1 Introduction

Consider the following weighted least gradient problem

$$\begin{aligned} \min \left\{ \int _{\Omega } a|Du|: u\in BV(\Omega ), u|_{\partial \Omega }=f \right\} , \end{aligned}$$
(1)

where \(\Omega \) is a bounded open set in \(\mathbb {R}^n\) (\(n\ge 2\)) with connected Lipschitz boundary \(\partial \Omega \), a is a bounded non-negative function, and \(f\in C(\partial \Omega )\). This problem was first studied for the case \(a\equiv 1\) in [17, 19], where existence of a unique minimizer was proved under the assumption that f is continuous and \(\partial \Omega \) has positive mean curvature on a dense subset of \(\partial \Omega \) (see the precise hypotheses (3.1) and (3.2) in [17]). As in [6], these assumptions on \(\partial \Omega \) are needed in the existence proofs but not for the uniqueness arguments. In [6] it has recently been proved that, if \(a\in C^{1,1}(\Omega )\) is positive and bounded away from zero and \(f\in C(\partial \Omega )\), then the weighted least gradient problem (1) has at most one minimizer in \(BV(\Omega )\). The counterexample to uniqueness in [6] for \(a \in C^{1,\alpha }(\Omega )\), \(0<\alpha <1\), shows that the regularity \(a\in C^{1,1}(\Omega )\) is sharp in general. The assumption \(a>0\) in \(\Omega \) is also crucial for the results in [6].

In contrast with the results in [6, 17], here we present uniqueness results that take into account the regularity of an existing minimizer. The main contribution of this paper is to show that if existence is known a priori, then the assumptions on the weight function a can be considerably weakened. The uniqueness result in this paper is based on a calibration argument and takes into account a priori existence of a \(C^1\) minimizer. The following theorem is a consequence of a more general result (Theorem 1.2 below) that will be proved in Sect. 3.

Theorem 1.1

Let \(\Omega \subset \mathbb {R}^n\) be a bounded Lipschitz domain with connected boundary, \(f\in C^1(\partial \Omega )\), and \(a\in C(\overline{\Omega })\). If \(a>0\) in \(\overline{\Omega }\) and (7) has a minimizer \(u\in C^1(\overline{\Omega })\) with \(|\nabla u|>0\) in \(\overline{\Omega }\), then u is the unique minimizer of (1) in \(BV(\Omega )\).

Our motivation comes from a hybrid inverse problem in medical imaging, which concerns determining the conductivity of a body from knowledge of the magnitude \(a=|J|\) (in \(\Omega \)) of one current density vector field J generated by imposing the voltage f on \(\partial \Omega \), see [12]. The interior data |J| can be obtained non-invasively via a magnetic resonance technique pioneered in [7]. In [12] this problem was reduced to the weighted least gradient problem (1), by showing that the voltage potential is a minimizer. More precisely, assume \(\Omega \subset \mathbb {R}^n\) is made of conductive materials with conductivity \(\sigma \). If the voltage f is imposed on \(\partial \Omega \), then the corresponding voltage potential u is the solution of the following conductivity equation

$$\begin{aligned} \left\{ \begin{array}{ll} \nabla \cdot \sigma \nabla u=0,&{}\quad \text{ in }\,\Omega ,\\ u=f. &{}\quad \text{ on }\,\partial \Omega \end{array} \right. \end{aligned}$$
(2)

Let \(J=-\sigma \nabla u\) be the current density generated by imposing the voltage f on \(\partial \Omega \). Then the voltage potential u is a minimizer of the the weighted least gradient problem

$$\begin{aligned} \min \left\{ \int _\Omega |J|\;|\nabla u|dx:\;u\in BV(\Omega ),\, u|_{\partial \Omega }=f\right\} , \end{aligned}$$

(see Proposition 1.2 in [12] and the density argument of Proposition 3 in [15]). More generally, if \(\Omega \) also contains perfectly conducting and insulating inclusions \({\mathcal {U_P}}\) and \({\mathcal {U_I}}\). Then the corresponding voltage potential u is the unique solution of the following equation

$$\begin{aligned} \left\{ \begin{array}{ll} \nabla \cdot \sigma \nabla u=0,&{}\quad \text{ in }\,\Omega {\setminus }\overline{{\mathcal {U_P}} \cup {\mathcal {U_I}}},\\ \nabla u=0, &{}\quad \text{ in } \ \ {\mathcal {U_P}},\\ u|_+=u|_-,&{}\quad \text{ on }\ \ \partial ( {\mathcal {U_P}} \cup {\mathcal {U_I}}),\\ \int _{\partial {\mathcal {U_P}}^j}\sigma \frac{\partial u}{\partial \nu }|_{+}ds=0,&{}\quad j=1,2,\ldots ,\\ \frac{\partial u}{\partial \nu }|_{+}=0,&{}\quad \text{ on }\; {\mathcal {U_I}},\\ u|_{\partial \Omega }=f, \end{array} \right. \end{aligned}$$
(3)

where \( {\mathcal {U_P}} \cap {\mathcal {U_I}}=\emptyset \) and \( {\mathcal {U_P}}=\bigcup _{j=1} {\mathcal {U_P}}^{j}\) is the partition of \( {\mathcal {U_P}}\) into open connected components (see the appendix in [10] for more details). Moreover, if \(\sigma \in C^{\alpha }(\Omega {\setminus } \overline{ {\mathcal {U_P}} \cup {\mathcal {U_I}}})\), \(f \in C^{1,\alpha }(\partial \Omega )\), and the boundaries of \( {\mathcal {U_P}}\), \( {\mathcal {U_I}}\), and \(\Omega \) are regular enough, then it follows from standard elliptic regularity results that \(u\in C^{1}(\bar{\Omega } {\setminus } ( {\mathcal {U_P}} \cup {\mathcal {U_I}}))\). Under certain assumptions, it is shown in Theorem 2.1 in [10] (see also the density argument of Proposition 3 in [15]) that the solution of the Eq. (3) is a minimizer of (1), where a is the magnitude of the corresponding current density vector field. Once u is determined, the shape and locations of perfectly conducting and insulation inclusion and the conductivity outside of the inclusions can be uniquely identified. Thus existence is known a priori, and the main issue is to prove uniqueness for the variational problem. Indeed if the conductivity to be recovered is \(C^{\alpha }\), then the assumption of Theorems 1.1 and 1.2 are naturally satisfied in our practical setting. See [9, 10, 13, 15] for further results on this problem with only partial data, with insulating or highly conductive inclusions, reconstruction algorithms, and stability, and also [14] for a review. From the point of view of this original application, the uniqueness result in [6] does not apply to the case of embedded insulated and perfectly conductive inclusions described in [10], where the weight a is merely continuous and may vanish in \(\Omega \).

The uniqueness results presented here yield global convergence in minimization schemes based on compactness (e.g. [15]) and allow for extending stability results based on the Fredholm’s alternative (e.g. in [8]) to the case of vanishing interior data.

