1,αC^{1,\alpha^{\prime}} graph near p0p_{0} for some α(0,1)\alpha^{\prime}\in(0,1). Moreover Fn\partial F_{n} is tangent to Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} and Ω×{r0}\Omega{\mkern-1.0mu\times\mkern-1.0mu}\{r_{0}\} at p0p_{0}. But Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} and Ω×{r0}\Omega{\mkern-1.0mu\times\mkern-1.0mu}\{r_{0}\} is orthogonally transverse at p0p_{0}. This gives a contradiction and the first case is impossible.
The second case is p0=(x0,r0)p_{0}=(x_{0},r_{0}) for some x0Ωx_{0}\in\Omega. By remark 4.12 [[Fn]]|Q¯ϕΩ¯×[n,n]\partial[[F_{n}]]|_{\bar{Q}_{\phi}}\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[-n,n]. Thus FnF_{n} is a minimal set in Ω¯×[n,n]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[-n,n] away from Γ\Gamma. Note that Ω×[n,n]\Omega{\mkern-1.0mu\times\mkern-1.0mu}[-n,n] be the intersection of two C2C^{2} domain. Again by Theorem 4.17 Fn\partial F_{n} is a C1,αC^{1,\alpha^{\prime}} graph in a neighborhood of p0p_{0} contained in Br(p0)B_{r}(p_{0}). But we should observe that Ω×{r0}\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\{r_{0}\} is also minimal in MϕM_{\phi} by Theorem 2.3. With respect to the upward normal vector, HFn0H_{\partial F_{n}}\leq 0 near p0p_{0} in the Lipschitz sense. By the maximum principle in [14, Theorem A.1] Fn\partial F_{n} coincides with Ω×{r0}\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\{r_{0}\} near p0p_{0}. From the connectedness of Ω×{r0}\Omega{\mkern-1.0mu\times\mkern-1.0mu}\{r_{0}\} we obtain that Ω×{r0}Fn\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\{r_{0}\}\subset\partial F_{n}. This is impossible since Fn\partial F_{n} is contained in Ω¯×[n,n]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[-n,n]. Thus the second case is also impossible. This means r0α1r_{0}\leq\alpha_{1}. Arguing a similar derivation we will obtain that

(4.20) min{rΩ¯:p=(x,r)Tn=[[F]]|Q¯ϕ}α0\min\{r\in\bar{\Omega}:p=(x,r)\in T_{n}=\partial[[F]]|_{\bar{Q}_{\phi}}\}\geq\alpha_{0}

Thus the above claim is true.
For each nn with n<α0α1<n-n<\alpha_{0}\leq\alpha_{1}<n, TnT_{n} is contained in Ω¯×[α0,α1]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[\alpha_{0},\alpha_{1}] and 𝕄(Tn)\mathbb{M}(T_{n}) is uniformly bounded and Tn=Γ\partial T_{n}=\Gamma. By Theorem 4.7 as nn goes to \infty the sequence {Tn}nmax{|α0|,|α1|}\{T_{n}\}_{n\geq\max\{|\alpha_{0}|,|\alpha_{1}|\}} will converge to a TT_{\infty} such that T|Q¯ϕΩ¯×[α0,α1]T_{\infty}|_{\bar{Q}_{\phi}}\subset\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[\alpha_{0},\alpha_{1}] and

𝕄(T):=min{𝕄(T):T=Γ,T𝒢}\mathbb{M}(T_{\infty}):=\min\{\mathbb{M}(T):\partial T=\Gamma,T\in\mathcal{G}\}

Moreover from the compactness of BV functions in Theorem 2.10 there is a Caccioppoli set FF_{\infty} such that T=[[F]]|Q¯ϕT_{\infty}=\partial[[F_{\infty}]]|_{\bar{Q}_{\phi}} which is contained in Ω¯×[α0,α1]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[\alpha_{0},\alpha_{1}] and FF_{\infty} is the subgraph of ψ(x)\psi(x) outside Q¯ϕ\bar{Q}_{\phi}. By Theorem 3.10 there is a BV function u(x)BV(Ω)u(x)\in BV(\Omega^{\prime}) such that u(x)=ψ(x)u(x)=\psi(x) outside Ω\Omega and

(4.21) Pϕ(U,Ω×)Pϕ(F,Ω×)P_{\phi}(U,\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})\leq P_{\phi}(F_{\infty},\Omega^{\prime}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})

where UU is the subgraph of u(x)u(x) and

(4.22) α0u(x)α1 on Ωu(x)=ψ(x) on Ω\Ω\alpha_{0}\leq u(x)\leq\alpha_{1}\text{ on }\Omega\quad u(x)=\psi(x)\text{ on }\Omega^{\prime}\backslash\Omega

Moreover ([[U]]|Q¯ϕ)=Γ\partial(\partial[[U]]|_{\bar{Q}_{\phi}})=\Gamma by remark 4.3. Since FF_{\infty} coincides with UU outside Q¯ϕ\bar{Q}_{\phi}, then

(4.23) 𝕄([[U]]|Q¯ϕ)𝕄(T)\mathbb{M}(\partial[[U]]|_{\bar{Q}_{\phi}})\leq\mathbb{M}(T_{\infty})

Thus [[U]]|Q¯ϕ\partial[[U]]|_{\bar{Q}_{\phi}} also solves the area minimizing problem (4.1). This shows the existence of u(x)u(x) with the property (2).
Now fix any v(x)BV(Ω)v(x)\in BV(\Omega^{\prime}) satisfying v(x)=ψ(x)v(x)=\psi(x) outside Ω¯\bar{\Omega}. Let VV be the subgraph of v(x)v(x). By remark 4.3 ([[V]]|Q¯ϕ)=Γ\partial(\partial[[V]]|_{\bar{Q}_{\phi}})=\Gamma. By the property (2)

(4.24) 𝕄([[U]]|Q¯ϕ)𝕄([[V]]|Q¯ϕ)\mathbb{M}(\partial[[U]]|_{\bar{Q}_{\phi}})\leq\mathbb{M}(\partial[[V]]|_{\bar{Q}_{\phi}})

Because u(x)=ψ(x)=v(x)u(x)=\psi(x)=v(x) outside Ω¯\bar{\Omega}, by Theorem 3.9 and (4.16) we obtain

(4.25) 𝔉ϕ(u,Ω)𝔉ϕ(v,Ω)\mathfrak{F}_{\phi}(u,\Omega^{\prime})\leq\mathfrak{F}_{\phi}(v,\Omega^{\prime})

Thus u(x)u(x) realizes the minimum of

(4.26) min{𝔉ϕ(v(x),Ω):v(x)BV(Ω),v(x)=ψ(x) outside Ω¯}\min\{\mathfrak{F}_{\phi}(v(x),\Omega^{\prime}):v(x)\in BV(\Omega^{\prime}),v(x)=\psi(x)\text{ outside }\bar{\Omega}\}

We conclude that u(x)u(x) statisfies the property (1).
At last we show that u(x)u(x) has the property (3). By the property (2) UU is a minimal set in Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. By Theorem 4.15, except a closed set SS with Hn6(S)=0H^{n-6}(S)=0, U\S\partial U\backslash S is connected and C1,αC^{1,\alpha} for some α(0,1)\alpha\in(0,1). Since U\S\partial U\backslash S is minimal in Ω×Mϕ\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}\subset M_{\phi} and ϕ(x)\phi(x) is C2C^{2}, the regularity of minimal surface equation (2.3) implies that it is C2,βC^{2,\beta} for any β(0,1)\beta\in(0,1). Let v\vec{v} be the normal vector of U\S\partial U\backslash S in Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} pointing to the positive infinity. Define Θ=v,r\Theta=\langle\vec{v},\partial_{r}\rangle. Thus Θ0\Theta\geq 0 on U\S\partial U\backslash S in Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} which is connected and C2,βC^{2,\beta}. By the Harnack principle in corollary A.2, we have Θ0\Theta\equiv 0 or Θ>0\Theta>0 on U\S\partial U\backslash S. Since U|Q¯ϕ\partial U|_{\bar{Q}_{\phi}} is contained in Ω¯×[α0,α1]\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}[\alpha_{0},\alpha_{1}] and Hn1(S)=0H^{n-1}(S)=0, only Θ>0\Theta>0 can happen in U\S\partial U\backslash S.
Let SS^{\prime} be the orthonormal projection of SS in Ω\Omega. Therefore Hn1(S)=0H^{n-1}(S^{\prime})=0 and uC2,β(Ω\S)u\in C^{2,\beta}(\Omega\backslash S^{\prime}). Moreover

(4.27) Ω\Sϕn(x)1+|Du|2𝑑vol𝔉ϕ(u,Ω)<\int_{\Omega\backslash S^{\prime}}\phi^{n}(x)\sqrt{1+|Du|^{2}}dvol\leq\mathfrak{F}_{\phi}(u,\Omega)<\infty

This implies that uW1,1(Ω)u\in W^{1,1}(\Omega). By the property (1) in Theorem 4.1 u(x)u(x) is also the critical point of the functional

(4.28) Ωϕn(x)1+|Dv|2𝑑vol for vW1,1(Ω)\int_{\Omega}\phi^{n}(x)\sqrt{1+|Dv|^{2}}dvol\text{ for }v\in W^{1,1}(\Omega)

From the removable singularity result of Simon [33, Theorem 1] u(x)u(x) is regular at every xx in Ω\Omega. Thus u(x)u(x) satisfies (2.3) in the Lipschitz sense. Since ϕ(x)C2\phi(x)\in C^{2}, by the classical regularity of uniformly elliptic equations u(x)u(x) is C2,β(Ω)C^{2,\beta}(\Omega) for any β(0,1)\beta\in(0,1). The proof is complete. ∎

5. the Dirichlet problem

In this section we apply Theorem 4.1 to solve the Dirichlet problem of minimal surface equations in MϕM_{\phi}.

