1 Introduction

In this paper we study an overdetermined problem for domains in a cone. This topic shares similarities with the question of characterising constant mean curvature hypersurfaces inside a cone (see [22, 23]) and hence with the isoperimetric problem. Thus we will also show some results for it.

Let D be a smooth domain on the unit sphere \({\mathbb {S}}^{N-1}\) and let \(\Sigma _D\) be the cone spanned by D, namely

$$\begin{aligned} \Sigma _D:=\{x \in {\mathbb {R}}^{N}; \ x=s q,\ q\in D,\ s\in (0,+\infty )\}. \end{aligned}$$
(1.1)

For a domain \(\Omega \subset \Sigma _D\) we set

$$\begin{aligned} \Gamma _\Omega :=\partial \Omega \cap \Sigma _D, \ \Gamma _{1,\Omega }:=\partial \Omega \cap \partial \Sigma _D, \end{aligned}$$

and assume that \({\mathcal {H}}_{N-1}(\Gamma _{1,\Omega })>0\), where \({\mathcal {H}}_{N-1}(\cdot )\) denotes the \((N-1)\)-dimensional Hausdorff measure. The set \(\Gamma _\Omega \) is usually called the relative (to \(\Sigma _D\)) boundary of \(\Omega \).

We consider the overdetermined mixed boundary value problem

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 1 &{} \hbox { in}\ \Omega ,\\ u = 0 &{} \hbox { on}\ \Gamma _\Omega ,\\ \frac{\partial u}{\partial \nu } = 0 &{} \hbox { on } \Gamma _{1,\Omega }\setminus \{0\},\\ \frac{\partial u}{\partial \nu } = -c<0 &{} \hbox { on } \Gamma _\Omega \end{array}\right. } \end{aligned}$$
(1.2)

for a constant \(c>0\), where \(\nu \) is the exterior unit normal. If \(\Gamma _\Omega \) is not smooth then the constant normal derivative condition is understood to hold on the regular part of \(\Gamma _\Omega \).

The overdetermined problem (1.2) arises naturally in the study of critical points of a relative torsional energy of subdomains of the cone \(\Sigma _D\) subject to a fixed volume contraint. Indeed, for any domain \(\Omega \), as in (1.2), let us consider the torsion problem with mixed boundary conditions

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 1 &{} \text {in} \Omega ,\\ u = 0 &{} \text {on }\Gamma _\Omega ,\\ \frac{\partial u}{\partial \nu } = 0 &{} \text {on }\Gamma _{1,\Omega }\setminus \{0\}. \end{array}\right. } \end{aligned}$$
(1.3)

It is easy to see that (1.3) has a unique weak solution \(u_\Omega \) in the Sobolev space \(H_0^1(\Omega ; \Sigma _D)\) (see Sect. 6 or [9]), which is obtained by minimizing the functional

$$\begin{aligned} J(v):= \frac{1}{2}\int _\Omega |\nabla v|^2 \ \mathrm{d}x - \int _\Omega v \ \mathrm{d}x. \end{aligned}$$
(1.4)

We then define the value

$$\begin{aligned} {\mathcal {E}}(\Omega ; \Sigma _D):=J(u_\Omega )=-\frac{1}{2} \int _\Omega |\nabla u_\Omega |^2 \ \mathrm{d}x=-\frac{1}{2}\int _\Omega u_\Omega \ \mathrm{d}x, \end{aligned}$$
(1.5)

and we call it the torsional energy of \(\Omega \) in \(\Sigma _D\). Note that the second and third equality in (1.5) hold since \(u_\Omega \) is a weak solution of (1.3). By definition, the domain-dependent functional \(\Omega \mapsto {\mathcal {E}}(\Omega ; \Sigma _D)\) represents a relative version of the classical torsional energy functional usually defined using the solution of the analogous Dirichlet problem.

Using domain derivative techniques, as for other similar problems in shape-optimization theory it can be proved that the critical points of the functional \({\mathcal {E}}(\Omega ; \Sigma _D)\) with respect to volume-preserving deformations which leave the cone invariant, correspond to domains \(\Omega \) for which \(\frac{\partial u_\Omega }{\partial \nu }\) is constant on \(\Gamma _\Omega \), i.e. \(u_\Omega \) satisfies the overdetermined problem (1.2) (see [23, Proposition 4.3] if \(\Gamma _\Omega \) is smooth and \(u_\Omega \) has some Sobolev regularity, or Proposition 7.4 in the present paper in the nonsmooth case).

In this paper we intend to study the existence and the properties of domains for which a solution of (1.2) exists. It is easy to see that for any spherical sector \(\Omega _{D,R}:=B_R\cap \Sigma _D\), where \(B_R=B_R(0)\) is the ball with radius \(R>0\) centered at the origin (which is the vertex of the cone), the radial function

$$\begin{aligned} u(x)=\frac{N^2 c^2-|x|^2}{2N} \end{aligned}$$
(1.6)

is a solution of (1.2) for \(\Omega =\Omega _{D,R}\). Therefore the question is whether the spherical sectors \(\Omega _{D,R}\) are the only domains for which (1.2) admits a solution. In the case of convex cones the answer was provided in [22], obtaining the following result (see [22, Theorem 1.1]):

Theorem 1.1

If \(\Sigma _D\) is convex, \(\Gamma _\Omega \) is smooth and u is a classical solution of (1.2) such that \(u \in W^{1,\infty }(\Omega )\cap W^{2,2}(\Omega )\), then

$$\begin{aligned} \Omega =\Sigma _D\cap B_R(P_0), \end{aligned}$$

where \(B_R(P_0)\) is the ball centered at \(P_0\) with radius \(R=Nc\), and either \(P_0=0\), i.e. \(\Omega =\Omega _{D,R}\), or \(P_0\in \partial \Sigma _D\) and \(\Omega \) is a half-ball lying on a flat part of \(\partial \Sigma _D\).

Hence, if \(\Sigma _D\) is a convex cone, not flat anywhere, then the radial domains \(\Omega _{D,R}\) are the only domains admitting solutions of (1.2). Let us observe that the assumption \(u \in W^{1,\infty }(\Omega )\cap W^{2,2}(\Omega )\) can be seen as a “gluing condition". Indeed it is automatically satisfied whenever \(\Gamma _\Omega \) and \(\partial \Sigma _D\) intersect orthogonally (see [22, Sect. 6]).

In the context of the variational formulation of problem (1.2) described above, the result of Theorem 1.1 gives a characterization of the smooth critical points of \({\mathcal {E}}(\Omega ; \Sigma _D)\), restricted to the class of subdomains of fixed volume, in the case of convex cones. In particular any local minimizer of \({\mathcal {E}}(\Omega ; \Sigma _D)\) with a volume constraint is a spherical sector. Actually, using symmetrization methods in cones [19, 24] it can be proved (see [23]) that this holds in a more general class of cones which are the ones having an isoperimetric property.

In contrast, the case of nonconvex cones is largely unexplored, which is the main motivation of the present paper. The variational formulation of the overdetermined problem suggests that to look for nonradial domains for which there exists a solution of (1.2) is equivalent to look for nonradial critical points of \({\mathcal {E}}(\Omega ; \Sigma _D)\) under a volume constraint. In particular, if there are cones for which a minimizer of \({\mathcal {E}}(\Omega ; \Sigma _D)\) (fixing the volume) exists and if we are able to show that it is not the spherical sector then we achieve our goal. This is the content of our first main result.

Let us denote by \(\uplambda _1(D)\) the first nontrivial eigenvalue of the Laplace-Beltrami operator \(-\Delta _{{\mathbb {S}}^{N-1}}\) on D with zero Neumann boundary condition.

Theorem 1.2

If D is a smooth domain of \({\mathbb {S}}^{N-1}\) such that

$$\begin{aligned} \uplambda _1(D)<N-1 \ \ \hbox {and} \ \ {\mathcal {H}}_{N-1}(D)<{\mathcal {H}}_{N-1}({\mathbb {S}}^{N-1}_+), \end{aligned}$$
(1.7)

where \({\mathbb {S}}^{N-1}_+\) is a half unit sphere, then there exists a bounded domain \(\Omega ^*\) which is a minimizer for \({\mathcal {E}}(\Omega ; \Sigma _D)\) with a fixed volume, but \(\Omega ^*\) is not a spherical sector \(\Omega _{D,R}\), for \(R>0\).

Moreover there exists a critical dimension \(d^*\) which can be either 5, 6 or 7, such that for the relative boundary \(\Gamma _{\Omega ^*}\) it holds that

  1. (i)

    \(\Gamma _{\Omega ^*}\) is smooth if \(N<d^*\);

  2. (ii)

    \(\Gamma _{\Omega ^*}\) can have countable isolated singularities if \(N=d^*\);

  3. (iii)

    \(\Gamma _{\Omega ^*}\) can have a singular set of dimension \(N-d^*\), if \(N>d^*\).