Throughout the paper we assume that \(\Omega \subset \mathbb {R}^n\) is a bounded open set with connected Lipschitz boundary \(\partial \Omega \) and f is continuous on \(\partial \Omega \). The following assumptions concern the most general class of weights a we allow. Let \(\mathcal {I}\subset \subset \Omega \) be an open set (possibly empty) with finitely many \(C^1\) connected components with finite perimeter, each of which is \(C^1\)-diffeomorphic with a ball. In addition, in two dimensions \(\mathcal {I}\) is assumed to have at most one such component. We assume that \(a=0\) in \(\mathcal {I}\), \(a \in C(\overline{\Omega }{\setminus } \mathcal {I})\), and that a may have at most countable many zeros in \(\Omega {\setminus } \overline{\mathcal {I}}\). In other words, the set of zeros

$$\begin{aligned} S:=\{x\in \overline{\Omega }: a(x)=0 \} \end{aligned}$$
(4)

satisfy

$$\begin{aligned} \bar{S}=\overline{\mathcal {I}} \cup \Gamma , \end{aligned}$$
(5)

where \(\Gamma \) is a countable set of points in \(\overline{\Omega }{\setminus } \overline{\mathcal {I}}\). Note that we do not assume that a vanishes on \(\partial \mathcal {I}\). Note also that a may be discontinuous at points on \(\partial \mathcal {I}\).

Definition 1

Let \(\Omega \subset \mathbb {R}^n\) be a bounded Lipschitz domain with connected boundary. We define \(BV_{loc}(\Omega {\setminus } S)\) to be the space of all functions \(u \in L^1(\Omega )\) such that

$$\begin{aligned} u \in BV(\Omega {\setminus } S') \ \ \textit{for\;all\;open\;sets}\ \ S' \ \ { with} \ \ S' \supset \supset S. \end{aligned}$$

Recall the following definition from [2]: For any \(u \in BV_{loc}(\Omega {\setminus } S)\) the total variation of u (with respect to the weight a) in \(\Omega \) is defined as

$$\begin{aligned} \int _{\Omega } |Du|_{a}=\sup _{b \in {\mathfrak {B}}_a} \int _{\Omega } u \nabla \cdot b \ \ dx, \end{aligned}$$
(6)

where

$$\begin{aligned} {\mathfrak {B}}_a:=\left\{ b\in L^\infty (\Omega ;{\mathbb {R}}^n): \;\nabla \cdot b \in L^n(\Omega ),\; \text{ supp }(b)\subset \subset \Omega ,\; |b|\le a\; \hbox {a.e. in}\; \Omega \right\} . \end{aligned}$$

This paper considers the question of uniqueness of solutions of the weighted least gradient problem

$$\begin{aligned} \min \left\{ \int _{\Omega }|Dv|_a: v\in BV_{loc}(\Omega {\setminus } S),\; v|_{\partial \Omega }= f \right\} , \end{aligned}$$
(7)

where the boundary condition is in the sense of the trace of functions in \(BV(\Omega )\). Note that, since \(a\equiv 0\) in the open set \({\mathcal {I}}\), then \(\int _{\Omega }|D v|_a\) is independent of \(v|_{\mathcal {I}}\).

Our main uniqueness result assumes the existence of a minimizer \(u \in C^1({\overline{\Omega }{\setminus } {\mathcal {I}}})\) of (7), whose set of singularities (possibly empty) satisfy

$$\begin{aligned} {\mathcal {P}}:=\{x\in \overline{\Omega } {\setminus } {\mathcal {I}}: |\nabla u|=0\} \ \ \hbox {is the union of countably many}\;C^1\hbox {-path-connected sets. } \end{aligned}$$
(8)

Now we are ready to state the more general result of this paper.

Theorem 1.2

Let \(\Omega \subset \mathbb {R}^n\) be a bounded Lipschitz domain with connected boundary, \(f\in C^1(\partial \Omega )\). Assume that the set of zeros of a satisfy the hypothesis (5) and \(a\in C(\overline{\Omega } {\setminus } {\mathcal {I}})\). If (7) has a minimizer \(u\in C^1(\overline{\Omega }{\setminus } {\mathcal {I}})\) that satisfies (8), then u is the unique minimizer of (7) in \(BV_{loc}(\Omega {\setminus } S)\).

Theorem 1.1 follows from the above theorem by taking \(S=\emptyset \). We state other special cases in the next corollaries.

Corollary 1.3

Let \(\Omega \subset \mathbb {R}^n\) be a bounded Lipschitz domain with connected boundary, \(f\in C^1(\partial \Omega )\), and \(a\in C(\overline{\Omega })\) with a countable set S of zeros. If (7) has a minimizer \(u\in C^1(\overline{\Omega })\) with a countable set of critical points, then u is the unique minimizer of (7) in \(BV_{loc}(\Omega {\setminus } S)\).

Corollary 1.4

Let \(\Omega \subset \mathbb {R}^n\) be a bounded Lipschitz domain with connected boundary, \(f\in C^1(\partial \Omega )\), and \(a\in C(\overline{\Omega })\). If S satisfies (4) and the least gradient problem (7) has a minimizer \(u\in C^1(\overline{\Omega })\) such that

$$\begin{aligned} \{x\in \overline{\Omega }: |\nabla u (x)|=0 \}, \end{aligned}$$

is the closure of an open set, then u is the unique minimizer of (7) in \(BV_{loc}(\Omega {\setminus } S)\).

In the original application in [10], one interprets the open subsets of \(\{x\in \overline{\Omega }: |\nabla u (x)|=0 \}\) in which \(a>0\) as perfect conductors. To illustrate a simple case with one perfectly conducting inclusion, consider the following example from [18].

Example 1.5

Let \(D=\{x\in R^2: \ \ x^2+y^2<1\}\) be the unit disk, \(f(x,y)=x^2-y^2\), and \({\mathcal {P}}=(-\frac{1}{\sqrt{2}}, \frac{1}{\sqrt{2}} )\times (-\frac{1}{\sqrt{2}}, \frac{1}{\sqrt{2}})\). It is shown in [18] (see also [10] for a different proof) that

$$\begin{aligned} u=\left\{ \begin{array}{lll} 2x^2-1, \ \ &{}\quad { if}\;|x|\ge \frac{1}{\sqrt{2}}, \ \ |y|\le \frac{1}{\sqrt{2}},\\ 0, \ &{}\quad { if}\; (x,y)\in P,\\ 1-2y^2, \ \ &{}\quad { if} \;|x|\le \frac{1}{\sqrt{2}}, \ \ |y|\ge \frac{1}{\sqrt{2}}\\ \end{array} \right. \end{aligned}$$

is a minimizer of the least gradient problem

$$\begin{aligned} min\left\{ \int _{D}|\nabla u| dx, \ \ u\in BV(D), \ \ and\ \ u|_{\partial D}=f\right\} . \end{aligned}$$
(9)

It is easy to see that u satisfies the assumptions of Corollary 1.4 with \(S={\mathcal {I}}=\emptyset \) and \({\mathcal {P}}=(-\frac{1}{\sqrt{2}}, \frac{1}{\sqrt{2}}) \times (-\frac{1}{\sqrt{2}},\frac{1}{\sqrt{2}})\). Hence u is the only minimizer in \(BV(\Omega )\).

To understand the main ideas of the proofs in Sects. 2 and 3, one can keep in mind the example above and assume \(S={\mathcal {I}}=\emptyset \) at the first reading, to avoid some of the technicalities.