Definition 5.1.

We say that Ω\Omega is ϕ\phi-mean convex if the mean curvature of Ω\partial\Omega satisfies HΩ+nDlogϕ(x),γ0H_{\partial\Omega}+n\langle D\log\phi(x),\gamma\rangle\geq 0. Here DD denotes the covariant derivative of NN and HΩ=div(γ)H_{\partial\Omega}=div(\gamma) for the outward normal vector γ\gamma on Ω\partial\Omega.

Remark 5.2.

By (2.2) (see also [40, Lemma 3.1]) Ω\Omega is ϕ\phi-mean convex if and only if Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is mean convex in MϕM_{\phi} with respect to the metric ϕ2(x)(σ+dr2)\phi^{2}(x)(\sigma+dr^{2}) for its outward normal vector.

5.1. The Dirichlet problem on ϕ\phi-mean convex domain

The first result of this section is given as follows.

Theorem 5.3.

Let Ω\Omega be a C2C^{2} bounded ϕ\phi-mean convex domain in NN and let ϕ(x)\phi(x) be a C2C^{2} positive function in NN. For any ψ(x)C(Ω)\psi(x)\in C(\partial\Omega) the Dirichlet problem

(5.1) {L(u)=div(Duω)+nDlogϕ(x),Duω=0, on Ωu(x)=ψ(x) on Ω\left\{\begin{split}L(u)=-div(\frac{Du}{\omega})&+n\langle D\log\phi(x),-\frac{Du}{\omega}\rangle=0,\text{ on }\Omega\\ &u(x)=\psi(x)\quad\text{ on }\partial\Omega\end{split}\right.

has a unique solution in C(Ω¯)C2(Ω)C(\bar{\Omega})\cap C^{2}(\Omega). Here ω=1+|Du|2\omega=\sqrt{1+|Du|^{2}}.

Remark 5.4.

Suppose Ω\Omega is C2,αC^{2,\alpha} for some α(0,1)\alpha\in(0,1). Casteras-Heinonen-Holopainen [8, Theorem 2] showed that if there is a positive constant F>0F>0 such that

  1. (1)

    m(x)C2(N)m(x)\in C^{2}(N), r(t)C1()r(t)\in C^{1}(\mathbb{R}) satisfying

    maxΩ¯|Dm(x)|+maxt|r(t)|F\max_{\bar{\Omega}}|Dm(x)|+\max_{t\in\mathbb{R}}|r^{\prime}(t)|\leq F
  2. (2)

    the Ricci curvature of Ω\Omega satisfies RicΩF2nRic_{\Omega}\geq-\frac{F^{2}}{n} and HΩFH_{\partial\Omega}\geq F

Then the Dirichlet problem

(5.2) div(Duω)+D(m(x)),Duω+r(u(x))1ω=0 on Ω-div(\frac{Du}{\omega})+\langle D(m(x)),-\frac{Du}{\omega}\rangle+r^{\prime}(u(x))\frac{1}{\omega}=0\text{ on }\Omega

with u(x)=ψ(x)u(x)=\psi(x) on Ω\partial\Omega for any ψ(x)C(Ω)\psi(x)\in C(\partial\Omega) has a solution in C2,α(Ω)C(Ω¯)C^{2,\alpha}(\Omega)\cap C(\bar{\Omega}).
In the case of r(t)0r(t)\equiv 0 Theorem 5.3 removes the curvature assumption on Ω\Omega in [8, Theorem 2]. Our mean curvature assumption on Ω\partial\Omega should be optimal. For example when ϕ(x)1\phi(x)\equiv 1 this is confirmed by Serrin [32] in Euclidean spaces.

Remark 5.5.

The above result can not be obtained by the continuous method in [16, section 11.2, 11.3, Chapter 18]. The reason is that the boundary assumption may not be preserved by these methods.

Proof.

The uniqueness of the Dirichlet problem (5.1) is obvious by the maximum principle. We only need to show the existence.
First we assume that ψ(x)C3(Ω)\psi(x)\in C^{3}(\partial\Omega). Without loss generality we can assume ψ(x)C1(Ω\Ω)\psi(x)\in C^{1}(\Omega^{\prime}\backslash\Omega) for some Ω\Omega^{\prime} strictly containing Ω\Omega. Let QϕQ_{\phi} be the set Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} and Q¯ϕ\bar{Q}_{\phi} be its closure. By Theorem 4.1 there is a u(x)BV(Ω)u(x)\in BV(\Omega^{\prime}) with u(x)=ψ(x)u(x)=\psi(x) outside Ω\Omega such that T=[[U]]|Q¯ϕT=\partial[[U]]|_{\bar{Q}_{\phi}} realizes the area minimizing problem (4.1). Moreover u(x)C2(Ω)u(x)\in C^{2}(\Omega). As a result UQϕ\partial U\cap Q_{\phi} is a minimal graph in QϕQ_{\phi}. By the conclusion (1) in Theorem 2.3 u(x)u(x) satisfies Lu=0Lu=0 on Ω\Omega. It only suffices to show that u(x)C(Ω¯)u(x)\in C(\bar{\Omega})
Fix any zΩz\in\partial\Omega. Now define

(5.3) A:=sup{limnsupknu(xk):{xn}n=1Ω,limndist(xn,z)=0}A:=sup\{\lim_{n}\sup_{k\geq n}u(x_{k}):\{x_{n}\}_{n=1}^{\infty}\in\Omega,\lim_{n\rightarrow\infty}dist(x_{n},z)=0\}

where distdist is the distance function in NN. By Theorem 4.1 minΩψu(x)maxΩψ\min_{\partial\Omega}\psi\leq u(x)\leq\max_{\partial\Omega}\psi for any xΩx\in\Omega. Thus AA is a finite number. Suppose A>ψ(z)A>\psi(z). There is a sequence {xn}n=1\{x_{n}\}_{n=1}^{\infty} in Ω\Omega such that limn+xn=z\lim_{n\rightarrow+\infty}x_{n}=z and limn+ψ(xn)=A\lim_{n\rightarrow+\infty}\psi(x_{n})=A. Since ψ(x)\psi(x) is continuous at zz, there is a neighborhood VV of the point (z,A)(z,A) such that VV is disjoint with Γ\Gamma, the graph of ψ(x)\psi(x) in Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}.
By Theorem 4.1 UU is a minimal set in VQ¯ϕV\cap\bar{Q}_{\phi} and passes through the point (z,A)(z,A). Since Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is C2C^{2}, Theorem 4.17 implies that U\partial U is a C1,αC^{1,\alpha} graph near (z,A)(z,A). Since UU is a minimal set in VQ¯ϕV\cap\bar{Q}_{\phi}, then HU0H_{\partial U}\leq 0 near (z,A)(z,A) with respect to the outward normal vector of Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} in the Lipschitz sense. Since Ω\Omega is ϕ\phi-mean convex, by remark 5.2 Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is a mean convex domain in MϕM_{\phi}. Because UVΩ×U\cap V\subset\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} and is tangent to Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. From the maximum principle in [14, Theorem A.1], U\partial U coincides with Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} in VV. This contradicts to the definition of AA. Thus

(5.4) limsupu(xn)ψ(z),{xn}n=1Ωlimnxn=z\lim\sup u(x_{n})\leq\psi(z),\forall\{x_{n}\}_{n=1}^{\infty}\in\Omega\,\lim_{n\rightarrow\infty}x_{n}=z

With a similar derivation we also obtain

(5.5) liminfu(xn)ψ(z),{xn}n=1Ωlimnxn=z\lim\inf u(x_{n})\geq\psi(z),\forall\{x_{n}\}_{n=1}^{\infty}\in\Omega\,\lim_{n\rightarrow\infty}x_{n}=z