In addition on the regular part of \(\Gamma _{\Omega ^*}\) the normal derivative \(\frac{\partial u_{\Omega ^*}}{\partial \nu }\) is constant, where \(u_{\Omega ^*}\) is the torsion function of \(\Omega ^*\).

The condition \(\uplambda _1(D)<N-1\) in (1.7) is the one which ensures that a spherical sector \(\Omega _{D,R}\) cannot be a local minimizer for \({\mathcal {E}}(\Omega ; \Sigma _D)\) among the class of smooth subdomains of \(\Sigma _D\) with fixed volume, because it implies that it is not a stable critical point with respect to volume-preserving deformations (see Theorem 5.1). To prove this, we restrict the torsional energy functional to the class of strictly star-shaped sets \(\Omega \) in \(\Sigma _D\) with fixed volume \(c>0\), and we show the instability of the spherical sector \(\Omega _{D,R}\) with \(|\Omega _{D,R}|=c\) within this class. The reason to consider strictly star-shaped domains is that the relative boundary \(\Gamma _\Omega \) of a strictly star-shaped set is a radial graph of a function \(\varphi \) on D. This allows to study \({\mathcal {E}}(\Omega ; \Sigma _D)\) as a functional on \(\varphi \in C^2({\overline{D}})\).

On the other hand, the condition \({\mathcal {H}}_{N-1}(D)<{\mathcal {H}}_{N-1}({\mathbb {S}}^{N-1}_+)\) is the one which allows to prove the existence of a minimizer for \({\mathcal {E}}(\Omega ; \Sigma _D)\) (see Theorem 6.8 and Corollary 6.9). In the Appendix we give examples of domains D on \({\mathbb {S}}^{N-1}\) satisfying both conditions in (1.7).

Let us observe that, since \(\Sigma _D\) is not bounded, the existence of a minimizer for \({\mathcal {E}}(\Omega ; \Sigma _D)\) is not obvious. To prove Theorem 1.2 we use the concentration-compactness principle of P. L. Lions (see [17]). It was first used in shape-optimization Dirichlet problems in [6]. Having mixed boundary conditions, we cannot make use of the same proof as in [6]. We also stress that, as the cone \(\Sigma _D\) is not convex and since we do not have any information on the contact angle between \(\Sigma _D\) and \(\Gamma _{\Omega ^*}\), some care is needed to prove that the normal derivative \(\frac{\partial u_{\Omega ^*}}{\partial \nu }\) of the torsion function \(u_{\Omega ^*}\) is constant on the regular part of \(\Gamma _{\Omega ^*}\) (see Proposition 7.4). Finally, the regularity statements follow from the results of [11, 15] and [27].

As announced we also consider the isoperimetric problem in the cone to get a analogous nonradiality result using the same strategy.

The isoperimetric problem in the cone consists in minimizing the relative perimeter \({\mathcal {P}}(E; \Sigma _D)\) among all possibile finite relative perimeter sets E contained in the cone \(\Sigma _D\), with a fixed volume. It was proved in [18], and later in [10, 13, 25], that if \(\Sigma _D\) is a convex cone then the only minimizer of \({\mathcal {P}}(E; \Sigma _D)\) with a fixed volume are the spherical sectors \(\Omega _{D,R}\). This holds also in “almost" convex cones as shown in [2] (see also [23]). If the cone is not convex, a counterexample is given in [18].

Here we show that under the same conditions (1.7), a minimizer of \({\mathcal {P}}(E;\Sigma _D)\), exists, but is not the spherical sector \(\Omega _{D,R}\). Thus we have

Theorem 1.3

If D is a smooth domain of \({\mathbb {S}}^{N-1}\) such that (1.7) holds then there exists a bounded set of finite perimeter \(E^*\) inside \(\Sigma _D\) which minimizes \({\mathcal {P}}(E;\Sigma _D)\) for any fixed volume and \(E^*\) is not a spherical sector \(\Omega _{D,R}\), \(R>0\). Moreover for the relative boundary \(\Gamma _{E^*}\) it holds that

  1. (i)

    \(\Gamma _{E^*}\) can have a closed singular set \({\widetilde{\Gamma }}_{E^*}\) of Hausdorff dimension less than or equal to \(N-7\);

  2. (ii)

    \(\Gamma _{E^*}\setminus {\widetilde{\Gamma }}_{E^*}\) is a smooth embedded hypersurface with constant mean curvature;

  3. (iii)

    if \(x\in \overline{\Gamma _{E^*}\setminus {\widetilde{\Gamma }}_{E^*}} \cap \partial \Sigma _D\) then \(\Gamma _{E^*}\setminus {\widetilde{\Gamma }}_{E^*}\) is a smooth CMC embedded hypersurface with boundary in a neighborhood of x and meets \(\partial \Sigma _D\) orthogonally.

As for Theorem 1.2, the condition \(\uplambda _1(D)<N-1\) is the one which ensures that \(\Omega _{D,R}\) cannot be a local minimizer (see Theorem 8.3) and to prove this we again work in the class of smooth star-shaped sets. Instead the existence follows by results obtained in [25], while the regularity of minimizers derives from classical results for isoperimetric problems.

As a consequence of Theorem 1.3 we get that whenever (1.7) holds there exists a CMC hypersurface in the cone, namely \(\Gamma _{E^*}\), intersecting \(\partial \Sigma _D\) orthogonally, which is not a spherical cap centered at the vertex of the cone. It is important to notice that \(\Gamma _{E^*}\) cannot be a smooth radial graph. Indeed, by [22, Theorem 1.3] and [23, Theorem 1.1], we know that if \(\Gamma _{E^*}\) was a CMC radial graph intersecting \(\partial \Sigma _D\) orthogonally then \(E^*\) would be a spherical sector \(\Omega _{D,R}\), and this holds in any cone without requiring convexity hypotheses. It would be very interesting to understand what kind of CMC hypersurface \(\Gamma _{E^*}\) could be.

Finally we observe that, from our results and [18, Theorem 1.1] (or [22, Theorem 1.1]), we easily recover the inequality \(\uplambda _1(D)\geqq N-1\) whenever D is convex. This was proved in [12, Theorem 4.3] (see also [1, Theorem 4.1]).

The paper is organized as follows: in Sect. 2 we provide some geometric preliminaries. In Sect. 3 we study the torsional energy functional \({\mathcal {E}}(\Omega ; \Sigma _D)\) on strictly star-shaped domains in the cone, while in Sect. 4 we derive the formulas for the first and second variations of \({\mathcal {E}}(\Omega ; \Sigma _D)\) when the volume is fixed. In Sect. 5 we prove that the first condition in (1.7) allows to prove that the spherical sector is not a local minimizer for \({\mathcal {E}}(\Omega ; \Sigma _D)\). The long Sect. 6 is devoted to study the question of the existence of minimizers of \({\mathcal {E}}(\Omega ; \Sigma _D)\) with a volume constraint. Their properties are described in Sect. 7 where the proof of Theorem 1.2 is deduced. Finally in Sect. 8 we study the isoperimetric problem and prove Theorem 1.3. In the Appendix we give examples of nonconvex domains satisying the condition (1.7).

2 Some Preliminaries

In this section we fix some notation and we collect, for the reader’s convenience, some definitions and known facts from Riemannian Geometry that will be used throughout the paper.

Given a smooth manifold M, we denote by \(T_pM\) the tangent space at \(p\in M\), by \({\mathcal {T}}(M)\) the space of tangent vector fields on M and by TM the tangent bundle.

We denote by \(\langle .,\rangle \) or \(\varvec{\cdot }\) the standard scalar product in \({\mathbb {R}}^{N}\), by \(|\cdot |\) the Euclidean norm, and by \(\nabla ^0\) the flat connection of \({\mathbb {R}}^{N}\). In the special case \(M=D\), where \(D\subset {\mathbb {S}}^{N-1}\) is a domain of the unit sphere in \({\mathbb {R}}^N\), we denote by \(\nabla \) the induced Levi-Civita connection on D,namely

$$\begin{aligned} \nabla _X Y:=(\nabla _{X}^0Y)^\top ,\ \text {for any} X,Y \in {{\mathcal {T}}}(D), \end{aligned}$$

where \(\top :T{\mathbb {R}}^{N}\rightarrow TD\) is the orthogonal projection. If we further assume that D is a proper and smooth domain of \({\mathbb {S}}^{N-1}\) it will be always understood that D is considered as a submanifold with boundary, equipped with the induced Riemannian metric.