2 Preliminaries

In this section we recall and present some preliminary results that will be used in the following sections. First we recall a useful representation formula from [2]. For \(u\in BV(\Omega )\)

$$\begin{aligned} \int _{A}|Du|_a=\int _{A} h(x,v^u)|Du|, \end{aligned}$$
(10)

where

$$\begin{aligned} h(x,v^u)=|Du|-\hbox {ess} \sup _{b\in {\mathfrak {B_a}}}( b \cdot v^u)(x) \ \ \ \ for \ \ |Du|-a.e. \ \ x\in \Omega \end{aligned}$$
(11)

and \(v^u\) denotes the Radon–Nikodym derivative

$$\begin{aligned} v^{u}(x)=\frac{d\, Du}{d\, |Du|}. \end{aligned}$$
(12)

In particular, if \(u\in BV(\Omega )\), and the coefficient a is continuous in the Borel measurable subset \(A \subset \Omega \), then

$$\begin{aligned} \int _{A}|Du|_a=\int _{A} a|Du|, \end{aligned}$$
(13)

as shown in [2]. The following lemma provides an extension of this formula.

Lemma 2.1

Let \(\Omega \subset \mathbb {R}^n\) be a bounded open region with Lipschitz boundary. Suppose that u is a minimizer of (7) and satisfies the assumptions of Theorem 1.2. Then

$$\begin{aligned} \int _{\Omega }|D u|_a=\int _{\Omega }a |\nabla u| dx. \end{aligned}$$

Proof

Since, \(a \in C(\Omega {\setminus } ({\mathcal {I}}\cup {\mathcal {P}}))\). Hence by [2, Proposition 7.1] we have that

$$\begin{aligned} h(x,v^u)= \left\{ \begin{array}{lll} a(x) \ \ &{}\hbox {in} \ \ &{}\Omega {\setminus } \overline{{\mathcal {I}}\cup {\mathcal {P}}}\\ 0 \ \ &{}\hbox {in}\ \ &{}{\mathcal {I}}. \\ \end{array} \right. \end{aligned}$$
(14)

Thus it follows from (10) that

$$\begin{aligned} \int _{\Omega } |Du|_a=\int _{\Omega {\setminus } \overline{{\mathcal {I}}\cup {\mathcal {P}}}} a|\nabla u|=\int _{\Omega } a|\nabla u|dx. \end{aligned}$$

\(\square \)

Let \(\nu \) denote the outer unit normal vector to \(\partial \Omega \) and

$$\begin{aligned} X:= \{b\in L^{\infty }(\Omega ;{\mathbb {R}}^n): \; \nabla \cdot b \in L^n(\Omega )\}. \end{aligned}$$

Then for every \(b\in X\) there exists a unique function \([b\cdot \nu ] \in L^{\infty }_{{\mathcal {H}}^{n-1}}(\partial \Omega )\) such that

$$\begin{aligned} \int _{\partial \Omega } [b\cdot \nu ] u d{\mathcal {H}}^{n-1}=\int _{\Omega } u \nabla \cdot b dx +\int _{\Omega } b \cdot \nabla u dx, \ \ \forall u \in C^1(\overline{\Omega }). \end{aligned}$$
(15)

Moreover, for \(u\in BV(\Omega )\) and \(b\in X\), the linear functional \(u\mapsto (b \cdot Du)\) gives rise to a Radon measure on \(\Omega \), and

$$\begin{aligned} \int _{\partial \Omega } [b\cdot \nu ]u d{\mathcal {H}}^{n-1}=\int _{\Omega } u \nabla \cdot b dx +\int _{\Omega } (b \cdot D u), \ \ \forall u \in BV(\Omega ), \end{aligned}$$
(16)

see [1, 3] for a proof. We will need the following generalization of (16) in the proof of our uniqueness result.

Lemma 2.2

Let S be as defined in (5) and \(b\in {\mathfrak {B}}_a\). If \(u\in L^{\infty }(\Omega )\) and \(\int _{\Omega }|Du|_a<\infty \), then

$$\begin{aligned} \int _{\partial \Omega } [{b}\cdot \nu ]u d{\mathcal {H}}^{n-1}=\int _{\Omega } u \nabla \cdot {b} dx +\int _{\Omega } ({b} \cdot D u), \end{aligned}$$
(17)

for some unique function \([b\cdot \nu ]\) in \(L^{\infty }_{{\mathcal {H}}^{n-1}}(\partial \Omega )\).

Proof

By hypothesis (5) S has finite perimeter in \(\Omega \). Define the set

$$\begin{aligned} {\mathcal {D}}:=\partial S \cap S. \end{aligned}$$

There exists a sequence \({\mathcal {D}}_i\) of open subsets of \(\Omega \) with uniformly bounded perimeter such that \({\mathcal {D}}_{i+1}\subset {\mathcal {D}}_{i}\) and \(u\in BV({\mathcal {D}}_i)\) for all \(i\in \mathbb {N}\), and \(\cap _{i=1}^{\infty }{\mathcal {D}}_i={\mathcal {D}}\). Now choose a sequence of cut-off functions \(\varphi _i\) such that \(\varphi \equiv 0\) on \({\mathcal {D}}_i\), \(\varphi \equiv 1\) on the compliment of an open set \( {{\mathcal {D'}}}_i \supset \supset {\mathcal {D}}_i\) with \(\lim _{i\longrightarrow \infty }{\mathcal {H}}^n({\mathcal {D'}}_i)={\mathcal {H}}^n({\mathcal {D}})\), and \(\int _{\Omega }|D\varphi _i|<c\) for all i. Indeed \(\varphi _i\) could be chosen to be smooth approximations of the characteristic functions \(\chi _{{\mathcal {D}}_i^c}\) in \(BV(\mathbb {R}^n)\) and c proportional to the perimeter of \({\mathcal {D}}_i\). Then \(\varphi _i u \in BV(\Omega )\) and hence by (16)

$$\begin{aligned} \int _{\partial \Omega } [{b}\cdot \nu ]\varphi _i u d{\mathcal {H}}^{n-1}= & {} \int _{\Omega } \varphi _i u \nabla \cdot {b} dx +\int _{\Omega } ({b} \cdot D (\varphi _i u))\nonumber \\= & {} \int _{\Omega } \varphi _i u \nabla \cdot {b} dx +\int _{\Omega } ({b} \cdot D \varphi _i)u+\int _{\Omega } \varphi _i ({b} \cdot D u) \end{aligned}$$
(18)

for all \(b\in {\mathfrak {B}}\) and \(i\in \mathbb {N}\). Note that \(u\in L^{\infty }(\Omega )\) and \(|b|\le a\) a.e. in \(\Omega \). It follows from the continuity of a in \({\mathcal {D}} \subset S=\{x\in \overline{\Omega }: a(x)=0\}\) that \(\lim _{i\rightarrow \infty } \parallel b\parallel _{L^{\infty }(D'_i{\setminus } D_i)} \rightarrow 0\). Hence

$$\begin{aligned} \left| \int _{\Omega } ({b} \cdot D \varphi _i)u \right| \le \int _{D'_i{\setminus } D_i} |({b} \cdot D \varphi _i)u| \le \int _{D'_i{\setminus } D_i}|D\varphi _i| \parallel u\parallel _{\infty } \parallel b\parallel _{L^{\infty }(D'_i{\setminus } D_i)} \longrightarrow 0, \end{aligned}$$

as \(i\longrightarrow \infty \). Since \({\mathcal {I}}\subset \subset \Omega \) and \(\Gamma \) [defined in (4)] is countable, by letting \(i\longrightarrow \infty \) in (18) we obtain (17). \(\square \)

The next two results yield a calibration which will be used in the uniqueness proof.