Combining the above two facts together yields that u(x)u(x) is continous for any fixed zΩz\in\partial\Omega, thus u(x)C(Ω¯)u(x)\in C(\bar{\Omega}). As a result we obtain the existence of the Dirichlet problem (5.1) when ψ(x)C3(Ω)\psi(x)\in C^{3}(\partial\Omega).
Now we consider the general case of ψ(x)C(Ω)\psi(x)\in C(\partial\Omega). Then there are two sequences {ψk±(x)}k=1\{\psi^{\pm}_{k}(x)\}_{k=1}^{\infty} in C3(Ω)C^{3}(\partial\Omega) such that

(5.6) ψk(x)ψk+1(x)ψ(x)ψl+1+(x)ψl+(x)\cdots\psi^{-}_{k}(x)\leq\psi^{-}_{k+1}(x)\leq\cdots\psi(x)\leq\cdots\leq\psi^{+}_{l+1}(x)\leq\psi^{+}_{l}(x)\leq\cdots

and both of them converge to ψ(x)\psi(x) in the C0(Ω)C^{0}(\partial\Omega). By the previous argument for any positive integer kk let uk±(x)u^{\pm}_{k}(x) be the solution of the Dirichlet problem (5.1) on Ω\Omega with uk(x)=ψk±(x)u_{k}(x)=\psi^{\pm}_{k}(x) on Ω\partial\Omega. Applying the maximum principle on (5.1), we obtain

  1. (1)

    maxxΩ¯|uk(x)|C\max_{x\in\bar{\Omega}}|u_{k}(x)|\leq C where C=maxk=1,,xΩ{ψ(x),ψk±(x)}C=\max_{k=1,\cdots,x\in\partial\Omega}\{\psi(x),\psi^{\pm}_{k}(x)\};

  2. (2)

    uk(x)uk+1(x)ul+1+(x)ul+(x)u^{-}_{k}(x)\leq u^{-}_{k+1}(x)\leq\cdots\leq u^{+}_{l+1}(x)\leq u^{+}_{l}(x)\cdots on Ω\Omega;

According to Theorem A.3, {uk±(x)}k=1\{u^{\pm}_{k}(x)\}_{k=1}^{\infty} has a local uniformly bound of their C2C^{2} norms in Ω\Omega. Now we can choose a subsequence, still written as {uk(x)}k=1\{u^{-}_{k}(x)\}_{k=1}^{\infty} such that limkuk(x)=u(x)\lim_{k\rightarrow\infty}u^{-}_{k}(x)=u(x) on Ω¯\bar{\Omega} and u(x)C2(Ω)u(x)\in C^{2}(\Omega) and satisfies Lu=0Lu=0 on Ω\Omega.
The only thing to left is to show that u(x)C(Ω¯)u(x)\in C(\bar{\Omega}) and u(x)=ψ(x)u(x)=\psi(x) on Ω\partial\Omega. By the conclusion (2) above applying the maximum principle on Lu=0Lu=0 yields that

(5.7) uk(x)u(x)ul+(x) on Ωu^{-}_{k}(x)\leq u(x)\leq u^{+}_{l}(x)\text{ on }\Omega

for any positive integers k,lk,l. This implies that for any zΩz\in\partial\Omega

(5.8) ψk(z)limxz,xΩinfu(x)limxz,xΩsupu(x)ψl+(z)\psi^{-}_{k}(z)\leq\lim_{x\rightarrow z,x\in\Omega}\inf u(x)\leq\lim_{x\rightarrow z,x\in\Omega}\sup u(x)\leq\psi^{+}_{l}(z)

Letting k,lk,l go to the positive infinity we obtain limxz,xΩu(x)=ψ(z)\lim_{x\rightarrow z,x\in\Omega}u(x)=\psi(z) for any zΩz\in\partial\Omega. Thus we define u(z)=ψ(z)u(z)=\psi(z) on Ω\partial\Omega and u(x)C(Ω¯)u(x)\in C(\bar{\Omega}). The proof is complete. ∎

5.2. The uniqueness of area minimizing currents

Now by Theorem 5.3 we shall obtain a uniqueness result for the area minimizing problem (4.1).

Theorem 5.6.

Suppose Ω\Omega is a ϕ\phi-mean convex C2C^{2} domain and ψ(x)C1(Ω)\psi(x)\in C^{1}(\partial\Omega). Set Γ={(x,ψ(x)):xΩ}\Gamma=\{(x,\psi(x)):x\in\partial\Omega\} . Let u(x)u(x) be the solution of the Dirichlet problem (5.1) with boundary data ψ(x)\psi(x). Let U(x)U(x) be the subgraph of u(x)u(x).

  1. (1)

    Then T=[[U]]|Q¯ϕT=\partial[[U]]|_{\bar{Q}_{\phi}} is the unique n-integer multiplicity current in 𝒢\mathcal{G} to realize

    (4.1) min{𝕄(T):T𝒢,T=Γ}\min\{\mathbb{M}(T):T\in\mathcal{G},\partial T=\Gamma\}
  2. (2)

    Let Ω\Omega^{\prime} be a domain with ΩΩ\Omega\subset\subset\Omega^{\prime}. Extend ψ(x)\psi(x) as a function C1(Ω\Ω)C^{1}(\Omega^{\prime}\backslash\Omega). Let u(x)=ψ(x)u(x)=\psi(x) outside Ω\Omega. Then u(x)u(x) is a unique function to realize

    (4.2) min{𝔉ϕ(v(x),Ω):v(x)=ψ(x) outside Ω}\min\{\mathfrak{F}_{\phi}(v(x),\Omega^{\prime}):v(x)=\psi(x)\text{ outside $\Omega$}\}
Remark 5.7.

This theorem is a finite version of [22, Theorem 4.1] in finite conformal cones in MϕM_{\phi}.

Proof.

By Theorem 5.3, let TT be a n-integer multiplicity current with compact support in 𝒢\mathcal{G} to realize (4.1). For any tt\in\mathbb{R}, let UtU_{t} be the subgraph of u(x)+tu(x)+t over Ω\Omega. By (5.1), [[Ut]]\partial[[U_{t}]] is a minimal graph over Ω\Omega in MϕM_{\phi} and the boundary of [[Ut]]\partial[[U_{t}]] is disjoint with Γ\Gamma if t0t\neq 0.
Now the following number

(5.9) t0=inf{t>0:[[Ut]]T=:s[t,+)}t_{0}=\inf\{t>0:\partial[[U_{t}]]\cap T=\emptyset:\forall\quad s\in[t,+\infty)\}

is well-defined since spt(T)spt(T) is compact. Suppose t0>0t_{0}>0. Then [[Ut0]]\partial[[U_{t_{0}}]] is tangent to TT at some point pΩ×p\in\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}.
By Lemma 4.11, T=([[F]])|Q¯ϕT=(\partial[[F]])|_{\bar{Q}_{\phi}} where FF is a minimal set in the set Ω¯×\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. From the definition of t0t_{0}, FΩ×Ut0F\cap\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}\subset U_{t_{0}}. By Theorem 4.17 F\partial F is C1,αC^{1,\alpha} near pp. In the following let HH denote the mean curvature. Thus HF0H_{\partial F}\leq 0 near pp with respect to the upward normal vector in the Lipschitz sense. Note that HUt0=0H_{\partial U_{t_{0}}}=0 near pp. Then the maximum principle in [14, Theorem A.1] implies that F\partial F coincides with Ut0\partial U_{t_{0}} near pp. By the connectedness of Ut0\partial U_{t_{0}}, we have Ut0T\partial U_{t_{0}}\subset T. Thus for some x0Ωx_{0}\in\partial\Omega, the point p0=(x0,u(x0)+t0)p_{0}=(x_{0},u(x_{0})+t_{0}) is contained in TT.
Since t0>0t_{0}>0, there is a neighborhood VV of p0p_{0} disjoint with Γ\Gamma. Then FF is a minimal set in VΩ¯×V\cap\bar{\Omega}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. By Remark 5.2, Ω×\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is mean convex in MϕM_{\phi}. Again by [14, Theorem A.1], Ω×\partial\Omega{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} coincides with TT near p0p_{0}. With the connectedness, the set {(x,r):xΩ,r>u(x)+t0}\{(x,r):x\in\partial\Omega,r>u(x)+t_{0}\} is contained in TT. This is impossible since spt(T)spt(T) is compact. Thus t0=0t_{0}=0 and spt(T){(x,r):xΩ¯,ru(x)}spt(T)\subset\{(x,r):x\in\bar{\Omega},r\leq u(x)\}. With a similar derivation, we can show that spt(T){(x,r):xΩ¯,ru(x)}spt(T)\subset\{(x,r):x\in\bar{\Omega},r\geq u(x)\}. Finally combining the above two facts yields that T=[[U]]|Q¯ϕT=\partial[[U]]|_{\bar{Q}_{\phi}} where UU is the subgraph of u(x)u(x). We obtain the conclusion (1).
Suppose there are two functions u1,u2u_{1},u_{2} in BV(Ω)BV(\Omega^{\prime}) to solve (4.2). Let UiU_{i} be the subgraph of ui(x)u_{i}(x) and Ti=[[Ui]]|Q¯ϕT_{i}=\partial[[U_{i}]]|_{\bar{Q}_{\phi}} for i=1,2i=1,2. By (2) in Theorem 4.1, TiT_{i} should solve (4.1). By the conclusion (1) we just obtained, T1=T2T_{1}=T_{2} and u1(x)=u2(x)u_{1}(x)=u_{2}(x) on Ω\Omega. We arrive the conclusion (2). The proof is complete. ∎