If \(\varphi :D\rightarrow {\mathbb {R}}\) is a smooth function, we adopt, respectively, the notations \(\mathrm{d}\varphi \), \(\nabla \varphi \), to indicate the differential and the gradient of \(\varphi \), which is the only vector field on D such that

$$\begin{aligned} \mathrm{d}\varphi [X] = \langle X, \nabla \varphi \rangle , \ \ \hbox {for any} \ X \in {\mathcal {T}}(D). \end{aligned}$$

We will also use sometimes the notation \(\nabla _{{\mathbb {S}}^{N-1}} \varphi \) instead of \(\nabla \varphi \) to make a distinction with respect to the usual gradient of real valued functions defined in open subsets of \({\mathbb {R}}^N\). The second covariant derivative of \(\varphi \) is defined as

$$\begin{aligned} \nabla _{X,Y} \varphi := \nabla _X \nabla _Y \varphi - \nabla _{{\nabla _X Y}} \varphi , \ \ \hbox {for any} \ X,Y \in {\mathcal {T}}(D), \end{aligned}$$
(2.1)

and the Hessian of \(\varphi \), denoted by \(\nabla ^2 \varphi \) or by \(D^2 \varphi \), is the symmetric 2-tensor given by

$$\begin{aligned} \nabla ^2 \varphi \ (X,Y):=\nabla _{X,Y} \varphi , \ \ \hbox {for any} \ X,Y \in {\mathcal {T}}(D). \end{aligned}$$

The Laplacian of \(\varphi \), denoted by \(\Delta \varphi \), is the trace of the Hessian. Again, when there is a chance of confusion with the standard Laplacian we will use the notation \(\Delta _{{\mathbb {S}}^{N-1}} \varphi \) instead of \(\Delta \varphi \).

Let \(\{e_1,\ldots ,e_{N-1}\}\) be a local orthonormal frame field for D. For any \(i,j\in \{1,\ldots ,N-1\}\) we define the connection form \(\omega _{ij}\) as

$$\begin{aligned} \omega _{ij}(X):=\langle \nabla _X e_j , e_i\rangle ,\ X \in {{\mathcal {T}}}(D). \end{aligned}$$
(2.2)

We recall that the connection forms are skew symmetric and in terms of the \(\omega _{ij}\)’s we can write

$$\begin{aligned} \nabla _{e_i}e_j = \sum _{k=1}^{N-1} \omega _{kj}(e_i) e_k. \end{aligned}$$
(2.3)

We denote by \(\varphi _i\) the covariant derivative \(\nabla _{e_i}\varphi \), and we recall that, by definition, \(\nabla _{e_i} \varphi =\mathrm{d}\varphi [e_i]\). It is easy to check that the gradient of \(\varphi \) can be written as

$$\begin{aligned} \nabla \varphi = \sum _{i=1}^{N-1} \varphi _i e_i. \end{aligned}$$

Finally, taking \(X=e_i\), \(Y=e_j\) in (2.1) and using (2.2) we have

$$\begin{aligned} \nabla _{e_i,e_j} \varphi = \nabla _{e_i}\varphi _j -\sum _{k=1}^{N-1}\omega _{kj}(e_i) \varphi _k. \end{aligned}$$
(2.4)

From now on we will use the notation \(\varphi _{ij}\) to denote \(\nabla _{e_i,e_j} \varphi \). In particular the Laplacian of \(\varphi \) can be written as \( \Delta \varphi = \sum _{i=1}^{N-1} \varphi _{ii}\).

Now we consider the special case of radial graphs.

Definition 2.1

Let \(D \subset {\mathbb {S}}^{N-1}\) be a domain and let \(\varphi \in C^2(D)\). We denote by \(\Gamma _\varphi \) the associated radial graph to \(\varphi \), namely

$$\begin{aligned} \Gamma _\varphi :=\{x\in {\mathbb {R}}^N; \ x=e^{\varphi (q)}q, \ q\in D\}. \end{aligned}$$

Clearly \(\Gamma _\varphi \) is a \((N-1)\)-dimensional manifold (of class \(C^2\)). We consider the map \({\mathcal {Y}}:D \rightarrow \Gamma _\varphi \) defined by

$$\begin{aligned} {\mathcal {Y}}(q):=e^{\varphi (q)}q, \ \ \ q\in D. \end{aligned}$$
(2.5)

For any fixed \(q\in D\), let \(\gamma _i:(-\delta ,\delta ) \rightarrow D\), \(\gamma _i=\gamma _i(t)\) be a curve contained in D and such that \(\gamma _i(0)=q\), \(\gamma ^\prime _i(0)=e_i(q)\), for \(i=1,\ldots ,N-1\). Since

$$\begin{aligned} \left. \frac{d({\mathcal {Y}}\circ \gamma _i)}{\mathrm{d}t}\right| _{t=0} = e^\varphi (\varphi _i q + e_i) \end{aligned}$$
(2.6)

then a local basis for \(T_{{\mathcal {Y}}(q)}\Gamma _\varphi \) is given by

$$\begin{aligned} E_i(q)= e^\varphi (e_i + \varphi _i q), \ \ i=1,\ldots ,N-1, \end{aligned}$$

and the components of the induced metric are

$$\begin{aligned} g_{ij}=\langle E_i, E_j\rangle =e^{2\varphi } (\langle e_i, e_j \rangle + \varphi _i \varphi _j \langle q, q\rangle )=e^{2\varphi } (\delta _{ij} + \varphi _i\varphi _j). \end{aligned}$$

We denote by \(\nu ({\mathcal {Y}}(q))\) the exterior unit normal at \({\mathcal {Y}}(q)\in \Gamma _\varphi \). It is easy to check that

$$\begin{aligned} \nu ({\mathcal {Y}}(q))= \frac{q- \sum _{i=1}^{N-1} \varphi _i e_i }{(1+|\nabla \varphi |^2)^{1/2}}= \frac{q- \nabla \varphi }{(1+|\nabla \varphi |^2)^{1/2}}. \end{aligned}$$
(2.7)

In addition by direct computation we see that the coefficients of the second fundamental form are

$$\begin{aligned} {\mathbf {I}}{\mathbf {I}}_{ij} =\frac{e^{\varphi }\left( \delta _{ij} + \varphi _i\varphi _j -\varphi _{ij}\right) }{(1+|\nabla \varphi |^2)^{1/2}}, \end{aligned}$$

for any \(i,j=1,\ldots ,N-1\) (see [20] or [4] for more details).

Finally, since the mean curvature at \({\mathcal {Y}}(q) \in \Gamma _\varphi \) is given by

$$\begin{aligned} N H({\mathcal {Y}}(q))= \sum _{i,j=1}^{N-1} g^{ij} {\mathbf {I}}{\mathbf {I}}_{ij}, \end{aligned}$$

where \((g^{ij})\) is the inverse matrix of \((g_{ij})\), namely

$$\begin{aligned} g^{ij}=e^{-2\varphi }\left( \delta _{ij}-\frac{\varphi _i\varphi _j}{1+|\nabla \varphi |^2}\right) , \end{aligned}$$
(2.8)

then, by a straightforward computation we see that \(\varphi \) must satisfy the following equation

$$\begin{aligned} \!\!\!\!\!\!\!\!\!\!\!\!\sum _{i,j=1}^{N-1} \left( (1+|\nabla \varphi |^2)\delta _{ij} - \varphi _i \varphi _j)\right) \varphi _{ij}&\!=&(N-1)(1+|\nabla \varphi |^2)\nonumber \\&- (N-1) e^\varphi (1+|\nabla \varphi |^2)^{3/2}H({\mathcal {Y}}(q)). \end{aligned}$$
(2.9)

Writing (2.9) in divergence form we obtain the well known equation for radial graphs of prescribed mean curvature (see [20] or [26])

$$\begin{aligned}&\!\!\!\!\!\!\!\!\!\!\!\!-\mathrm {div}_{{\mathbb {S}}^{N-1}}\left( \frac{\nabla \varphi }{\sqrt{1+|\nabla \varphi |^{2}}}\right) +\frac{N-1}{\sqrt{1+|\nabla \varphi |^{2}}} =(N-1) e^{\varphi }H(e^{\varphi }q)\quad \text{ in } D. \end{aligned}$$
(2.10)

3 Torsional Energy for Domains in Cones

In this section we define and study the torsional energy for smooth domains in cones and then we focus on the class of strictly star-shaped domains.

Let D be a smooth proper domain of \({\mathbb {S}}^{N-1}\) and let \(\Sigma _D\) be the cone spanned by D. For a bounded domain \(\Omega \subset \Sigma _D\) we set:

$$\begin{aligned} \Gamma _\Omega :=\partial \Omega \subset \Sigma _D, \ \ \ \Gamma _{1,\Omega }:=\partial \Omega \cap \partial \Sigma _D, \end{aligned}$$

and assume that \({\mathcal {H}}_{N-1}(\Gamma _{1,\Omega })>0\) and that \(\Gamma _\Omega \) is a smooth \((N-1)\)-dimensional manifold whose boundary \(\partial \Gamma _\Omega =\partial \Gamma _{1,\Omega }\subset \partial \Sigma _D\setminus \{0\}\) is a smooth \((N-2)\)-dimensional manifold. The set \(\Gamma _\Omega \) is often called the relative (to \(\Sigma _D\)) boundary of \(\Omega \).