Suppose \(a\in L^2(\Omega )\) and fix \(v_f \in H^1(\Omega )\) with \(v_f|_{\partial \Omega }=f\). Consider the weighted least gradient problem

$$\begin{aligned} (P) \ \ \ \ \inf _{v\in H^1_0(\Omega )} \int _{\Omega } a|\nabla v+\nabla v_f|dx. \end{aligned}$$

In [9] it is shown that the dual problem to (P) is

$$\begin{aligned} (D) \ \ \ \ \max \left\{ <\nabla v_f, b>: \ \ b\in L^2(\Omega ;{\mathbb {R}}^n),\ \ |b(x)|\le a(x) \ \ a.e. \ \ \hbox {and} \ \ \nabla \cdot b\equiv 0\right\} . \end{aligned}$$

Let v(P) and v(D) be the optimal values of the primal and dual problems. It is shown in [9] that \(v(P)=v(D)\) and the dual problem (D) has an optimal solution. The following proposition is an immediate consequence of Proposition 2.1 and Corollary 2.3 in [9].

Proposition 2.1

Let \(a\in L^2(\Omega )\) be a non-negative function and \(v_f \in H^1(\Omega )\) with \(v_f|_{\partial \Omega }=f\). Then the optimal values of the primal problem (P) and dual problem (D) are equal, and the dual problem (D) has an optimal solution J with \(\nabla \cdot J\equiv 0\) in \(\Omega \). Moreover, if \(v\in H^1_0(\Omega )\) is an optimal solution of the primal problem (P), then

$$\begin{aligned} J(x)=a(x)\frac{\nabla ( v(x)+v_f(x))}{|\nabla (v(x)+ v_f(x))|} \ \ { if} \ \ |\nabla ( v(x)+v_f(x))|\ne 0, \end{aligned}$$

for all \(x\in \Omega \).

The following result is an immediate consequence of Proposition 2.1.

Corollary 2.3

Let \(\Omega \subset \mathbb {R}^n\) be a bounded Lipschitz domain and assume that assumptions of Theorem 1.2 are satisfied. Then there exists an optimal solution \(J\in L^{2}(\Omega ;{\mathbb {R}}^n)\) of the dual problem (D) such that \(\nabla \cdot J\equiv 0\) in \(\Omega \), \(|J|\le a\) a.e. in \(\Omega \). Moreover

$$\begin{aligned} J(x)= \left\{ \begin{array}{ll} a(x)\frac{\nabla u}{|\nabla u|}\ \ &{}\quad { if} \ \ |\nabla u|\ne 0\\ 0 \ \ &{}\quad { if}\ \ |\nabla u|=0, \\ \end{array} \right. \end{aligned}$$
(19)

where u is the solution of (7) described in the statement of Theorem 1.2.

3 Uniqueness of the minimizers

In this section we prove Theorem 1.2. To understand the main ideas of the proofs in this section one may assume \(S={\mathcal {I}}=\emptyset \) at first reading.

Let u be the minimizer of the weighted least gradient problem (7) assumed in the statement of Theorem 1.2, and suppose \(u^*\in BV_{loc}(\Omega {\setminus } S)\) is another minimizer. We will show that \(u=u^*\) a.e. in \(\Omega {\setminus } S\). We will prove Theorem 1.2 in four steps. First (Step I) we prove that \(\frac{\nabla u}{|\nabla u|}=v^{u^*}\) \(Du^*\)-a.e. in \(\Omega {\setminus } \overline{{\mathcal {I}}\cup {\mathcal {P}}}\), where \(v^{u^*}\) is the Radon–Nikodym derivative in (12). In Step II we prove that almost every level set of \(u^*\) is also a level set of u. In Step III, we prove that almost every level set of \(u^*\) reaches \(\partial \Omega \). Finally in Step IV, we show that on almost every level set of \(u^*\), u and \(u^*\) take the same values and therefore \(u=u^*\) a.e. in \(\Omega {\setminus } S\).

Step I. :

First notice that \(u^*\) is bounded above and below almost everywhere. Indeed if we define

$$\begin{aligned} \bar{u}(x)= \left\{ \begin{array}{ll} u^*(x)\ \ &{}\quad \hbox {if} \ \ m_f\le u^*(x) \le M_f\\ M_f \ \ &{}\quad \hbox {if}\ \ u^*(x)>M_f, \\ m_f\ \ &{}\quad \hbox {if}\ \ u^*(x)<m_f, \end{array} \right. \end{aligned}$$
(20)

where \(M_f\) and \(m_f\) are the maximum and minimum values of f on \(\partial \Omega \), then \(\bar{u} \in BV_{loc}(\Omega {\setminus } S)\) and

$$\begin{aligned} \int _{\Omega }|D\bar{u}|_a \le \int _{\Omega }|Du^*|_a. \end{aligned}$$
(21)

Moreover the inequality is strict if \(\{x\in \Omega : u^*(x)>M_f\}\) or \(\{x\in \Omega : u^*(x)<m_f\}\) has positive measure. Therefore we have \(\hbox {range}(u^*)= \hbox {range}(f)\).

Next we prove that

$$\begin{aligned} \frac{\nabla u}{|\nabla u|}=\frac{d Du^*}{d|Du^*|} \end{aligned}$$

\(|Du^*|-a.e.\) in \(\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}\).

Lemma 3.1

Suppose that the assumptions of Theorem 1.2 are satisfied and u is the minimizer of (7) assumed in the statement of Theorem 1.2. If \(u^*\) is another minimizer of (7), then

$$\begin{aligned} \frac{\nabla u}{|\nabla u|}=v^{u^*} \ \ \ \ |Du^*|-a.e.\ \ \hbox {in} \ \ \Omega {\setminus } \overline{{\mathcal {I}}\cup {\mathcal {P}}}. \end{aligned}$$

Proof

It follows from the definition of \(h(x, v^{u^*})\) that

$$\begin{aligned} h(x,v^{u^*})\ge J\cdot v^{u^*}, \ \ \ \ |Du^*|-a.e. \ \ \hbox {in} \ \ \Omega , \end{aligned}$$

where J is the solution of the dual problem (D) in Corollary 2.3. Hence

$$\begin{aligned} \int _{\Omega } |Du^*|_a= & {} \int _{\Omega }h(x, v^{u^*})|Du^*| \ge \int _{\Omega } J \cdot v^{u^*} |Du^*|\\= & {} \int _{\Omega } J\cdot Du^*= \int _{\partial \Omega } J\cdot \nu f d{\mathcal {H}}^{n-1}\\= & {} \int _{\Omega }\nabla u\cdot J dx= \int _{\Omega }|J| |\nabla u| \\= & {} \int _{\Omega } |Du|_a=\int _{\Omega } |Du^*|_a, \end{aligned}$$

where the third and fifth equalities follow form Lemmas 2.2 and 2.1, respectively. Therefore

$$\begin{aligned} h(x,v^{u^*})=J\cdot v^{u^*},\ \ |Du^*|-a.e. \ \ \hbox {in} \ \ \Omega . \end{aligned}$$