6. Infinity boundary cases

In this section we apply the results in previous sections to generalize the existence and the uniqueness of area minimizing graphs with the infinity star-shaped boundary in hyperbolic spaces.
Throughout this section, we assume that NN is a compact nn-dimensional Riemannian manifold with its metric σ\sigma and compact embedded C2C^{2} boundary N\partial N. We define

(6.1) Nr:={d(x)>r:xN}N_{r}:=\{d(x)>r:x\in N\}

where d(x)d(x) denotes d(x,N)d(x,\partial N), the distance between xx and Ω\partial\Omega. The following lemma is obvious.

Lemma 6.1.

Since N\partial N is C2C^{2} embedded and compact, there is a r0>0r_{0}>0 such that for any xN\Nr0x\in N\backslash N_{r_{0}}, there is a unique yNy\in\partial N such that d(x)d(x) is equal to d(x,y)d(x,y).

From now on fix r0r_{0}. Thus d(x)d(x) is C2C^{2} on N\Nr0N\backslash N_{r_{0}}. Suppose ϕ(x)\phi(x) is a positive C3C^{3} function on NN with

(6.2) ϕ(x)=h(d(x)) on N\Nr0limd(x)0ϕ(x)=limd(x)0h(d(x))=+\phi(x)=h(d(x))\text{ on }N\backslash N_{r_{0}}\quad\lim_{d(x)\rightarrow 0}\phi(x)=\lim_{d(x)\rightarrow 0}h(d(x))=+\infty

where h(r):(0,r0)+h(r):(0,r_{0})\rightarrow\mathbb{R}^{+} is a positive C2C^{2} function. Then

(6.3) Mϕ:=(N×,ϕ2(σ+dr2)))M_{\phi}:=(N{\mkern-1.0mu\times\mkern-1.0mu},\phi^{2}(\sigma+dr^{2})))

is a complete Riemannian manifold. A natural compactification is

(6.4) M¯ϕ=Mϕ(N×)\bar{M}_{\phi}=M_{\phi}\cup(\partial N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R})

equipped with the product metric topology of (N×,σ+dr2)(N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R},\sigma+dr^{2}).

Definition 6.2.

Let SS be a complete kk-integer multiplicity current in MϕM_{\phi}. Its infinity asymptotic boundary in N×\partial N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is the set S¯\S\bar{S}\backslash S where S¯\bar{S} is the closure of SS in the product metric topology.

Definition 6.3.

Let 𝕄\mathbb{M} be the mass of currents in MϕM_{\phi}. We say that TT is a local area minimizing nn-integer multiplicity current in MϕM_{\phi} if spt(T)spt(T) is contained in N×[a,b]N{\mkern-1.0mu\times\mkern-1.0mu}[a,b] for some finite interval [a,b][a,b], for any nn-integer multiplicity current TT^{\prime} satisfying T=TT=T^{\prime} outside a compact set in MϕM_{\phi}, then 𝕄(T)𝕄(T)\mathbb{M}(T)\leq\mathbb{M}(T^{\prime}).

The main result of this section is given as follows.

Theorem 6.4.

Let Nr,ϕ,r0N_{r},\phi,r_{0} be given in (6.1), (6.2) and Lemma 6.1 respectively. Suppose for any r(0,r0)r\in(0,r_{0}) NrN_{r} is ϕ\phi-mean convex. Fix any ψ(x)C1(N)\psi(x)\in C^{1}(\partial N). Let Γ\Gamma denote {(x,ψ(x)):xN}\{(x,\psi(x)):x\in\partial N\} in N×\partial N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}. Then there is a unique local area minimizing nn-integer multiplicity current TT in MϕM_{\phi} with the infinity boundary Γ\Gamma. Moreover TT is a graph over NN.

Remark 6.5.

Recall that n+1\mathbb{H}^{n+1} is the upper half space {(x,y):xn,y>0}\{(x,y):x\in\mathbb{R}^{n},y>0\} equipped with the metric

(6.5) dsH2=dx2+dy2y2 dx2 is the Euclidean metric in nds^{2}_{H}=\frac{dx^{2}+dy^{2}}{y^{2}}\quad\text{ $dx^{2}$ is the Euclidean metric in $\mathbb{R}^{n}$}

Let S+nS^{n}_{+} be the upper half hemisphere in n+1\mathbb{R}^{n+1} with the induced metric σn\sigma_{n}. We introduce the pole coordinate on S+nS^{n}_{+} (the pole at the north pole) (θ,φ)Sn1×[0,π2](\theta,\varphi)\in S^{n-1}{\mkern-1.0mu\times\mkern-1.0mu}[0,\frac{\pi}{2}]. Thus σn=dφ2+sin2(φ)dθ2\sigma_{n}=d\varphi^{2}+\sin^{2}(\varphi)d\theta^{2}. Note that dx2+dy2=s2σn+ds2dx^{2}+dy^{2}=s^{2}\sigma_{n}+ds^{2} for s(0,)s\in(0,\infty) and xn+1=scos(φ)x_{n+1}=s\cos(\varphi). Letting r=lnsr=\ln s we can represent Hyperbolic spaces n+1\mathbb{H}^{n+1} as

(6.6) (S+n×,1cos2(φ)(σn+dr2))(S^{n}_{+}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R},\frac{1}{\cos^{2}(\varphi)}(\sigma_{n}+dr^{2}))

which is a special case of (1.1).
The above theorem is to generalize Hardt-Lin’s result in Hyperbolic spaces n+1\mathbb{H}^{n+1} [22, Theorem 4.1]. We verify this claim as follows. Fixing any number φ0(0,π2)\varphi_{0}\in(0,\frac{\pi}{2}) we define a domain

(6.7) Sφ0:={(θ,φ)S+n:φ[0,φ0)}S_{\varphi_{0}}:=\{(\theta,\varphi)\in S^{n}_{+}:\varphi\in[0,\varphi_{0})\}

By [39, Proposition 2.1] the mean curvature of Sφ0\partial S_{\varphi_{0}} with respect to φ\frac{\partial}{\partial\varphi} is (n1)cos(φ0)sin(φ0)(n-1)\frac{\cos(\varphi_{0})}{\sin(\varphi_{0})}. By (2.2) the mean curvature of Sφ0×\partial S_{\varphi_{0}}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} in n+1\mathbb{H}^{n+1} is

(6.8) HSφ0×=cos(φ){(n1)cos(φ)sin(φ)+nDlog1cos(φ),φ}|φ=φ0=cos(φ0){(n1)cos(φ0)sin(φ0)+nsin(φ0)cos(φ0)}\begin{split}H_{\partial S_{\varphi_{0}}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}}&=\cos(\varphi)\{(n-1)\frac{\cos(\varphi)}{\sin(\varphi)}+n\langle D\log\frac{1}{\cos(\varphi)},\frac{\partial}{\partial\varphi}\rangle\}|_{\varphi=\varphi_{0}}\\ &=\cos(\varphi_{0})\{(n-1)\frac{\cos(\varphi_{0})}{\sin(\varphi_{0})}+n\frac{\sin(\varphi_{0})}{\cos(\varphi_{0})}\}\end{split}

Thus Sφ0×S_{\varphi_{0}}{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} is mean convex in MϕM_{\phi} and Sφ0S_{\varphi_{0}} is ϕ\phi-mean convex by Remark 5.2. Then n+1\mathbb{H}^{n+1} satisfies the condition in the above theorem. Thus [22, Theorem 4.1] is a speical case of Theorem 6.4.

Now we are ready to show Theorem 6.4.

Proof.