We consider the following mixed boundary value problem:

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 1 &{} \text {in }\Omega ,\\ u = 0 &{} \text {on}\ \Gamma _\Omega ,\\ \frac{\partial u}{\partial \nu } = 0 &{} \text {on}\ \Gamma _{1,\Omega }\setminus \{0\}. \end{array}\right. } \end{aligned}$$
(3.1)

It is easy to see that (3.1) admits a unique weak solution \(u_\Omega \) in the space \(H_0^{1}(\Omega \cup \Gamma _{1,\Omega })\) which is the Sobolev space of functions in \(H^{1}(\Omega )\) whose trace vanishes on \(\Gamma _\Omega \). Indeed \(u_\Omega \) is the only minimizer of the functional

$$\begin{aligned} J(v):= \frac{1}{2}\int _\Omega |\nabla v|^2 \ \mathrm{d}x - \int _\Omega v \ \mathrm{d}x \end{aligned}$$

in the space \(H_0^{1}(\Omega \cup \Gamma _{1,\Omega })\) and we remark that \(u_\Omega >0\) a.e. in \(\Omega \), by the maximum principle (we refer to [22, 23] for more details).

Usually, the function \(u_\Omega \) is called torsion function of \(\Omega \) and its energy \(J(u_\Omega )\) represents the torsional energy of the domain \(\Omega \). This allows to consider the functional

$$\begin{aligned} {\mathcal {E}}(\Omega ; \Sigma _D)=J(u_\Omega ) \end{aligned}$$

which is defined on the domains contained in \(\Sigma _D\).

From the weak formulation of (3.1) we have

$$\begin{aligned} \int _\Omega |\nabla u_\Omega |^2 \ \mathrm{d}x = \int _\Omega u_\Omega \ \mathrm{d}x, \end{aligned}$$

which implies that

$$\begin{aligned} {\mathcal {E}}(\Omega ;\Sigma _D)=-\frac{1}{2} \int _\Omega |\nabla u_\Omega |^2 \ \mathrm{d}x = -\frac{1}{2}\int _\Omega u_\Omega \ \mathrm{d}x. \end{aligned}$$
(3.2)

Now we focus on the special case when \(\Omega \) is strictly star-shaped with respect to the origin which is the vertex of the cone \(\Sigma _D\). Thus we consider the relative boundary \(\Gamma _\Omega \) as the radial graph in \(\Sigma _D\) of a function \(\varphi \in C^2({{\overline{D}}}, {\mathbb {R}})\) as defined in Sect. 2. Therefore we denote \(\Omega \) by \(\Omega _\varphi \) which can be described as:

$$\begin{aligned} \Omega _\varphi :=\{x\in \Sigma _D; \ x=s q, \ 0<s<e^{\varphi (q)}, \ q\in D\}. \end{aligned}$$
(3.3)

We restrict the torsional energy functional \({\mathcal {E}}\) to this class of domains and we denote it by \({\mathscr {E}}\), i.e. we set

$$\begin{aligned} {\mathscr {E}}(\varphi ):={\mathcal {E}}(\Omega _\varphi ; \Sigma _D). \end{aligned}$$

We observe that \({\mathscr {E}}\) is a functional defined on \(C^2({\overline{D}},{\mathbb {R}})\) and we compute its first and second derivatives. To this aim we point out that taking variations of \(\varphi \) in \(C^2({\overline{D}},{\mathbb {R}})\) corresponds to taking variations of \(\Omega _\varphi \) in the class of strictly star-shaped domains (of class \(C^2\)).

Let us set for simplicity

$$\begin{aligned} \Gamma _\varphi :=\Gamma _{\Omega _\varphi }, \ \ \ \Gamma _{1,\varphi }:=\Gamma _{1,\Omega _\varphi }. \end{aligned}$$

If \(v \in C^2({{\overline{D}}}, {\mathbb {R}})\) and \(t \in (-\delta , \delta )\), where \(\delta >0\) is a fixed number, we consider the domain variations \(\Omega _{\varphi +tv} \subset \Sigma _D\), \({t\in (-\delta ,\delta )}\). Let \(\xi :(-\delta ,\delta )\times \Sigma _D \rightarrow \Sigma _D\) be the map defined by

$$\begin{aligned} \xi (t,x)=e^{tv\left( \frac{x}{|x|}\right) }x. \end{aligned}$$

It is elementary to check that, for a fixed \(t\in (-\delta ,\delta )\) the restriction

$$\begin{aligned} \xi |_{\Omega _\varphi }(t,\cdot ):\Omega _\varphi \rightarrow \Omega _{\varphi +tv} \end{aligned}$$
(3.4)

is a diffeomorphism whose inverse \(\left( \xi |_{\Omega _\varphi }\right) ^{-1}:\Omega _{\varphi +tv} \rightarrow \Omega _\varphi \) is given by

$$\begin{aligned} \left( \xi |_{\Omega _\varphi }\right) ^{-1}(x)=e^{-tv(\frac{x}{|x|})}x=\xi (-t,x). \end{aligned}$$

Moreover by definition we have \(\xi (t,x) \in \partial \Sigma _D\setminus \{0\}\) for all \((t,x) \in (-\delta ,\delta ) \times \partial \Sigma _D\setminus \{0\}\). In particular \(\xi \) is the flow associated to the vector field V on \(\Sigma _D\) given by

$$\begin{aligned} V(x):=v\left( \frac{x}{|x|}\right) x, \end{aligned}$$
(3.5)

since \(\xi (0,x)=x\) and \(\frac{\mathrm{d}\xi }{\mathrm{d}t}(t,x)=e^{tv\left( \frac{x}{|x|}\right) } v\left( \frac{x}{|x|}\right) x=V(\xi (t,x))\), and \((\Omega _{\varphi +tv})_{t\in (-\delta ,\delta )}\) is a deformation of \(\Omega _\varphi \) associated to the vector field V (see [16, Definition 1.1]). We now compute the derivative of \({\mathscr {E}}\) with respect to a variation \(v \in C^2({{\overline{D}}},{\mathbb {R}})\).

Lemma 3.1

Let \(\varphi \in C^2({{\overline{D}}},{\mathbb {R}})\) and assume that \(u_{\Omega _\varphi }\in W^{1,\infty }(\Omega _\varphi )\cap W^{2,2}(\Omega _\varphi )\). Then, for any \(v \in C^2({{\overline{D}}},{\mathbb {R}})\), it holds that

$$\begin{aligned} {\mathscr {E}}^\prime (\varphi )[v]=-\frac{1}{2} \int _D e^{N\varphi }\ v \left( \frac{\partial u_{\Omega _\varphi } }{\partial \nu }\left( e^{\varphi }q\right) \right) ^2 \ \mathrm{d}\sigma , \end{aligned}$$

where \(\mathrm{d}\sigma \) is the \((N-1)\)-dimensional area element of \({\mathbb {S}}^{N-1}\).

Proof

Let \(\varphi \in C^2({{\overline{D}}},{\mathbb {R}})\) as in the statement and let \(v \in C^2({{\overline{D}}},{\mathbb {R}})\). By definition we have

$$\begin{aligned} {\mathscr {E}}(\varphi +tv)={\mathcal {E}}(\Omega _{\varphi +tv};\Sigma _D)= -\frac{1}{2}\int _{\Omega _{\varphi +tv}} u_{\Omega _{\varphi +tv}} \ \mathrm{d}x, \end{aligned}$$
(3.6)

where \(u_{\Omega _{\varphi +tv}}\) is the only (positive) weak solution to

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 1 &{} \text {in }{\Omega _{\varphi +tv}},\\ u = 0 &{} \text {on }\ \Gamma _{\varphi +tv},\\ \frac{\partial u}{\partial \nu } = 0 &{} \text {on}\ \Gamma _{1,\varphi +tv}\setminus \{0\}. \end{array}\right. } \end{aligned}$$
(3.7)

Writing (3.6) in polar coordinates we obtain that

$$\begin{aligned} {\mathscr {E}}(\varphi +tv)= -\frac{1}{2}\int _D\int _0^{e^{\varphi +tv}} \rho ^{N-1} u_{\Omega _{\varphi +tv}}(\rho q) \ \mathrm{d}\rho \mathrm{d}\sigma . \end{aligned}$$