Since a is continuous in \(\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}\), as in (14) we have

$$\begin{aligned} J\cdot v^{u^*}=h(x, v^{u^*})=a(x), \ \ |Du^*|-a.e. \ \ \hbox {in} \ \ \Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}. \end{aligned}$$
(22)

Since \(|v^{u^*}|=1\) and \(|J|\le a\), \(|Du^*|-a.e.\) in \(\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}\), we get

$$\begin{aligned} \frac{J}{|J|} =v^{u^*}, \ \ |Du^*|-a.e. \ \ \hbox {in} \ \ \Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}. \end{aligned}$$

On the other hand \(|\nabla u|\ne 0\) \(|Du^*|\)-a.e. on \(\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}\). Hence it follows from Corollary 2.3 that

$$\begin{aligned} \frac{\nabla u}{|\nabla u|}=\frac{J}{|J|} =v^{u^*}, \ \ |Du^*|-a.e. \ \ \hbox {in} \ \ \Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}. \end{aligned}$$

The proof is now complete. \(\square \)

Step II:

For \(\lambda \in \hbox {range}(u^*)\), let

$$\begin{aligned} E_{\lambda }=\{x\in \Omega {\setminus } \overline{{\mathcal {I}}}: \ \ u^*(x) \ge \lambda \} \end{aligned}$$

and define

$$\begin{aligned} E'_\lambda :=\left\{ x\in \mathbb {R}^n: \lim _{r \rightarrow 0} \frac{{\mathcal {H}}(B(r,x)\cap E_\lambda )}{{\mathcal {H}}(B(r))}=1\right\} .\ \end{aligned}$$
(23)

By changing \(u^*\) in a set of measure zero, we may assume that \(E_\lambda =E'_{\lambda }\). Throughout this paper we shall always assume that \(E_\lambda =E'_{\lambda }\). We also define

$$\begin{aligned} {\mathcal {Z}}=\{x\in \overline{\Omega }{\setminus } {\mathcal {I}}: u(x)\in \overline{\Gamma \cup {\mathcal {P}}}\}, \end{aligned}$$
(24)

where \(\Gamma \) is defined in (4) and \({\mathcal {P}}\) is the set of critical points of u satisfying (8). Notice that if \(x \notin {\mathcal {Z}}\), then \(|\nabla u(x)|>0\).

Let \(\Lambda \) be the set of all \(\lambda \in \hbox {range}(u^*)\) such that every connected component \(\Sigma \) of \(\partial E_\lambda \) with \(\Sigma \cap {\mathcal {Z}}=\emptyset \) is a \(C^1\) hypersurface. In the next lemma we prove that

$$\begin{aligned} {\mathcal {H}}^1(\hbox {range}(u^*){\setminus } \Lambda )=0. \end{aligned}$$
(25)

Lemma 3.2

Assume that the assumptions of Theorem 1.2 are satisfied and u is the corresponding minimizer of (7). Let \(u^* \in BV_{loc}(\Omega {\setminus } S)\) be another minimizer and \(\Sigma \) be a connected component of \(\partial E_\lambda \) for some \(\lambda \in \hbox {range}(u^*)\). Then for almost every \(\lambda \in \hbox {range} (u^*)\) one and only one of the following statements hold:

  1. (i)

    \(\Sigma \subset {\mathcal {Z}}\).

  2. (ii)

    \(\Sigma \cap {\mathcal {Z}}=\emptyset \), \(\Sigma \) is a \(C^1\) hypersurface, and u is constant on \(\Sigma \).

Proof

By the co-area formula we have

$$\begin{aligned} 0=\int _{\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}} \varphi \left[ \frac{\nabla u}{|\nabla u|} - v^{u^*}\right] |Du^*| = \int _{0}^{\infty } \int _{\partial ^* E_{\lambda } \cap (\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}})} \varphi \left[ \frac{\nabla u}{|\nabla u|} -v^{u^*}\right] d {\mathcal {H}}^{n-1} d\lambda \end{aligned}$$
(26)

for every smooth vector field \(\varphi \), where \(\partial ^*E_\lambda \) is the reduced boundary of \(E_\lambda \). Therefore \(\frac{\nabla u}{|\nabla u|}=v^{u^*}\), \({\mathcal {H}}^{n-1}-a.e.\) in \(\partial ^* E_\lambda \cap (\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}})\) for almost every \(\lambda \in \hbox {range} (u^*)\). Since \(|D\chi _{E_{\lambda }}|\) is the \((n-1)-\)dimensional Hausdorff measure restricted to \(\partial ^* E_\lambda \) (see Chapter 4 in [4]), for almost every \(\lambda \in \hbox {range}(u^*)\) the generalized normal \(\nu (x)\) exists for \(|D\chi _{E_{\lambda }}|-a.e.\) \(x\in \partial E_\lambda \cap (\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}})\) and coincides there with the continuous vector field \(\frac{\nabla u}{|\nabla u|}\).

Now let \(x\in \partial E_\lambda \cap \Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}}\) such that \(x \notin {\mathcal {Z}}\). Since \({\mathcal {Z}}\) is closed, there exists \(\epsilon >0\) such that \(B_{\epsilon }(x)\cap {\mathcal {Z}}=\emptyset \). By Theorem 4.8 in [4], \(\partial E_\lambda \cap B_{\epsilon }(x)\) can be represented as the graph of a Lipschitz continuous function g. Thus the derivative of g coincides almost everywhere with a continuous function and therefore g must be \(C^1\). Hence we conclude that if \(x \in [\partial E_\lambda \cap (\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}})] {\setminus } Z\), then \(\partial E_\lambda \) is a \(C^1\) hyperspace near x for almost every \(\lambda \in \hbox {range}(u^*)\).

Next we show that u is constant on every \(C^1\) connected open subset of \(\Sigma \) of \(\partial E_\lambda \cap (\Omega \backslash \overline{{\mathcal {I}} \cup {\mathcal {P}}})\). Let \(\gamma :(-\epsilon ,+\epsilon ) \rightarrow \Sigma \) be an arbitrary \(C^1\) curve. Then

$$\begin{aligned} \frac{d}{dt}u(\gamma (s))=|\nabla u(\gamma (s))|\nu (\gamma (s)).\gamma '(s)=0, \end{aligned}$$

because either \(|\nabla u(\gamma (s))|=0\) or \(\nu (\gamma (s)).\gamma '(s)=0\) on \(\Sigma \). Thus u is constant along \(\gamma \) and hence u is constant on \(\Sigma \). Therefore it follows from continuity of u and the definition of the set \({\mathcal {Z}}\) that, for almost every \(\lambda \in \hbox {range}(u^*)\), if \(\Sigma \not \subset Z\) is a connected component of \(\partial E_\lambda \cap (\Omega {\setminus } \overline{{\mathcal {I}} \cup {\mathcal {P}}})\), then \(\Sigma \) is a \(C^1\) hypersurface, u is constant on \(\Sigma \), and \(\Sigma \cap {\mathcal {Z}}=\emptyset \). The proof is now complete. \(\square \)

Step III:

We show next that every connected component of \(\partial E_{\lambda }\) intersects the boundary \(\partial \Omega \).