First we show the existence of a minimal graph in MϕM_{\phi} with the infinity boundary Γ\Gamma. The local area minimizing property can be easily obtained by Theorem 4.1.
Step one: Existence. First assume ψ(x)C3(N)\psi(x)\in C^{3}(\partial N). Since N\partial N is compact, by Lemma 6.1 we extend ψ(x)\psi(x) into a C2C^{2} function on NN such that

(6.9) ψ(x)=ψ(y) for any xN\Nr with d(x,y)=d(x,N)\psi(x)=\psi(y)\text{ for any $x\in N\backslash N_{r}$ with $d(x,y)=d(x,\partial N)$}

Because NrN_{r} for each r<r0r<r_{0} is ϕ\phi-mean convex, by Theorem 5.3 there is a ur(x)u_{r}(x) in C2(Ω)C(Ω¯)C^{2}(\Omega)\cap C(\bar{\Omega}) to solve the Dirichlet problem

(6.10) {Lu:=div(Duω)+nDlogϕ(x),Duω=0, on Nru(x)=ψ(x) on Nr\left\{\begin{split}Lu:=-div(\frac{Du}{\omega})&+n\langle D\log\phi(x),-\frac{Du}{\omega}\rangle=0,\text{ on }N_{r}\\ &u(x)=\psi(x)\quad\text{ on }\partial N_{r}\end{split}\right.

Let μ\mu be a positive constant such that maxxNN|ψ(x)|μ\max_{x\in N\cup\partial N}|\psi(x)|\leq\mu. By the maximum principle, for each r(0,r0)r\in(0,r_{0}),

(6.11) maxNr|ur|μ\max_{N_{r}}|u_{r}|\leq\mu

For any fixed embedded open ball Bε(x)B_{\varepsilon}(x) with dist(x,N)>2εdist(x,\partial N)>2\varepsilon, Theorem A.3 implies that

(6.12) maxB¯ε(x){|ur|(x),|Dur|(x),|D2ur|(x)}μ1\max_{\bar{B}_{\varepsilon}(x)}\{|u_{r}|(x),|Du_{r}|(x),|D^{2}u_{r}|(x)\}\leq\mu_{1}

where μ1\mu_{1} is a constant depending on ψ\psi, ϕ\phi on B32ε(x)B_{\frac{3}{2}\varepsilon}(x) and μ\mu. Thus after choosing a subsequence from {ur(x)}r<r0\{u_{r}(x)\}_{r<r_{0}}, denoted by uk(x)u_{k}(x), uk(x)u_{k}(x) converges uniformly to u(x)u_{\infty}(x) in the C2C^{2} sense on any compact set of NN disjoint with N\partial N. Thus

(6.13) Lu(x)=0 on NLu_{\infty}(x)=0\quad\text{ on }N

Next we show limxz,xNu(x)=ψ(z)\lim_{x\rightarrow z,x\in N}u_{\infty}(x)=\psi(z) for any zΩz\in\partial\Omega. To achieve this goal we construct the supersolution and subsolution of Lu=0Lu=0 in (6.10) on NrN_{r} for some r(0,r0)r\in(0,r_{0}).

Lemma 6.6.

Let ψ(x)\psi(x) satisfy (6.9), let μ\mu be a constant given in (6.11) and r0r_{0} be given in Definition 6.1. Suppose for any r(0,r0)r\in(0,r_{0}) NrN_{r} is ϕ\phi-mean convex. Then there are three positive constants r1(0,r0)r_{1}\in(0,r_{0}) and ν>0\nu>0 and κ>0\kappa>0 such that

(6.14) u±(x)=ψ(x)±φ(d(x)) on N\Nr1 φ(r):=1νlog(1+κr)\displaystyle u_{\pm}(x)=\psi(x)\pm\varphi(d(x))\text{ on $N\backslash N_{r_{1}}$\quad}\varphi(r):=\frac{1}{\nu}\log(1+\kappa r)
(6.15) Lu+(x)0,Lu(x)0 on N\Nr1\displaystyle Lu_{+}(x)\geq 0,\quad Lu_{-}(x)\leq 0\text{ on $N\backslash N_{r_{1}}$}
(6.16) u+(x)2μ,u(x)2μ on {xN:d(x,N)=r1}\displaystyle u_{+}(x)\geq 2\mu,\quad u_{-}(x)\leq-2\mu\text{ on $\{x\in N:d(x,\partial N)=r_{1}\}$}
Proof.

Let {x1,,xn}\{x_{1},\cdots,x_{n}\} be a local coordinate in NN and {i}i=1\{\partial_{i}\}_{i=1}^{\infty} be the corresponding frame. We write i,j\langle\partial_{i},\partial_{j}\rangle as σij\sigma_{ij} and (σij)=(σij)1(\sigma^{ij})=(\sigma_{ij})^{-1}. For any smooth function ff, fif_{i}, fijf_{ij} denote the first and second covariant derivatives of ff and fi=σikfkf^{i}=\sigma^{ik}f_{k} with the sum over kk.
Note that

(6.17) u:=ω3Lu=(1+|Du|2){Δu+nDlogϕ(x),Du}+uiujuij\mathcal{E}u:=\omega^{3}Lu=(1+|Du|^{2})\{-\Delta u+n\langle D\log\phi(x),-Du\rangle\}+u^{i}u^{j}u_{ij}

Here ω=1+|Du|2\omega=\sqrt{1+|Du|^{2}}, Δu=div(Du)\Delta u=div(Du) and DD denote the Laplacian and the covariant derivative of NN respectively. Now assume u±(x)u_{\pm}(x) are given in (6.14) where ν,κ\nu,\kappa are determined later.
By (6.2) and (6.9) on N\Nr0N\backslash N_{r_{0}} we have Dψ,Dd=0\langle D\psi,Dd\rangle=0 and

(6.18) Dlogϕ(x),Du+(x)=φDlogϕ(x),Dd\langle D\log\phi(x),-Du_{+}(x)\rangle=\varphi^{\prime}\langle D\log\phi(x),-Dd\rangle\quad

where d(x)=d(x,N)d(x)=d(x,\partial N) and φ=1ν(1+κd(x))\varphi^{\prime}=\frac{1}{\nu(1+\kappa d(x))}. With (6.18) we compute u+\mathcal{E}u_{+} on N\Nr0N\backslash N_{r_{0}} as follows.

(u+)\displaystyle\mathcal{E}(u_{+}) =φ(1+|Du+|2)(Δd+nDlogϕ,Dd)(1+|Dψ|2)Δψ\displaystyle=\varphi^{\prime}(1+|Du_{+}|^{2})(-\Delta d+n\langle D\log\phi,-Dd\rangle)-(1+|D\psi|^{2})\Delta\psi
+ψiψjψij+(φ)2(Δψ+didjψij)(1+|Dψ|2)φ′′\displaystyle+\psi^{i}\psi^{j}\psi_{ij}+(\varphi^{\prime})^{2}(-\Delta\psi+d^{i}d^{j}\psi_{ij})-(1+|D\psi|^{2})\varphi^{\prime\prime}

Since NrN_{r} is ϕ\phi-mean convex for each r(0,r0)r\in(0,r_{0}) and φ>0\varphi^{\prime}>0, the first term above is nonnegative. By (6.14) φ′′=νφ2\varphi^{\prime\prime}=-\nu\varphi^{\prime 2}. Now assuming φ1\varphi^{\prime}\geq 1 (u+)\mathcal{E}(u_{+}) on N\Nr0N\backslash N_{r_{0}} satisfies

(6.19) u+νφ2C(φ2)\mathcal{E}u_{+}\geq\nu\varphi^{\prime 2}-C(\varphi^{\prime 2})

for some positive constant CC only depending on ψ(x)\psi(x). Now take ν=C\nu=C we obtain u+0\mathcal{E}u_{+}\geq 0 on N\Nr0N\backslash N_{r_{0}}. This shows (6.15) for u+u_{+} and we have to determine κ\kappa and r1(0,r0)r_{1}\in(0,r_{0}) in a such way φ(r)1\varphi^{\prime}(r)\geq 1 and φ(r1)μ1:=2μ+maxN¯ψ(x)\varphi(r_{1})\geq\mu_{1}:=2\mu+\max_{\bar{N}}\psi(x). We have

(6.20) φ(r)=1νκ1+κr>1νκ1+κr11\displaystyle\varphi^{\prime}(r)=\frac{1}{\nu}\frac{\kappa}{1+\kappa r}>\frac{1}{\nu}\frac{\kappa}{1+\kappa r_{1}}\geq 1
(6.21) φ(r1)=1νlog(1+κr1)μ1\displaystyle\varphi(r_{1})=\frac{1}{\nu}\log(1+\kappa r_{1})\geq\mu_{1}

All conditions are satisfied provided we take r1r_{1} small enough and large κ\kappa satisfying

(6.22) κmax{ν1r1ν,eμ1νr1}\kappa\geq\max\{\frac{\nu}{1-r_{1}\nu},\frac{e^{\mu_{1}\nu}}{r_{1}}\}

This gives (6.16) for u+(x)u_{+}(x).
A similar derivation yields the conclusions for u(x)u_{-}(x). The proof is complete. ∎

Now we continue to show Theorem 6.4. Now for any r(0,r1)r\in(0,r_{1}), with the maximum principle, (6.10) and (6.11) Lemma 6.6 implies that u(x)ur(x)u+(x)u_{-}(x)\leq u_{r}(x)\leq u_{+}(x) on Nr\Nr0N_{r}\backslash N_{r_{0}}. Let rr go to 0 we obtain that

(6.23) u(x)u(x)u+(x) on N\Nr1.u_{-}(x)\leq u_{\infty}(x)\leq u_{+}(x)\quad\text{ on $N\backslash N_{r_{1}}$}.