Let \({\hat{\Phi }}:(-\delta ,\delta )\rightarrow H^{1}_0(\Omega _\varphi \cup \Gamma _{1,\varphi })\) be the map defined by

$$\begin{aligned} {\hat{\Phi }}(t):={\hat{u}}_t, \end{aligned}$$

where \({\hat{u}}_t:=u_{\Omega _{\varphi +tv}}\circ \xi (t,\cdot )|_{\Omega _\varphi } \in H^{1}_0(\Omega _\varphi \cup \Gamma _{1,\varphi })\), \(\xi |_{\Omega _\varphi }(t,\cdot ):\Omega _\varphi \rightarrow \Omega _{\varphi +tv}\) is the diffeomorphism given by (3.4). From the proof of [23, Proposition 4.3] we know that \({\hat{\Phi }}\) is differentiable and thus we infer that \(u_{\Omega _{\varphi +tv}}\) is differentiable with respect to t. Hence, by the Leibniz integral rule for differentiation of integral functions we get that

$$\begin{aligned} \begin{array}{lll} \displaystyle \frac{\mathrm{d}}{\mathrm{d}t}\left( {\mathscr {E}}(\varphi +tv)\right) &{}=&{}\displaystyle -\frac{1}{2}\int _D e^{(N-1)(\varphi +tv)} e^{\varphi +tv} v\ u_{\Omega _{\varphi +tv}}(e^{\varphi +tv} q) \ \mathrm{d}\sigma \\ &{}&{} \displaystyle -\frac{1}{2}\int _D \int _0^{e^{\varphi +tv}} \rho ^{N-1} \frac{\mathrm{d}}{\mathrm{d}t}\left( u_{\Omega _{\varphi +tv}}\right) (\rho q) \ \mathrm{d}\rho \mathrm{d}\sigma . \end{array} \end{aligned}$$

In view of (3.7) we have \(u_{\Omega _{\varphi +tv}}(e^{\varphi +tv} q)=0\) on D for any \(t\in (-\delta , \delta )\). In particular computing at \(t=0\) we have

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathscr {E}}^\prime (\varphi )[v]= \frac{\mathrm {d}}{\mathrm {d}t}\left. \left( {\mathscr {E}}(\varphi +tv)\right) \right| _{t=0}&{}{}=&{}{}\displaystyle -\frac{1}{2}\int _D \int _0^{e^{\varphi (q)}} \rho ^{N-1} \frac{\mathrm {d}}{\mathrm {d}t}\left. \left( u_{\Omega _{\varphi +tv}}\right) \right| _{t=0}(\rho q) \ \mathrm {d}\rho \mathrm {d}\sigma \\ &{}{}=&{}{}\displaystyle -\frac{1}{2}\int _{\Omega _\varphi } \frac{\mathrm {d}}{\mathrm {d}t}\left. \left( u_{\Omega _{\varphi +tv}}\right) \right| _{t=0} \ \mathrm {d}x. \end{array}\nonumber \\ \end{aligned}$$
(3.8)

Setting \(u^\prime :=\left. \frac{\mathrm{d}}{\mathrm{d}t}\left( u_{\Omega _{\varphi +tv}}\right) \right| _{t=0}\) and arguing as in the proof of [23, Proposition 4.3], where the assumption \(u_{\Omega _\varphi }\in W^{1,\infty }(\Omega _\varphi )\cap W^{2,2}(\Omega _\varphi )\) is used, we infer that \(u^\prime \in H^{1}_0(\Omega _\varphi \cup \Gamma _{1,\varphi })\) satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u^\prime = 0 &{} \text {in}\ \Omega _{\varphi },\\ u^\prime = -\frac{\partial u_{\Omega _\varphi }}{\partial \nu } \langle V,\nu \rangle &{} \text {on }\Gamma _{\varphi },\\ \frac{\partial u^\prime }{\partial \nu } = 0 &{} \text {on}\ \Gamma _{1,\varphi }\setminus \{0\}. \end{array}\right. } \end{aligned}$$
(3.9)

In particular, in view of (3.5) and since \(\Gamma _{\varphi }\) is a radial graph we have \(\nu (x)=\frac{\frac{x}{|x|}-\nabla _{{\mathbb {S}}^{N-1}}\varphi \left( \frac{x}{|x|}\right) }{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi \left( \frac{x}{|x|}\right) |^2}}\), for any \(x\in \Gamma _\varphi \) (see (2.7)), and thus

$$\begin{aligned} \langle V,\nu \rangle =\frac{|x|}{\sqrt{1+\left| \nabla _{{\mathbb {S}}^{N-1}} \varphi \left( \frac{x}{|x|}\right) \right| ^2}} v\left( \frac{x}{|x|}\right) \ \ \hbox { on}\ \Gamma _\varphi . \end{aligned}$$
(3.10)

Rewriting (3.8) in terms of \(u^\prime \), applying Green’s second identity (which holds also in conic domains, since it is a consequence of the divergence theorem, see e.g. [22, Lemma 2.1]) and taking into account (3.1) (with \(\Omega =\Omega _\varphi \)), (3.9) and (3.10) we get that

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathscr {E}}^\prime (\varphi )[v]&{}=&{}\displaystyle -\frac{1}{2}\int _{\Omega _\varphi } u^\prime \ \mathrm{d}x.\\ &{}=&{}\displaystyle \frac{1}{2}\int _{\Omega _\varphi } u^\prime {\Delta u_{\Omega _\varphi }} \ \mathrm{d}x -\frac{1}{2}\int _{\Omega _\varphi } \underbrace{\Delta u^\prime }_{=0} u_{\Omega _\varphi } \ \mathrm{d}x\\ &{}=&{}\displaystyle \frac{1}{2}\int _{\Gamma _\varphi } u^\prime \frac{\partial u_{\Omega _\varphi }}{\partial \nu } \ \mathrm{d}\sigma _{\Gamma _\varphi } + \frac{1}{2}\int _{\Gamma _{1,\varphi }\setminus \{0\}} u^\prime \underbrace{\frac{\partial u_{\Omega _\varphi }}{\partial \nu }}_{=0} \ \mathrm{d}\sigma _{\Gamma _{1,\varphi }\setminus \{0\}}\\ &{}=&{}\displaystyle - \frac{1}{2}\int _{\Gamma _\varphi } \left( \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(x)\right) ^2 \frac{|x|}{\sqrt{1+\left| \nabla _{{\mathbb {S}}^{N-1}} \varphi \left( \frac{x}{|x|}\right) \right| ^2}} v\left( \frac{x}{|x|}\right) \ \mathrm{d}\sigma _{\Gamma _\varphi }, \end{array} \end{aligned}$$
(3.11)

where \(\mathrm{d}\sigma _{\Gamma _\varphi }\), \(\mathrm{d}\sigma _{\Gamma _{1,\varphi }\setminus \{0\}}\) are the \((N-1)\)-dimensional area elements of \(\Gamma _\varphi \), \(\Gamma _{1,\varphi }\setminus \{0\}\), respectively. Finally, writing \(x=e^{\varphi (q)}q\), \(q\in D\), observing that \(\mathrm{d}\sigma _{\Gamma _\varphi }=e^{(N-1)\varphi }\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}}\varphi |^2} \mathrm{d}\sigma \) and \(\frac{x}{|x|}=q\), then from (3.11) we obtain that

$$\begin{aligned} {\mathscr {E}}^\prime (\varphi )[v] = - \frac{1}{2}\int _{D} e^{N\varphi } v\left( \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\right) ^2 \ \mathrm{d}\sigma , \end{aligned}$$

and this completes the proof. \(\square \)

For the second variation of the functional \({\mathscr {E}}\) we have

Lemma 3.2

Let \(\varphi \) be as in Lemma 3.1. Then, for any \(v,w \in C^2({{\overline{D}}}, {\mathbb {R}})\), it holds that

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathscr {E}}^{\prime \prime }(\varphi )[v,w]&{}=&{}\displaystyle - \frac{N}{2}\int _{D} e^{N\varphi }\, v\, w\left( \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\right) ^2 \mathrm{d}\sigma \\ &{}&{} \displaystyle - \int _{D} e^{N\varphi }\, v\, \frac{\partial u_{\Omega _\varphi }}{\partial \nu } (e^{\varphi }q)\, \frac{\partial u^\prime _w}{\partial \nu } (e^{\varphi }q) \ \mathrm{d}\sigma \\ &{}&{}\displaystyle - \int _{D} e^{N\varphi } v\, w\, \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\, D^2u_{\Omega _\varphi }(e^{\varphi (q)}q) e^{\varphi }q \varvec{\cdot } \nu \ \mathrm{d}\sigma \\ &{}&{}\displaystyle + \int _{D} e^{N\varphi } v\, \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\, \frac{\nabla u_{\Omega _\varphi }(e^{\varphi }q) \varvec{\cdot } \nabla _{{\mathbb {S}}^{N-1}} w}{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}}\varphi |^2}} \ \mathrm{d}\sigma \\ &{}&{}\displaystyle +\int _{D} e^{N\varphi } v\left( \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\right) ^2 \frac{\nabla _{{\mathbb {S}}^{N-1}}\varphi \varvec{\cdot } \nabla _{{\mathbb {S}}^{N-1}} w}{(1+|\nabla _{{\mathbb {S}}^{N-1}}\varphi |^2)} \ \mathrm{d}\sigma , \end{array} \end{aligned}$$
(3.12)

where \(u^\prime _w=\left. \frac{\mathrm{d}}{\mathrm{d}s}\left( u_{\Omega _{\varphi +sw}}\right) \right| _{s=0}\) is the solution to (3.9) with V given by \(V(x)=w\left( \frac{x}{|x|}\right) x\).