Proposition 3.1

Assume that the assumptions of Theorem 1.2 are satisfied and u is the corresponding minimizer of (7). Suppose \(\Sigma _{\lambda }\) is a \(C^1\) connected component of \(\partial E_{\lambda }=\partial \{x \in \Omega {\setminus } {\mathcal {I}}: \ \ u^*(x)>\lambda \}\) and \(\Sigma _{\lambda }\cap {\mathcal {Z}}=\emptyset \). Then

$$\begin{aligned} \overline{\Sigma }_{\lambda }\cap \partial \Omega \ne \emptyset . \end{aligned}$$

Proof

Assume \(\overline{\Sigma }_{\lambda }\cap \partial \Omega = \emptyset \). Then one of the followings statements hold:

  1. (I)

    \(\overline{\Sigma }_\lambda \) is a manifold without boundary in \(\overline{\Omega } {\setminus } {\mathcal {I}}\).

  2. (II)

    \(\overline{\Sigma }_\lambda \cap \partial {\mathcal {I}} \ne \emptyset \).

Case I:

Assume that \(\overline{\Sigma }_\lambda \) is a manifold without boundary in \(\overline{\Omega }\). Then, since \(\partial \Omega \) is connected, \(\partial \Omega \cup \Sigma _\lambda \) is a compact manifold with two connected components. By the Alexander duality theorem for \(\partial \Omega \cup \Sigma _\lambda \) (see, e.g., Theorem 27.10 in [5]) we have that \(\mathbb {R}^n {\setminus }(\partial \Omega \cup \Sigma _\lambda )\) is partitioned into three open connected components: \(\mathbb {R}^n=(\mathbb {R}^n{\setminus } \overline{\Omega }) \cup O_1\cup O_2\). Since \(\Sigma _\lambda \subset \Omega \) we have \(O_1 \cup O_2=\Omega {\setminus } \Sigma _\lambda \) and then \(\partial O_i \subset \partial \Omega \cup \Sigma _\lambda \), for \(i=1,2\). We claim that at least one of the \(\partial O_1\) or \(\partial O_2\) is in \(\Sigma _\lambda \). Assume not, i.e. for \(i=1,2\), \(\partial O_i \cap \partial \Omega \ne \emptyset \). Since \(\partial \Omega \) is connected (by assumption) we have that \(O_1 \cup O_2\cup \partial \Omega \) is connected which implies that \(O_1 \cup O_2\cup (\mathbb {R}^n{\setminus } \Omega )\) is also connected. Again by applying the Alexander duality theorem for \(\Sigma _\lambda \subset \mathbb {R}^n\), we have that \(\mathbb {R}^n {\setminus } \Sigma _\lambda \) has exactly two open connected components, one of which is unbounded: \(\mathbb {R}^n {\setminus } \Sigma _\lambda =O_{\infty }\cup O_0\). Since \(O_1\cup O_2 \cup (\mathbb {R}^n \backslash \Omega )\) is connected and unbounded, we have that \(O_1\cup O_2 \cup (\mathbb {R}^n \backslash \Omega ) \subset O_{\infty }\), which leaves \(O_0 \subset \mathbb {R}^n {\setminus } (O_1 \cup O_2 \cup (\mathbb {R}^n {\setminus } \Omega )) \subset \Sigma _\lambda \). This is impossible since \(O_0\) is open and \(\Sigma _\lambda \) is a hypersurface. Therefore either \(\partial O_1\) or \(\partial O_2\) or both lie in \(\Sigma _\lambda \). Assume \(\partial O_1\subset \Sigma _\lambda \). We claim that u is constant in \(O_1\). Indeed, by Lemma 3.2, \(u=c\) on \(\Sigma _{\lambda }\) for some c. Hence the new map \(\tilde{u}\) defined by

$$\begin{aligned} \tilde{u}:=\left\{ \begin{array}{ll} u,\ \ &{}\quad x \in \Omega {\setminus } O_1,\\ c, \ \ \ \ &{}\quad x\in \overline{O_1}, \end{array} \right. \end{aligned}$$

is in \(BV_{loc}(\Omega {\setminus } S)\) and decreases the energy, which contradicts the minimality of u. Therefore \(u=c\) in \(O_1\). This is a contradiction since we have assumed \(\overline{\Sigma }_{\lambda } \cap {\mathcal {Z}}=\emptyset \).

Case II:

Suppose \(\overline{\Sigma }_\lambda \cap \partial {\mathcal {I}} \ne \emptyset \). We first consider the dimensions \(n\ge 3\). Let

$$\begin{aligned} \epsilon ^*:= { min}\left\{ \min _{i\ne j} { dist}({\mathcal {I}}_i, {\mathcal {I}}_j), \min _{i} { dist}({\mathcal {I}}_i, \partial \Omega )\right\} , \end{aligned}$$

where \({\mathcal {I}}_i\), \(1\le i\le m\), are the open connected components of the set \({\mathcal {I}}\). For any \(0<\epsilon <\epsilon ^*:\) define

$$\begin{aligned} {\mathcal {I}}^{\epsilon }=\{x\in \Omega : { dist}(x, {\mathcal {I}})<\epsilon \}. \end{aligned}$$

Then \({\mathcal {I}}^{\epsilon }\) is an open set with the same number of disjoint open connected components as \({\mathcal {I}}\). Now let \(\Sigma _{\lambda }^{\epsilon }= \Sigma _\lambda {\setminus } {\mathcal {I}}^{\epsilon }\) which we know is \(C^1\) on \(\Omega {\setminus } {\mathcal {I}}^{\epsilon }\). Since \(\partial \Sigma _{\lambda }^{\epsilon } \subset \partial {\mathcal {I}}^{\epsilon }\) and \(\partial {\mathcal {I}}^\epsilon {\setminus } \partial \Sigma _{\lambda }^{\epsilon } \) is open, each connected component of \(\partial \Sigma _\lambda ^{\epsilon }\) is the boundary of an open set in \(\partial {\mathcal {I}}^{\epsilon }\) with connected boundary. Suppose M is a connected component of \(\partial \Sigma _{\lambda }^{\epsilon }\). Then \(M \subset \partial {\mathcal {I}}^{\epsilon }_i\) for some \(1 \le i \le m\), \({\mathcal {I}}^{\epsilon }_i\) is \(C^1\)-diffeomorphic image of the unit ball for \(\epsilon \) small, and M is an orientable manifold without boundary in \(\partial {\mathcal {I}}^{\epsilon }\). Therefore it follows from Alexander’s duality theorem that

$$\begin{aligned} \partial {\mathcal {I}}_i^{\epsilon } {\setminus } M= V_1 \cup V_2, \end{aligned}$$

where \(V_1, V_2\) are disjoint open connected (with respect to the topology of \(\partial {\mathcal {I}}^{\epsilon }\)) sets. Since \(\Sigma ^{\epsilon }_{\lambda }\) can be extended to a \(C^1\) hypersurface \(\Sigma _\lambda \) inside \( {\mathcal {I}}^{\epsilon } {\setminus } {\mathcal {I}}\), we can extend \(\Sigma _\lambda ^{\epsilon }\) inside \({\mathcal {I}}^{\epsilon }_i\) to obtain a \(C^1\) hypersurface \(H^{\epsilon }_\lambda \) such that