Since both u+(x)u_{+}(x) and u(x)u_{-}(x) are continuous on NNN\cup\partial N and equal to ψ(x)\psi(x) on N\partial N, thus

(6.24) limxN,xzNu(x)=ψ(z)\lim_{x\in N,x\rightarrow z\in\partial N}u_{\infty}(x)=\psi(z)

Let UU be the subgraph of u(x)u_{\infty}(x) in MϕM_{\phi}. Let T=[[U]]T=\partial[[U]] be the corresponding integer multiplicity current. Thus with respect to the product topology of N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R}, T=Γ\partial T=\Gamma.
Now we obtain the existence of a local integer multiplicity current in MϕM_{\phi} with the desirable infinity boundary Γ\Gamma for any ψC3(N)\psi\in C^{3}(\partial N). Moreover it is a minimal graph over NN in MϕM_{\phi}.
Step Three: General Case As for ψ(x)C(N)\psi(x)\in C(\partial N), we can construct two monotone sequences {ψk1(x)}k=1\{\psi^{1}_{k}(x)\}_{k=1}^{\infty} and {ψk2(x)}k=1\{\psi^{2}_{k}(x)\}_{k=1}^{\infty} in C3(N)C^{3}(\partial N) such that the former one converges increasingly to ψ(x)\psi(x) and the latter one converges decreasingly to ψ(x)\psi(x) in the C0(N)C^{0}(\partial N) sense. Let {uki(x)}k=1\{u^{i}_{k}(x)\}_{k=1}^{\infty} be the solutions of Lu=0Lu=0 on NN with asymptotic value {ψki(x)}k=1\{\psi^{i}_{k}(x)\}_{k=1}^{\infty} for i=1,2i=1,2. From the maximum principle and the translating invariant of minimal graphs, we have {uk1(x)}\{u^{1}_{k}(x)\} is an increasing sequence on NN and uk1(x)ul2(x)u^{1}_{k}(x)\leq u^{2}_{l}(x) for any k,lk,l. Note that both sequences are uniformly bounded. By the interior estimate of Lu=0Lu=0 in NN, uk(x)u_{k}(x) locally converges to a C2C^{2} function u(x)u_{\infty}(x) in NN satisfying Lu=0Lu_{\infty}=0. Thus for any k,lk,l, on NN we obtain

(6.25) uk1(x)u(x)ul2(x)u^{1}_{k}(x)\leq u_{\infty}(x)\leq u^{2}_{l}(x)

Thus (6.24) still holds for ψ(x)C(N)\psi(x)\in C(\partial N). As a result for any ψ(x)C(N)\psi(x)\in C(\partial N) we show there is a minimal graph Σ=(x,u(x))\Sigma=(x,u(x)) in MϕM_{\phi} with the infinity boundary Γ\Gamma. It is also a local area minimizing integer multiplicity current according to Theorem 4.1.
Step Four: Uniqueness. Suppose TT is a local area minimizing nn-integer multiplicity current with infinity boundary Γ=(x,ψ(x))\Gamma=(x,\psi(x)). For any tt\in\mathbb{R}, we define ft(x,t)=(x,r+t)f_{t}(x,t)=(x,r+t) for any xNx\in N and tt\in\mathbb{R}. By (1.1), ftf_{t} is an isometry of MϕM_{\phi}. Thus ft,#Tf_{t,\#}T is also a local area minimizing nn-integer multiplicity current in MϕM_{\phi} with infinity boundary Γt=(x,ψ(x)+t)\Gamma_{t}=(x,\psi(x)+t). Recall that spt(T)N×[a,b]spt(T)\subset N{\mkern-1.0mu\times\mkern-1.0mu}[a,b]. We claim that ft,#(T)T=f_{t,\#}(T)\cap T=\emptyset for any t0t\neq 0. Otherwise there is a t0t\neq 0 such that the regular part of ft,#(T)f_{t,\#}(T) and TT intersects transversely. Since ΓtΓ=\Gamma_{t}\cap\Gamma=\emptyset for any t0t\neq 0. By the area minimzing property, we can replace a piece of TT by the piece of ft,#f_{t,\#} that is cut off by TT and still have a local area minimizing property. But then the singular set of the resulting integer multiplicity current would include the intersection of TT and ft,#(T)f_{t,\#}(T) with n1n-1 dimension. This contradicts to the n7n-7 dimensional singular set of local area minimizing integer multiplicity currents. Thus TT is a graph over NN. Arguing similar as in the proof of the property (3) in Theorem 4.1, TT is a C2C^{2} minimal graph in MϕM_{\phi} with the infinity boundary Γ\Gamma. Since NNN\cap\partial N is compact, applying the maximum principle into Lu=0Lu=0 yields that the uniqueness of the minimal graph.
The proof is complete. ∎

Appendix A Interior estimate of mean curvature equations

Throughout this section let NN be a complete Riemannian manifold with a metric σ\sigma and f(x)f(x) be a C3C^{3} function in NN. Let M0M_{0} be the product manifold N×N{\mkern-1.0mu\times\mkern-1.0mu}\mathbb{R} equipped with the metric σ+dr2\sigma+dr^{2}. Let DD be the covariant derivative of NN. Let SS be an orientable C2C^{2} hypersurface in M0M_{0} with the normal vector v\vec{v} for any β1(0,1)\beta_{1}\in(0,1). We have the following theorem about its angle function v,r\langle\vec{v},\partial_{r}\rangle. Comparing to [8, section 2.3] we use a moving frame method from [11, section 2] to obtain the interior estimate of mean curvature equations.

Lemma A.1.

Let Θ=v,r\Theta=\langle\vec{v},\partial_{r}\rangle. Suppose the mean curvature of SS with respect v\vec{v} satisfies that

(A.1) HS+Df,v=0H_{S}+\langle Df,\vec{v}\rangle=0

where DD is the gradient of ff on NN. Then in the Lipschitz sense it holds that

(A.2) ΔΘ+(|A|2+R¯ic(v,v)Hess(f)(v,v)Θ+Θ,Df=0\begin{split}\Delta\Theta&+(|A|^{2}+\bar{R}ic(\vec{v},\vec{v})-Hess(f)(\vec{v},\vec{v}\rangle)\Theta\\ &+\langle\nabla\Theta,Df\rangle=0\end{split}

where |A|2|A|^{2} is the second fundamental form of SS, \nabla is the covariant derivative of SS and R¯ic\bar{R}ic is the Ricci curvature of the product manifold M0M_{0}, HessHess is the Hessian of ff in M0M_{0}.

Proof.

Notice that SS is a C3C^{3} hypersurface. We denote {ei}i=1n\{e_{i}\}_{i=1}^{n} by a local orthonormal frame of SS. Note that the derivation in Lemma 2.2, [41] is true for any dimension. That is by Lemma 2.2 in [41] we have

(A.3) ΔΘ+(|A|2+R¯ic(v,v))ΘH,r=0\Delta\Theta+(|A|^{2}+\bar{R}ic(\vec{v},\vec{v}))\Theta-\langle\nabla H,\partial_{r}\rangle=0

where HH is the mean curvature of SS. Let D¯\bar{D} is the corvariant derivative of M0M_{0}. Since H=Df,v-H=\langle Df,\vec{v}\rangle,

(A.4) H,r=ei,rD¯eiDf,v+ei,rDf,ekhik-\langle\nabla H,\partial_{r}\rangle=\langle e_{i},\partial_{r}\rangle\langle\bar{D}_{e_{i}}Df,\vec{v}\rangle+\langle e_{i},\partial_{r}\rangle\langle Df,e_{k}\rangle h_{ik}

where D¯eiv=hikek\bar{D}_{e_{i}}\vec{v}=h_{ik}e_{k}. Note that r,eiei=rv,rv\langle\partial_{r},e_{i}\rangle e_{i}=\partial_{r}-\langle\vec{v},\partial_{r}\rangle\vec{v}. Thus

(A.5) ei,rD¯eiDf,v=ΘHess(f)v,v\langle e_{i},\partial_{r}\rangle\langle\bar{D}_{e_{i}}Df,\vec{v}\rangle=-\Theta Hess(f)\langle\vec{v},\vec{v}\rangle

Here we use the fact D¯rDf=0\bar{D}_{\partial_{r}}Df=0 . Since the metric of M0M_{0} is the product metric, D¯Xr=0\bar{D}_{X}\partial_{r}=0 for any tangent vector field XX. Thus Θ,Df=Df,ekhikei,r\langle\nabla\Theta,Df\rangle=\langle Df,e_{k}\rangle h_{ik}\langle e_{i},\partial_{r}\rangle. Combining this with (A.4), (A.5) we obtain

(A.6) H,r=ΘHess(f)v,v+Θ,Df-\langle\nabla H,\partial_{r}\rangle=-\Theta Hess(f)\langle\vec{v},\vec{v}\rangle+\langle\nabla\Theta,Df\rangle

Putting this into (A.3) we obtain the conclusion in the case fC3(N)f\in C^{3}(N).
In the case of fC2(N)f\in C^{2}(N), SS is a C2,βC^{2,\beta} hypersurface for any β\beta in (0,1)(0,1). The result comes from the classical approximating method. The proof is complete. ∎

As a corollary we obtain the following Harnack type result.