Proof

Let us fix \(v, w \in C^2({{\overline{D}}}, {\mathbb {R}})\), by definition and by Lemma 3.1 we have

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathscr {E}}^{\prime \prime }(\varphi )[v,w]= & {} \displaystyle \left. \frac{\mathrm{d}}{\mathrm{d}s}\left( - \frac{1}{2}\int _{D} e^{N(\varphi +sw)} v\left( \frac{\partial u_{\Omega _{\varphi +sw}}}{\partial \nu }(e^{\varphi +sw}q)\right) ^2 \ \mathrm{d}\sigma \right) \right| _{s=0},\ \end{array} \nonumber \\ \end{aligned}$$
(3.13)

and thus

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathscr {E}}^{\prime \prime }(\varphi )[v,w] &{}{}=&{}{}\displaystyle - \frac{1}{2}\int _{D} e^{N\varphi }N v\, w\, \left( \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\right) ^2\, \ \mathrm {d}\sigma \\ &{}{}&{}{}\displaystyle - \int _{D} e^{N\varphi } v\, \frac{\partial u_{\Omega _{\varphi }}}{\partial \nu }(e^{\varphi }q)\left. \frac{\mathrm {d}}{\mathrm {d}s}\left( \frac{\partial u_{\Omega _{\varphi +sw}}}{\partial \nu }(e^{\varphi +sw}q)\right) \right| _{s=0} \ \mathrm {d}\sigma . \end{array} \end{aligned}$$
(3.14)

Since \(\Gamma _{\varphi +sw}\) is a radial graph, then, in view of (2.7), we have

$$\begin{aligned} \begin{array}{lll} \displaystyle \frac{\partial u_{\Omega _{\varphi +sw}}}{\partial \nu }(e^{\varphi +sw}q)= & {} \displaystyle \nabla u_{\Omega _{\varphi +sw}}(e^{\varphi +sw}q) \varvec{\cdot } \frac{q-\nabla _{{\mathbb {S}}^{N-1}}(\varphi +sw)}{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}} (\varphi +sw)|^2}}. \end{array} \end{aligned}$$
(3.15)

As in the proof of Lemma 3.1 we consider the map \({\hat{\Phi }}:(-\delta ,\delta )\rightarrow H^{1}_0(\Omega _\varphi \cup \Gamma _{1,\varphi })\), defined by

$$\begin{aligned} {\hat{\Phi }}(s)= {\hat{u}}_s:=u_{\Omega _{\varphi +sw}}\circ \xi (s,\cdot )|_{\Omega _\varphi }. \end{aligned}$$

Moroever, let \(G:H^{1}_0(\Omega _\varphi \cup \Gamma _{1,\varphi }) \rightarrow L^2(\Omega _\varphi \cup \Gamma _{1,\varphi }, {\mathbb {R}}^N)\), given by \(G(f):=\nabla f\). Since G is a bounded linear operator, then G is differentiable, \(G^\prime (f)[g]=\nabla g\) for any \(g \in H^{1}_0(\Omega _\varphi \cup \Gamma _{1,\varphi })\). In addition, as \({\hat{\Phi }}\) is differentiable (see the proof of [23, Proposition 4.3] for the details), then the composition \(G\circ {\hat{\Phi }}: (-\delta ,\delta )\rightarrow L^2(\Omega _\varphi \cup \Gamma _{1,\varphi }, {\mathbb {R}}^N)\) is differentiable and

$$\begin{aligned} (G\circ {\hat{\Phi }})^\prime (s)=G^\prime ({\hat{\Phi }}(s))[{\hat{\Phi }}^\prime (s)]=\nabla {\hat{\Phi }}^\prime (s) \ \ \forall s\in (-\delta ,\delta ). \end{aligned}$$

In terms of \({\hat{u}}_s\), this means that

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}s}\left( \nabla {\hat{u}}_s\right) = \nabla \left( \frac{\mathrm{d}{\hat{u}}_s}{\mathrm{d}s}\right) . \end{aligned}$$
(3.16)

In addition, since \(u_{\Omega _{\varphi +sw}}={\hat{u}}_s \circ \xi (-s,\cdot )|_{\Omega _{\varphi +sw}}\) it follows that also \(s\mapsto \nabla u_{\Omega _{\varphi +sw}}\) is differentiable. We claim that

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}s}\left( \nabla u_{\Omega _{\varphi +sw}}\right) =\nabla \left( \frac{\mathrm{d}}{\mathrm{d}s} u_{\Omega _{\varphi +sw}}\right) . \end{aligned}$$
(3.17)

Indeed, setting \(\xi _s:=\xi (-s,\cdot )|_{\Omega _{\varphi +sw}}\), since

$$\begin{aligned} \frac{\partial }{\partial x_i} u_{\Omega _{\varphi +sw}} =\frac{\partial }{\partial x_i}\left( {\hat{u}}_s \circ \xi _s\right) =\nabla {\hat{u}}_s(\xi _s) \varvec{\cdot } \frac{\partial \xi _s}{\partial x_i}, \end{aligned}$$

then, using (3.16), we get that

$$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}s}\left( \frac{\partial }{\partial x_i}u_{\Omega _{\varphi +sw}}\right)&= \left( \nabla \left( \frac{d {\hat{u}}_s}{\mathrm {d}s}\right) (\xi _s)+ D^2{\hat{u}}_s(\xi _s) \frac{\mathrm {d}\xi _s}{\mathrm {d}s}\right) \mathbf {\cdot } \frac{\partial \xi _s}{\partial x_i}\\&\quad +\nabla {\hat{u}}_s(\xi _s) \mathbf {\cdot }\frac{\partial }{\partial x_i}\frac{\mathrm {d}\xi _s}{\mathrm {d}s}=\frac{\partial }{\partial x_i}\left( \frac{\mathrm {d}}{\mathrm {d}s} u_{\Omega _{\varphi +sw}}\right) \end{aligned}$$

and, thus by a straightforward computation, we obtain

$$\begin{aligned}&\frac{\mathrm{d}}{\mathrm{d}s}\left( \frac{\partial }{\partial x_i}u_{\Omega _{\varphi +sw}}\right) = \left( \nabla \left( \frac{d {\hat{u}}_s}{\mathrm{d}s}\right) (\xi _s)+ D^2{\hat{u}}_s(\xi _s) \frac{\mathrm{d}\xi _s}{\mathrm{d}s}\right) \varvec{\cdot } \frac{\partial \xi _s}{\partial x_i}\\&\quad +\nabla {\hat{u}}_s(\xi _s) \varvec{\cdot }\frac{\partial }{\partial x_i}\frac{\mathrm{d}\xi _s}{\mathrm{d}s}=\frac{\partial }{\partial x_i}\left( \frac{\mathrm{d}}{\mathrm{d}s} u_{\Omega _{\varphi +sw}}\right) \end{aligned}$$

for any \(i=1,\ldots ,N\), which proves Claim (3.17).

Thanks to (3.15) and (3.17) we have

$$\begin{aligned} \begin{array}{lll} &{}{}&{}{}\displaystyle \frac{\mathrm {d}}{\mathrm {d}s}\left. \left( \frac{\partial u_{\Omega _{\varphi +sw}}}{\partial \nu }(e^{\varphi +sw}q)\right) \right| _{s=0}\\[4pt] &{}{}&{}{}\quad =\displaystyle \left( \nabla u^\prime _w(e^{\varphi }q)+D^2 u_{\Omega _\varphi }(e^{\varphi }q)e^{\varphi }w q\right) \mathbf {\cdot } \frac{q-\nabla _{{\mathbb {S}}^{N-1}}\varphi }{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2}}\\[4pt] &{}{}&{}{}\quad \displaystyle + \nabla u_{\Omega _{\varphi }}(e^{\varphi }q) \mathbf {\cdot } \left( - \frac{\nabla _{{\mathbb {S}}^{N-1}}w}{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2}} - \frac{(q-\nabla _{{\mathbb {S}}^{N-1}}\varphi ) (\nabla _{{\mathbb {S}}^{N-1}} \varphi \mathbf {\cdot } \nabla _{{\mathbb {S}}^{N-1}}w)}{(1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2)^{3/2}}\right) \\ &{}{}&{}{}\quad = \left( \nabla u^\prime _w(e^{\varphi }q)+D^2 u_{\Omega _\varphi }(e^{\varphi }q)e^{\varphi }w q\right) \mathbf {\cdot } \nu \\ &{}{}&{}{}\quad \displaystyle + \nabla u_{\Omega _{\varphi }}(e^{\varphi }q) \mathbf {\cdot } \left( -\frac{\nabla _{{\mathbb {S}}^{N-1}}w}{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2}} -\frac{\nabla _{{\mathbb {S}}^{N-1}} \varphi \mathbf {\cdot } \nabla _{{\mathbb {S}}^{N-1}}w}{(1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2)}\, \nu \right) \\ &{}{}&{}{}\quad =\displaystyle \frac{\partial u^\prime _w}{\partial \nu }(e^{\varphi }q)+ w\, D^2 u_{\Omega _\varphi }(e^{\varphi }q)e^{\varphi }q \mathbf {\cdot } \nu \\ &{}{}&{}{}\quad \displaystyle - \frac{ \nabla u_{\Omega _{\varphi }}(e^{\varphi }q) \mathbf {\cdot } \nabla _{{\mathbb {S}}^{N-1}}w}{\sqrt{1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2}} - \frac{\partial u_{\Omega _\varphi }}{\partial \nu }(e^{\varphi }q)\frac{\nabla _{{\mathbb {S}}^{N-1}} \varphi \mathbf {\cdot } \nabla _{{\mathbb {S}}^{N-1}}w}{(1+|\nabla _{{\mathbb {S}}^{N-1}} \varphi |^2)}. \end{array}\nonumber \\ \end{aligned}$$
(3.18)

Finally, combining (3.14) and (3.18) we readily obtain (3.12). The proof is complete.