$$\begin{aligned} H^{\epsilon }_\lambda \cap (\Omega {\setminus } {\mathcal {I}}^{\epsilon })=\Sigma _\lambda ^{\epsilon } \cap (\Omega {\setminus } {\mathcal {I}}^{\epsilon }) \end{aligned}$$

and \(\partial (H^{\epsilon }_\lambda \cap {\mathcal {I}}^{\epsilon })=M\). Repeating this argument for other connected components of \(\partial \Sigma _\lambda ^{\epsilon }\) leads to a \(C^1\) orientable hypersurface \({\mathcal {S}}^{\epsilon }_\lambda \) with no boundary such that \(\partial \Omega \cap {\mathcal {S}}^{\epsilon }_\lambda =\emptyset \) and \({\mathcal {S}}^{\epsilon }_\lambda \cap (\Omega {\setminus } {\mathcal {I}}^{\epsilon }) =\partial \Sigma ^{\epsilon }_\lambda \). Now apply Alexander’s duality theorem to get the partition

$$\begin{aligned} \mathbb {R}^n {\setminus } {\mathcal {S}}^{\epsilon }_\lambda =O_0^{\epsilon } \cup O^{\epsilon }_{\infty }, \end{aligned}$$

where \(O_0^{\epsilon }\) and \(O^{\epsilon }_{\infty }\) are open subsets of \(\mathbb {R}^n\) and \(O^{\epsilon }_{\infty }\) is unbounded. If \(\epsilon '< \epsilon \), then \(\Sigma ^{\epsilon }_\lambda \subset \Sigma ^{\epsilon '}_\lambda \) and \({\mathcal {I}}^{\epsilon '} \subset {\mathcal {I}}^{\epsilon }\). Hence

$$\begin{aligned} {\mathcal {I}}^{\epsilon } {\setminus } \bar{O}_0^{\epsilon } \subset {\mathcal {I}}^{\epsilon '} {\setminus } \bar{O}_0^{\epsilon '}. \end{aligned}$$

Now let

$$\begin{aligned} O=\cup _{0<\epsilon <\epsilon ^{*}} \left( {\mathcal {I}}^{\epsilon } {\setminus } \bar{O}_0^{\epsilon }\right) . \end{aligned}$$

Then O is open and \(\partial O \subset \Sigma _\lambda \cup \bar{{\mathcal {I}}}\). We claim that u is constant in O. Indeed by Lemma 3.2, \(u=c\) on \(\Sigma _{\lambda }\) for some constant c. Define

$$\begin{aligned} \tilde{u}:=\left\{ \begin{array}{ll} u,\ \ &{}\quad x \in \Omega \backslash O,\\ c, \ \ \ \ &{}\quad x\in O.\\ \end{array} \right. \end{aligned}$$
(27)

Then \( \tilde{u} \in BV_{loc}(\Omega {\setminus } S)\) which contradicts the minimality of u. Hence u is constant in O which is a contradiction because we have assumed \(\overline{\Sigma }_{\lambda } \cap {\mathcal {Z}}=\emptyset .\) Now assume \(n=2\). Since \(\overline{\Sigma }_\lambda \cap \partial \Omega =\emptyset \) and \({\mathcal {I}}\) has only one connected component, there exists two distinct point \(a,b \in \bar{\Sigma }_\lambda \cap \partial {\mathcal {I}}\) such that

$$\begin{aligned} \partial {\mathcal {I}} {\setminus } \{a,b\}= V_1 \cup V_2. \end{aligned}$$

Note that \(\Sigma _\lambda \cup V_1\) is a continuous closed curve in \(\mathbb {R}^2\). By the Jordan Curve Theorem there exists a bounded open set O such that \(\partial O=\Sigma _\lambda \cup V_1\). Define \(\tilde{u}\) by (27), then with a similar argument we reach a contradiction. In both cases (I) and (II) the contradiction follows from the assumption that \(\overline{\Sigma }_\lambda \cap \partial \Omega = \emptyset \).\(\square \)

Step IV:

Since \(f \in C^1(\partial \Omega )\), f can be extended to a function in \(C^1(\mathbb {R}^n {\setminus } \Omega ) \cap BV(\mathbb {R}^n {\setminus } \Omega )\). We will denote the extension of f to \(\Omega ^c\) by f, again. We will also denote the continuous extension of \(u^*\) to \(\mathbb {R}^n\) with \(u^*=f\) on \(\Omega ^c\) by \(u^*\) again. Define

$$\begin{aligned} F_\lambda =\{x\in \mathbb {R}^n{\setminus } \bar{{\mathcal {I}}}: \ \ u^*(x)\ge \lambda \} \end{aligned}$$

and let the corresponding \(F'_\lambda \) be defines as (23).

Remark 3.3

Let \(\Lambda \subset \hbox {range}(u^*)\) be the set defined by Lemma 3.2 and \(\lambda \in \Lambda \). By Lemma 3.2 every connected component of \(\partial F'_\lambda \cap (\Omega {\setminus } {\mathcal {Z}})\) is a \(C^1\) hypersurface. Since \(F_\lambda \cap (\Omega {\setminus } {\mathcal {Z}}) \) differs from \(F'_\lambda \cap (\Omega {\setminus } {\mathcal {Z}}) \) on a set of measure zero, we may assume that \(F_\lambda \cap (\Omega {\setminus } {\mathcal {Z}}) \) is open.

The proof of the following lemma is very similar to that of Theorem 3.7 in [17]. We include the proof for the convenience of the reader.

Lemma 3.4

Let \(\Omega \) be a bounded domain with connected Lipschitz boundary. If \(x \in \partial ^*F_\lambda \cap \partial \Omega \), where \(\partial ^* F_\lambda \) is the reduced boundary of \(F_\lambda \), and

then \(\lambda =f(x)\).

Proof

Assume \(f(x)<\lambda \). Then

$$\begin{aligned} 0= & {} \lim _{r\rightarrow 0} \frac{1}{|B_r(x)\cap \Omega |} \left( \int _{B_r(x)\cap \Omega \cap \{u^*< \lambda \}} |u^*(y)-f(x)|dy\right. \\&\left. + \int _{B_r(x)\cap \Omega \cap \{u^*\ge \lambda \}} |u^*(y)-f(x)|dy\right) \\\ge & {} \limsup _{r\rightarrow 0}\frac{1}{|B_r(x)\cap \Omega |} \int _{B_r(x)\cap \Omega \cap \{u^*\ge \lambda \}} |u^*(y)-f(x)|dy\\\ge & {} (\lambda -f(x)) \limsup _{r\rightarrow 0} \frac{|B_r(x)\cap \Omega \cap \{u^* \ge \lambda \}|}{|B_r(x)\cap \Omega |}. \end{aligned}$$

Consequently

$$\begin{aligned} \limsup _{r\rightarrow 0} \frac{|B_r(x)\cap \Omega \cap \{u^* \ge \lambda \}|}{|B_r(x)\cap \Omega |}=0. \end{aligned}$$