Corollary A.2.

Let Ω\Omega be a bounded domain in NN and ff be a C3C^{3} function in NN. Suppose SS is a C2C^{2} connected orientable hypersurface satisfying (A.1) with its normal vector v\vec{v}. If Θ=v,r0\Theta=\langle\vec{v},\partial_{r}\rangle\geq 0 on SS, Θ0\Theta\equiv 0 or Θ>0\Theta>0 on the whole SS.

Proof.

Suppose ff is C3C^{3}. Then SS is C3C^{3}. Because Ω\Omega is bounded, there is a positive constant C>0C>0 such that

(A.7) ΔΘCΘ+Θ,Df0\Delta\Theta-C\Theta+\langle\nabla\Theta,Df\rangle\leq 0

Since Θ0\Theta\geq 0 on SS, the weak maximum principle implies that Θ0\Theta\equiv 0 or Θ>0\Theta>0 on the whole SS.
When ff is C2C^{2}, SS is C2,βC^{2,\beta}. Then Θ\Theta satisfies (A.7) in the Lipschitz sense. By the weak Harnack inequality in [38], Θ>0\Theta>0 or Θ0\Theta\equiv 0 on the whole SS. The proof is complete. ∎

We have the following interior estimate of mean curvature equations.

Theorem A.3.

Let fC3(N)f\in C^{3}(N). Fix x0Nx_{0}\in N and let Bρ(x0)B_{\rho}(x_{0}) be an embedded ball centered at x0x_{0} with radius ρ\rho. Suppose u(x)C2(Bρ(x0))u(x)\in C^{2}(B_{\rho}(x_{0})) satisfying

(A.8) div(Duω)+Df,Duω=0-div(\frac{Du}{\omega})+\langle Df,-\frac{Du}{\omega}\rangle=0

where DuDu is the gradient of uu, ω=1+|Du|2\omega=\sqrt{1+|Du|^{2}} and divdiv is the divergence of NN. If maxxBρ(x0)|u(x)|c0\max_{x\in B_{\rho}(x_{0})}|u(x)|\leq c_{0} for some positive constant c0c_{0}, then

(A.9) maxBρ2(x0)|Du|C\max_{B_{\frac{\rho}{2}}(x_{0})}|Du|\leq C

where CC is a constant only depending on the Ricci curvature, the C2C^{2} norm of ff, c0c_{0} and ρ\rho.

Proof.

Our proof follows from the idea of Lemma 2.1 in [11] by Eichmair.
First we assume f(x)f(x) is a C3C^{3} function in NN. By the classical Schauder estimate (x)(x) is C3,βC^{3,\beta} for any β(0,1)\beta\in(0,1). We denote by Λ\Lambda the C2C^{2} norm of ff on Ω¯\bar{\Omega}. i.e.

(A.10) Λ:=maxxΩ¯{|f(x)|,|Df(x)|,|D2f(x)|}\Lambda:=\max_{x\in\bar{\Omega}}\{|f(x)|,|Df(x)|,|D^{2}f(x)|\}

Let Σ\Sigma be the graph of u(x)u(x) with its upward normal vector v=rDuω\vec{v}=\frac{\partial_{r}-Du}{\omega} in M0M_{0}. By (A.8), the mean curvature of Σ\Sigma with respect to v\vec{v}, HΣH_{\Sigma}, in M0M_{0} satisfies that HΣ+Df,v=0H_{\Sigma}+\langle Df,\vec{v}\rangle=0. Let Θ=v,r=1ω\Theta=\langle\vec{v},\partial_{r}\rangle=\frac{1}{\omega} which is a C2,βC^{2,\beta} function. By Lemma A.1, we have

(A.11) ΔΘ+(|A|2+R¯ic(v,v)Hess(f)(v,v)ΘΘ,Df=0\begin{split}\Delta\Theta&+(|A|^{2}+\bar{R}ic(\vec{v},\vec{v})-Hess(f)(\vec{v},\vec{v}\rangle)\Theta\\ &-\langle\nabla\Theta,Df\rangle=0\end{split}

Thus ω\omega satisfies that

(A.12) Δωω2|ω|2ω2c1+ωω,Df\frac{\Delta\omega}{\omega}\geq 2\frac{|\nabla\omega|^{2}}{\omega^{2}}-c_{1}+\langle\frac{\nabla\omega}{\omega},Df\rangle

where c10c_{1}\geq 0 is a constant only depending on the Ricci curvature and Λ\Lambda.
Let d(x,x0)d(x,x_{0}) be the distance function between x0x_{0} and xx. Now we define

(A.13) q(x):=1+u(x)2c032ρ2d2(x,x0)q(x):=1+\frac{u(x)}{2c_{0}}-\frac{3}{2\rho^{2}}d^{2}(x,x_{0})

Define B:={xN:q(x)>0}B:=\{x\in N:q(x)>0\}. Thus Bρ2(x0)BBρ(x0)B_{\frac{\rho}{2}}(x_{0})\subset B\subset B_{\rho}(x_{0}). Set η(x):=eKq(x)1\eta(x):=e^{Kq(x)}-1 where KK is a positive constant determined later. Thus the maximum of ηω\eta\omega is obtained at a point in BB, for example x1Bρ(x0)x_{1}\in B_{\rho}(x_{0}). At this point, ηω+ωη=0\eta\nabla\omega+\omega\nabla\eta=0 and

0\displaystyle 0 1ωΔ(ηω)=ηωΔω+2η,ωω+Δη\displaystyle\geq\frac{1}{\omega}\Delta(\eta\omega)=\frac{\eta}{\omega}\Delta\omega+2\frac{\langle\nabla\eta,\nabla\omega\rangle}{\omega}+\Delta\eta
c1η+ηωω,Df+Δη by (A.12)\displaystyle\geq-c_{1}\eta+\eta\langle\frac{\nabla\omega}{\omega},Df\rangle+\Delta\eta\text{\,by\,}\eqref{eq:estimate:omega}
(A.14) c1c1eKq(x)+η,Df+Δη\displaystyle\geq c_{1}-c_{1}e^{Kq(x)}+\langle\nabla\eta,Df\rangle+\Delta\eta

Observe that

(A.15) η,DfeKq(x)(K2|q(x)|224Λ2)\langle\nabla\eta,Df\rangle\geq e^{Kq(x)}(-\frac{K^{2}|\nabla q(x)|^{2}}{2}-4\Lambda^{2})

Since Δη=KΔq(x)+K2|q(x)|2\Delta\eta=K\Delta q(x)+K^{2}|\nabla q(x)|^{2}, from (A.14) and (A.15) we obtain

(A.16) 0\displaystyle 0 eKq(x)(KΔq(x)+12K2|q(x)|24Λ2c1)\displaystyle\geq e^{Kq(x)}(K\Delta q(x)+\frac{1}{2}K^{2}|\nabla q(x)|^{2}-4\Lambda^{2}-c_{1})

Note that for any C2C^{2} function h(x)h(x), Δh=Hess(h)(ei,ei)HΣDh,v)\Delta h=Hess(h)(e_{i},e_{i})-H_{\Sigma}\langle Dh,\vec{v}\rangle) where {ei}i=1n\{e_{i}\}_{i=1}^{n} is the orthonormal frame on SS. Thus

Δu=1ωDf,Duω,\displaystyle\Delta u=\frac{1}{\omega}\langle Df,-\frac{Du}{\omega}\rangle,
Δd2(x,x0)=Hess(d2)(ei,ei)+Df,DuωDd2(x,x0),Duω\displaystyle\Delta d^{2}(x,x_{0})=Hess(d^{2})(e_{i},e_{i})+\langle Df,-\frac{Du}{\omega}\rangle\langle Dd^{2}(x,x_{0}),-\frac{Du}{\omega}\rangle