\(\square \)

4 Volume-Constrained Critical Points for the Torsional Energy of Star-Shaped Domains

For any \(\varphi \in C^2({{\overline{D}}},{\mathbb {R}})\), the volume of the associated star-shaped domain \(\Omega _\varphi \) (see (3.3)) is given by

$$\begin{aligned} {\mathcal {V}}(\varphi )=|\Omega _\varphi |=\frac{1}{N}\int _D e^{N\varphi } \ \mathrm{d}\sigma , \end{aligned}$$
(4.1)

where \(\mathrm{d}\sigma \) is the \((N-1)\)-dimensional area element of \({\mathbb {S}}^{N-1}\). It is easy to check that \({\mathcal {V}}\) is of class \(C^2\) and for any \(v,w \in C^2({{\overline{D}}},{\mathbb {R}})\) it holds

$$\begin{aligned} {\mathcal {V}}^\prime (\varphi )[v]= \int _D e^{N\varphi }\, v \ \mathrm{d}\sigma , \end{aligned}$$
(4.2)

and

$$\begin{aligned} {\mathcal {V}}^{\prime \prime }(\varphi )[v,w]= N \int _D e^{N\varphi }\,v\, w \ \mathrm{d}\sigma . \end{aligned}$$
(4.3)

For a number \(c>0\) we define

$$\begin{aligned} M=\{\varphi \in C^2({{\overline{D}}}, {\mathbb {R}}); \ {\mathcal {V}}(\varphi )=c\}. \end{aligned}$$
(4.4)

Clearly M is a smooth manifold and for any \(\varphi \in M\) it holds

$$\begin{aligned} T_\varphi M=\left\{ v \in C^2({{\overline{D}}}, {\mathbb {R}}); \ \int _D e^{N\varphi }\, v \ \mathrm{d}\sigma = 0\right\} . \end{aligned}$$
(4.5)

We consider the restriction of the torsional energy to the domains corresponding to functions \(\varphi \in M\), namely the functional defined by

$$\begin{aligned} I(\varphi ):=\left. {\mathscr {E}}(\varphi )\right| _{\varphi \in M}=\left. {\mathcal {E}}(\Omega _\varphi ; \Sigma _D)\right| _{\varphi \in M}. \end{aligned}$$
(4.6)

If \(\varphi \in M\) is critical point of I then there exists \(\uplambda \in {\mathbb {R}}\) such that

$$\begin{aligned} {\mathscr {E}}^\prime (\varphi ) = \uplambda {\mathcal {V}}^\prime (\varphi ). \end{aligned}$$
(4.7)

As a straightforward consequence of Lemma 3.1 and (4.2) we have

Lemma 4.1

Let \(\varphi \in M\) be a critical point for I and assume that \(u_{\Omega _\varphi }\in W^{1,\infty }(\Omega _\varphi )\cap W^{2,2}(\Omega _\varphi )\). Then the Lagrange multiplier \(\uplambda \) is negative and

$$\begin{aligned} \frac{\partial u_{\Omega _\varphi }}{\partial \nu }\equiv - \sqrt{-2\uplambda }\ \ \hbox { on}\ \Gamma _\varphi . \end{aligned}$$

Proof

Let \(\varphi \in M\) be a critical point for I and assume that \(u_{\Omega _\varphi }\in W^{1,\infty }(\Omega _\varphi )\cap W^{2,2}(\Omega _\varphi )\), then, from (4.7) and exploiting Lemma 3.1 and (4.2), we have

$$\begin{aligned} -\frac{1}{2} \int _D e^{N\varphi }\, v \left( \frac{\partial u_{\Omega _\varphi } }{\partial \nu }\left( e^{\varphi }q\right) \right) ^2 \ \mathrm{d}\sigma = \uplambda \int _D e^{N\varphi }\, v \ \mathrm{d}\sigma , \end{aligned}$$

for any \(v \in C^2({{\overline{D}}}, {\mathbb {R}})\). Hence we readily obtain that

$$\begin{aligned} \int _D e^{N\varphi }\, v\left[ \left( \frac{\partial u_{\Omega _\varphi } }{\partial \nu }\left( e^{\varphi }q\right) \right) ^2 +2 \uplambda \right] \ \mathrm{d}\sigma =0, \end{aligned}$$

and from the arbitrariness of \(v \in C^2({{\overline{D}}}, {\mathbb {R}})\) we easily deduce that \(\uplambda <0\) and

$$\begin{aligned} \left( \frac{\partial u_{\Omega _\varphi } }{\partial \nu }\right) ^2 =-2 \uplambda \ \ \text{ on }\ \Gamma _\varphi . \end{aligned}$$
(4.8)

Now, recalling that \(u_{\Omega _\varphi }\) is the only (positive) weak solution to

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 1 &{} \text {in }\Omega _\varphi ,\\ u = 0 &{} \text {on}\ \Gamma _\varphi ,\\ \frac{\partial u}{\partial \nu } = 0 &{} \text {on}\ \Gamma _{1,\varphi }\setminus \{0\}, \end{array}\right. } \end{aligned}$$
(4.9)

then from standard regularity estimates we infer that \(u_{\Omega _\varphi }\) is smooth in \(\Omega _\varphi \), and from Hopf’s lemma we get that \(\frac{\partial u_{\Omega _\varphi } }{\partial \nu }<0\) on \(\Gamma _\varphi \). Hence, in view of (4.8), we obtain

$$\begin{aligned} \frac{\partial u_{\Omega _\varphi }}{\partial \nu } =-\sqrt{-2 \uplambda } \ \hbox { on}\ \Gamma _\varphi . \end{aligned}$$

\(\square \)

Remark 4.2

From Lemma 4.1 we deduce that each critical point of I produces a star-shaped domain \(\Omega _\varphi \) for which the overdetermined problem (1.2) has a solution. We recall that, as shown in [23, Proposition 4.3], each critical point of the functional \({\mathcal {E}}(\Omega ; \Sigma _D)\) on the whole family of domains in \(\Sigma _D\), with a volume constraint, is a domain for which (1.2) has a solution. Hence Lemma 4.1 shows that the same statement holds even if the variations are taken only in the class of star-shaped domains.

In the next result we compute the second derivative of I at critical point along variations in \(T_\varphi M\).

Lemma 4.3

Let \(\varphi \in M\) be a critical point for I and let \(v,w \in T_\varphi M\). Then

$$\begin{aligned} I^{\prime \prime }(\varphi )[v,w]={\mathscr {E}}^{\prime \prime }(\varphi )[v,w] - \uplambda {\mathcal {V}}^{\prime \prime }(\varphi )[v,w], \end{aligned}$$

where \(\uplambda \) is the Lagrange multiplier.

Proof

By definition if \(\varphi \in M\) is a critical point for I, the second variation \(I^{\prime \prime }(\varphi )[v,w]\) along the variations \(v,w \in T_\varphi M\) is given by

$$\begin{aligned} I^{\prime \prime }(\varphi )[v,w]=\left. \frac{\partial ^2 I(\Psi (t,s))}{\partial t \partial s}\right| _{(t,s)=(0,0)}, \end{aligned}$$

where \(\Psi :(-\epsilon ,\epsilon )\times (-\epsilon ,\epsilon )\rightarrow M\) is a smooth surface in M such that

$$\begin{aligned} \Psi (0,0)=\varphi , \ \frac{\partial \Psi }{\partial t}(0,0)=v, \ \frac{\partial \Psi }{\partial s}(0,0)=w. \end{aligned}$$

We recall that by definition it holds \(I(\Psi (t,s))= {\mathscr {E}}(\Psi (t,s))\). Since

$$\begin{aligned} \frac{\partial }{\partial s} \left( {\mathscr {E}}(\Psi (t,s))\right) ={\mathscr {E}}^\prime (\Psi (t,s))\left[ \frac{\partial \Psi }{\partial s}(t,s)\right] , \end{aligned}$$

we have

$$\begin{aligned} \frac{\partial }{\partial t} \frac{\partial }{\partial s}\left( {\mathscr {E}}(\Psi (t,s))\right) ={\mathscr {E}}^{\prime \prime }(\Psi (t,s))\left[ \frac{\partial \Psi }{\partial s}(t,s), \frac{\partial \Psi }{\partial t}(t,s)\right] + {\mathscr {E}}^\prime (\Psi (t,s))\left[ \frac{\partial ^2\Psi }{\partial t\partial s}(t,s)\right] .\nonumber \\ \end{aligned}$$
(4.10)