On the other hand since f is the trace of \(u^*\in BV(\mathbb {R}^n {\setminus } \Omega )\) on \(\partial \Omega \), with a similar argument we conclude that

$$\begin{aligned} \limsup _{r\rightarrow 0} \frac{|B_r(x)\cap (\mathbb {R}^n {\setminus } \Omega )\cap {\{u^*}\ge \lambda \}|}{|B_r(x)\cap (\mathbb {R}^n{\setminus } \Omega )|}=0. \end{aligned}$$

Therefore

$$\begin{aligned} \lim _{r\rightarrow 0} \frac{|B_r(x) \cap \{u^* \ge \lambda \}|}{|B_r|}=0 \end{aligned}$$

and hence \(x \not \in \partial ^* E_\lambda \) which is a contradiction. Similarly \(f(x)>\lambda \) leads to a contradiction. Thus \(f(x)=\lambda \). \(\square \)

Proposition 3.2

Assume that the assumptions of Theorem 1.2 hold, and let \(u^*\) be a the corresponding minimizer of (7). Then for almost every \(\lambda \in \Lambda \)

$$\begin{aligned} u(\partial F_{\lambda }\cap (\overline{\Omega }{\setminus } {\mathcal {Z}}))=\{\lambda \}, \end{aligned}$$

where \(\Lambda \) and \({\mathcal {Z}}\) are defined by (25) and (24), respectively.

Proof

In view of Remark 3.3 and Proposition 3.1, we may assume that \(F_\lambda \cap (\Omega {\setminus } {\mathcal {Z}})\) is open and every connected component of \(\partial F_\lambda \cap (\Omega {\setminus } {\mathcal {Z}})\) is a \(C^1\) hypersurface intersecting \(\partial \Omega \). Now let \(\Sigma \) be a connected component of \(\partial F_\lambda \cap (\Omega {\setminus } {\mathcal {Z}})\). By Proposition 3.1, \(\bar{\Sigma }\cap \partial \Omega \ne \emptyset \). Let \(x_0\in \bar{\Sigma }\cap \partial \Omega \ne \emptyset \). Since \(x_0 \not \in {\mathcal {Z}}\), \(|\nabla u(x_0)|> 0\). On the other hand, note that if \(x_0 \in \bar{\Sigma }\cap \partial \Omega {\setminus } \partial ^* F_\lambda \), then \(x_0\) is not a regular point of the function \(u^* \in BV(\mathbb {R}^n)\), i.e. \(u^*\) is discontinuous at \(x_0\), which is not of jump type (see §4.4 in [16] for a precise definition of regular point of BV functions). Since the set of points which are not regular points of \(u^*\) has \((n-1)-\)dimensional measure zero (see §4.5 [16]), for almost every \(\lambda \in \Lambda \) and \({\mathcal {H}}^{n-2}\)-a.e. \(x_0 \in \bar{\Sigma }\cap \partial \Omega \), \(x_0 \in \partial ^* F_\lambda \cap \partial \Omega \). Thus by Lemma 3.4 we conclude that \(u(\Sigma )=\{\lambda \}\). \(\square \)

It is now straightforward to deduce uniqueness from the results established above. To make the argument rigorous it helps to work with super level sets of the solutions as in [6, 17]. Note however that we do not rely on maximum principles for minimum surfaces that are at the core of the uniqueness proofs in [6, 17].

Proof of Theorem 1.2

First we prove that \(u^*=u\) a.e. in \(\Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}})\). Suppose this is not true, then without loss of generality we may assume that there exists \(\alpha >0\) such that

$$\begin{aligned} {\mathcal {H}}^{n}(N)>0, \end{aligned}$$

where

$$\begin{aligned} N:=\{x\in \Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}}): \ \ u^*(x)\ge u(x)+\alpha \}, \end{aligned}$$

because otherwise the function f in (7) can be replaced by \(-f\). Let

$$\begin{aligned} \lambda ^* =\sup \left\{ \lambda : \ \ {\mathcal {H}}^{n} (\{x\in \Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}}): \ \ u(x)\ge \lambda \} \cap N )\ge \frac{{\mathcal {H}}^{n}(N)}{2}\right\} . \end{aligned}$$

Since \(u\in L^1(\Omega {\setminus } \bar{{\mathcal {I}}})\), \(\lambda ^* <\infty \). For \(0<\beta <1\) define

$$\begin{aligned} E_1= \{x\in \Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}}) : \ \ u^*(x)\ge \lambda ^*+ (1-\beta ) \alpha \}. \end{aligned}$$

By Lemma 3.2 and Proposition 3.1 there exists \(0<\beta <1\) such that \(\lambda ^*+(1-\beta )\alpha \in \Lambda \). Also it follows from the definition of \(\lambda ^*\) that \({\mathcal {H}}^{n}(K)>0\), where

$$\begin{aligned} K:=\{x\in \Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}}): \ \ \lambda ^*-\beta \alpha< u(x) < \lambda ^* \} \cap N. \end{aligned}$$

Now let \(E_2=\{x\in \Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}}): \ \ u(x)\ge \lambda ^* \}\). It is easy to see that \(K \subset E_1 {\setminus } \bar{E}_2 \subset \Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}})\). On the other hand by Remark 3.3 we may assume that \(E_1\) is open and hence \(E_1 {\setminus } \bar{E}_2\) is a non-empty open set. Also

$$\begin{aligned} \partial (E_1{\setminus } \bar{E}_2) \subset \left( \partial E_1 \cap \overline{E_2^c}\right) \cup (E_1 \cap \partial E_2) \end{aligned}$$

and in particular, \(\partial (E_1{\setminus } \bar{E}_2) \subset \partial E_1 \cup \partial E_2\). Notice that \(\partial (E_1{\setminus } \bar{E}_2) \not \subset \partial E_2\), because otherwise \(u=\lambda ^*\) in \(E_1{\setminus } \bar{E}_2\) which is in contradiction with the assumption \(E_1{\setminus } \bar{E}_2 \subset (\Omega {\setminus } {\mathcal {Z}})\). Let

$$\begin{aligned} x_0 \in \partial (E_1{\setminus } \bar{E}_2). \end{aligned}$$

Then \(x_0 \in \partial E_1 \cap \overline{E_2^c}\). By Proposition 3.2 we have

$$\begin{aligned} u(x_0)\in u(\partial E_1)=\{ \lambda ^*+(1-\beta )\alpha \}. \end{aligned}$$
(28)

On the other hand

$$\begin{aligned} u(x_0) \in u(\overline{E_2^c}) \subset (-\infty , \lambda ^*] \end{aligned}$$

which is in contradiction with (28). Hence \(u^*=u\) a.e. in \(\Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}})\).

To finish the proof let \(\Sigma \) be a connected component of \({\mathcal {Z}}\). Since, \(\hbox {int} (\overline{u({\mathcal {Z}})})=\emptyset \), u is continuous, \(u=u^*\) in \(\Omega {\setminus } ({\mathcal {Z}}\cup \bar{{\mathcal {I}}})\), and \(u^*\) minimizes (7), \(u=u^*\) a.e. in \(\Sigma \). The proof is now complete. \(\square \)

Remark 3.5

Note that in domension \(n=2\), if the number of components of \({\mathcal {I}}\) is bigger than one, then there may exists level curves going from one component to the other, and not intersecting \(\partial \Omega \). So the uniqueness argument fails. In higher dimensions this can not happen.