As a result,

(A.17) Δq(x)c2\Delta q(x)\geq c_{2}

where c2c_{2} is a constant depending on Λ\Lambda, the Ricci curvature on Bρ(x0)B_{\rho}(x_{0}). Let DD be the covariant derivative on MM. Note that

(A.18) |q(x)|214c0|Du|2c3|\nabla q(x)|^{2}\geq\frac{1}{4c_{0}}|Du|^{2}-c_{3}

Here c3c_{3} is a fixed constant depending on ρ\rho. Now combining (A.16) with (A.17), (A.18) together we obtain

(A.19) 0c3K+14c0|Du|2c21K2(4Λ2c1)0\geq\frac{c_{3}}{K}+\frac{1}{4c_{0}}|Du|^{2}-c_{2}-\frac{1}{K^{2}}(4\Lambda^{2}-c_{1})

When taking KK sufficiently large, we obtain at x1x_{1}, |Du|212c3|Du|^{2}\leq\frac{1}{2}c_{3}. Note that KK only depends on Λ\Lambda, c1c_{1} and c2c_{2}, the Ricci curvature of MM on Bρ(x0)B_{\rho}(x_{0}). Thus

(A.20) (eKq(x)1)ωe32K(1+c32)(e^{Kq(x)}-1)\omega\leq e^{\frac{3}{2}K}(1+\frac{c_{3}}{2})

on Bρ(x0)B_{\rho}(x_{0}). Note that for any xBρ2(x0)x\in B_{\frac{\rho}{2}}(x_{0}), by (A.13) q(x)18q(x)\geq\frac{1}{8}. Thus maxBρ2(x0)|Du|C\max_{B_{\frac{\rho}{2}}(x_{0})}|Du|\leq C. Here CC is a constant only depending on c0c_{0},Λ\Lambda and ρ\rho.
The proof is complete.

References

  • [1] A. Aiolfi, J. Ripoll, and M. Soret. The Dirichlet problem for the minimal hypersurface equation on arbitrary domains of a Riemannian manifold. Manuscripta Math., 149(1-2):71–81, 2016.
  • [2] F. Almgren. Review: Enrico Giusti, Minimal surfaces and functions of bounded variation. Bull. Amer. Math. Soc. (N.S.), 16(1):167–171, 01 1987.
  • [3] Y. N. Alvarez. PDE and hypersurfaces with prescribed mean curvature, Nov. 2019. arXiv:1911.12679.
  • [4] Y. N. Alvarez and R. Sa Earp. Sharp solvability criteria for Dirichlet problems of mean curvature type in Riemannian manifolds: non-existence results. Calc. Var. Partial Differential Equations, 58(6):Paper No. 197, 12, 2019.
  • [5] M. T. Anderson. Complete minimal varieties in hyperbolic space. Invent. Math., 69(3):477–494, 1982.
  • [6] M. T. Anderson. Complete minimal hypersurfaces in hyperbolic nn-manifolds. Comment. Math. Helv., 58(2):264–290, 1983.
  • [7] J. Bemelmans and U. Dierkes. On a singular variational integral with linear growth. I. Existence and regularity of minimizers. Arch. Rational Mech. Anal., 100(1):83–103, 1987.
  • [8] J.-B. Casteras, E. Heinonen, and I. Holopainen. Dirichlet problem for FF-minimal graphs. J. Anal. Math., 138(2):917–950, 2019.
  • [9] D. De Silva and J. Spruck. Rearrangements and radial graphs of constant mean curvature in hyperbolic space. Calc. Var. Partial Differential Equations, 34(1):73–95, 2009.
  • [10] F. Duzaar and K. Steffen. λ\lambda minimizing currents. Manuscripta Math., 80(4):403–447, 1993.
  • [11] M. Eichmair. The Plateau problem for marginally outer trapped surfaces. J. Differential Geom., 83(3):551–583, 2009.
  • [12] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
  • [13] H. Federer and W. H. Fleming. Normal and integral currents. Ann. of Math. (2), 72:458–520, 1960.
  • [14] Q. Gao and H. Zhou. The area minimizing problem in conformal cones, 2020. arXiv:2001.06207.
  • [15] C. Gerhardt. Closed hypersurfaces of prescribed mean curvature in locally conformally flat Riemannian manifolds. J. Differential Geom., 48(3):587–613, 1998.
  • [16] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [17] E. Giusti. Generalized solutions for the mean curvature equation. Pacific J. Math., 88(2):297–321, 1980.
  • [18] E. Giusti. Minimal surfaces and functions of bounded variation, volume 80 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 1984.
  • [19] B. Guan and J. Spruck. Hypersurfaces of constant mean curvature in hyperbolic space with prescribed asymptotic boundary at infinity. Amer. J. Math., 122(5):1039–1060, 2000.
  • [20] E. M. Guio and R. Sa Earp. Existence and non-existence for a mean curvature equation in hyperbolic space. Commun. Pure Appl. Anal., 4(3):549–568, 2005.
  • [21] E. M. Guio and R. Sa Earp. Errata: “Existence and non-existence for a mean curvature equation in hyperbolic space” [Commun. Pure Appl. Anal. 4 (2005), no. 3, 549–568; mr2167187]. Commun. Pure Appl. Anal., 7(2):465, 2008.
  • [22] R. Hardt and F.-H. Lin. Regularity at infinity for area-minimizing hypersurfaces in hyperbolic space. Invent. Math., 88(1):217–224, 1987.
  • [23] J. M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.
  • [24] F. Lin and X. Yang. Geometric measure theory—an introduction, volume 1 of Advanced Mathematics (Beijing/Boston). Science Press Beijing, Beijing; International Press, Boston, MA, 2002.
  • [25] F.-H. Lin. Regularity for a class of parametric obstacle problems integrand integral current prescribed mean curvature minimal surface system. ProQuest LLC, Ann Arbor, MI, 1985. Thesis (Ph.D.)–University of Minnesota.
  • [26] F.-H. Lin. On the Dirichlet problem for minimal graphs in hyperbolic space. Invent. Math., 96(3):593–612, 1989.
  • [27] L. Lin and L. Xiao. Modified mean curvature flow of star-shaped hypersurfaces in hyperbolic space. Comm. Anal. Geom., 20(5):1061–1096, 2012.
  • [28] M. McGonagle and L. Xiao. Minimal Graphs and Graphical Mean Curvature Flow in Mn×M^{n}\times\mathbb{R}. J. Geom. Anal., 30(2):2189–2224, 2020.
  • [29] M. Miranda. Superfici cartesiane generalizzate ed insiemi di perimetro localmente finito sui prodotti cartesiani. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 18:515–542, 1964.
  • [30] B. Nelli and J. Spruck. On the existence and uniqueness of constant mean curvature hypersurfaces in hyperbolic space. In Geometric analysis and the calculus of variations, pages 253–266. Int. Press, Cambridge, MA, 1996.
  • [31] J. Ripoll and M. Telichevesky. On the asymptotic plateau problem for CMC hypersurfaces in hyperbolic space. Bull. Braz. Math. Soc. (N.S.), 50(2):575–585, 2019.
  • [32] J. Serrin. The problem of Dirichlet for quasilinear elliptic differential equations with many independent variables. Philos. Trans. Roy. Soc. London Ser. A, 264:413–496, 1969.
  • [33] L. Simon. On a theorem of de Giorgi and Stampacchia. Math. Z., 155(2):199–204, 1977.
  • [34] L. Simon. Lectures on geometric measure theory, volume 3 of Proceedings of the Centre for Mathematical Analysis, Australian National University. Australian National University, Centre for Mathematical Analysis, Canberra, 1983.
  • [35] L. Simon. The minimal surface equation. In Geometry, V, volume 90 of Encyclopaedia Math. Sci., pages 239–272. Springer, Berlin, 1997.
  • [36] J. Spruck. Interior gradient estimates and existence theorems for constant mean curvature graphs in Mn×M^{n}\times\mathbb{R}. Pure Appl. Math. Q., 3(3, Special Issue: In honor of Leon Simon. Part 2):785–800, 2007.
  • [37] I. Tamanini. Boundaries of Caccioppoli sets with Hölder-continuous normal vector. J. Reine Angew. Math., 334:27–39, 1982.
  • [38] N. S. Trudinger. On Harnack type inequalities and their application to quasilinear elliptic equations. Comm. Pure Appl. Math., 20:721–747, 1967.
  • [39] H. Zhou. Inverse mean curvature flows in warped product manifolds. J. Geom. Anal., 28(2):1749–1772, 2018.
  • [40] H. Zhou. The boundary behavior of domains with complete translating, minimal and CMC graphs in N2×N^{2}\times\mathbb{R}. Sci. China Math., 62(3):585–596, 2019.
  • [41] H. Zhou. Generalized solutions to the Dirichlet problem of translating mean curvature equations., 2019. arXiv:1902.01512.