On the other hand, since \(\Psi (t,s) \in M\) we have \({\mathcal {V}}(\Psi (t,s))=c\) for any \((t,s)\in (-\epsilon ,\epsilon )\times (-\epsilon ,\epsilon )\), and thus differentiating with respect to t we infer that \({\mathcal {V}}^\prime (\Psi (t,s))[\frac{\partial \Psi }{\partial s}(t,s)]=0\). Differentiating again with respect to s we obtain

$$\begin{aligned} {\mathcal {V}}^{\prime \prime }(\Psi (t,s))\left[ \frac{\partial \Psi }{\partial s}(t,s), \frac{\partial \Psi }{\partial t}(t,s)\right] +{\mathcal {V}}^\prime (\Psi (t,s))\left[ \frac{\partial }{\partial t}\frac{\partial \Psi }{\partial s}(t,s)\right] =0 \end{aligned}$$
(4.11)

Hence, computing (4.10) at \((t,s)=(0,0)\), since \(\varphi =\Psi (0,0)\) is a critical point of I and taking into account that (4.7), (4.11), we get that

$$\begin{aligned} \frac{\partial }{\partial t}\frac{\partial }{\partial s}\left. \left( {\mathscr {E}}(\Psi (t,s))\right) \right| _{(t,s)=(0,0)}= & {} {\mathscr {E}}^{\prime \prime }(\Psi (t,s))\left[ v,w\right] +{\mathscr {E}}^{\prime }(\Psi (t,s))\left[ \frac{\partial }{\partial t}\frac{\partial \Psi }{\partial s}(0,0)\right] \\= & {} {\mathscr {E}}^{\prime \prime }(\Psi (t,s))\left[ v,w\right] +\uplambda {\mathcal {V}}^\prime (\varphi ) \left[ \frac{\partial }{\partial t}\frac{\partial \Psi }{\partial s}(0,0)\right] \\= & {} {\mathscr {E}}^{\prime \prime }(\Psi (t,s))\left[ v,w\right] -\uplambda {\mathcal {V}}^{\prime \prime }(\varphi ) \left[ v,w\right] , \end{aligned}$$

which proves the desired relation. \(\square \)

Remark 4.4

When \(\varphi \equiv 0\) then \(\Omega _\varphi \) is the unit spherical sector \(\Omega _{D,1}=\Sigma _D\cap B_1\), where \(B_1=B_1(0)\) is the unit ball in \({\mathbb {R}}^N\) centered at the origin. We denote it by \(\Omega _0\), while \(\Gamma _0\) will be its relative boundary. In this case the torsion function \(u_{\Omega _0}\) is known to be the radial function \(u_{\Omega _0}(x)=\frac{1-|x|^2}{2N}\). Then we can choose \(c=|\Omega _0|\) in the definition of M and the tangent space to M at \(\varphi \equiv 0\) is \(T_0M=\{v \in C^2({{\overline{D}}}, {\mathbb {R}});\ \int _D v \ \mathrm{d}\sigma =0 \}\). It is easy to check that \(\nabla u_{\Omega _0}=-\frac{1}{N}x\), for \(x \in \Sigma _D\cap B_1\), and \(\frac{\partial u_{\Omega _0}}{\partial \nu }=-\frac{1}{N}\) on \(\Gamma _0\), so that \(\Omega _0\) is a critical point for I with \(\uplambda =-\frac{1}{2N^2}\). Finally \(D^2 u_{\Omega _0}(x)=-\frac{1}{N} {\mathbb {I}}_N\), for \(x \in \Sigma _D\cap B_1\), where \({\mathbb {I}}_N\) is the identity matrix of order N, and thus we readily have that \(u_{\Omega _0}\in W^{1,\infty }(\Omega _0)\cap W^{2,2}(\Omega _0)\).

For the second variation we have

Proposition 4.5

For any \(v \in T_0M\) it holds that

$$\begin{aligned} \displaystyle I^{\prime \prime }(0)[v,v]=\displaystyle - \frac{1}{N^2}\int _{D} v^2 \ \mathrm{d}\sigma \displaystyle +\frac{1}{N} \int _{D} v \frac{\partial u^\prime }{\partial \nu } \ \mathrm{d}\sigma , \end{aligned}$$
(4.12)

where \(u^\prime =\left. \frac{\mathrm{d}}{\mathrm{d}t}\left( u_{\Omega _{0+tv}}\right) \right| _{t=0}\) (see (3.9)) and \(\frac{\partial u^\prime }{\partial \nu }\) is the normal derivative of \(u^\prime \) on \(\Gamma _0=D\).

Proof

First we observe that, taking \(\varphi \equiv 0\), from (3.9) and (3.5) we have that \(u^\prime \) satisfies

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u^\prime = 0 &{} \hbox { in}\ \Omega _0,\\ u^\prime = \frac{1}{N} v &{} \hbox { on}\ \Gamma _0,\\ \frac{\partial u^\prime }{\partial \nu } = 0 &{} \hbox { on}\ \Gamma _{1,0}. \end{array}\right. } \end{aligned}$$
(4.13)

Then, taking \(v \in C^2({{\overline{D}}}, {\mathbb {R}})\) such that \(\int _D v \ \mathrm{d}\sigma =0\), from Lemma 3.2, Lemma 4.3, Remark 4.4 and (4.3), (4.13) we obtain

$$\begin{aligned} \begin{array}{lll} \displaystyle I^{\prime \prime }[v,v]&{}{}=&{}{}\displaystyle - \frac{N}{2}\int _{D} v^2 \left( -\frac{1}{N}\right) ^2 \ \mathrm {d}\sigma - \int _{D} v \left( -\frac{1}{N}\right) \frac{\partial u^\prime }{\partial \nu } \ \mathrm {d}\sigma \\ &{}{}&{}{}\displaystyle - \int _{D} v^2 \left( -\frac{1}{N}\right) \left( -\frac{1}{N}q \mathbf {\cdot } q \right) \ \mathrm {d}\sigma + \int _{D} v \left( -\frac{1}{N}\right) \nabla u_{\Omega _0} \mathbf {\cdot } \nabla _{{\mathbb {S}}^{N-1}} v\ \mathrm {d}\sigma \\ &{}{}&{}{}\displaystyle - \left( -\frac{1}{2N^2}\right) N \int _D v^2 \ \mathrm {d}\sigma \\ &{}{}=&{}{}\displaystyle \left( - \frac{1}{2N}-\frac{1}{N^2}+\frac{1}{2N}\right) \int _{D} v^2 \ \mathrm {d}\sigma +\frac{1}{N} \int _{D} v \frac{\partial u^\prime }{\partial \nu } \ \mathrm {d}\sigma , \end{array}\nonumber \\ \end{aligned}$$
(4.14)

since \(\nabla u_{\Omega _0} \varvec{\cdot } \nabla _{{\mathbb {S}}^{N-1}} v\equiv 0\) in D because \(\nabla u_{\Omega _0}\) is proportional to the radial direction. \(\square \)

Remark 4.6

We observe that thanks to (4.12), since \(u^\prime =\frac{1}{N} v\) on \(\Gamma _0\), by (4.13) and recalling that \(\Gamma _0=D\), we can write

$$\begin{aligned} I^{\prime \prime }(0)[v,v]=\displaystyle - \int _{D} (u^\prime )^2 \ \mathrm{d}\sigma + \int _{D} u^\prime \frac{\partial u^\prime }{\partial \nu } \ \mathrm{d}\sigma . \end{aligned}$$

Then by Green’s identity and (4.13) we infer that

$$\begin{aligned} I^{\prime \prime }(0)[v,v]=\displaystyle - \int _{D} (u^\prime )^2 \ \mathrm{d}\sigma + \int _{\Omega _0} |\nabla u^\prime |^2 \ \mathrm{d}x. \end{aligned}$$

5 A Condition for Instability

In this section we provide conditions on the domain \(D\subset {\mathbb {S}}^{N-1}\) such that the corresponding spherical sector (i.e. the domain \(\Omega _0\) associated to the function \(\varphi \equiv 0\), see (3.3)) is not a local minimizer for the torsional energy functional under a volume constraint. This is achieved by showing that \(\Omega _0\) is an unstable critical point of I, i.e. its Morse index is positive.

More precisely, let M be the manifold defined in (4.4), with \(c=|\Omega _0|\) and let I be as in (4.6). As observed in Remark 4.4 the function \(\varphi \equiv 0\) belongs to M and \(\Omega _0\) is a critical point for I. The main result of this section is the following:

Theorem 5.1

Let \(D\subset {\mathbb {S}}^{N-1}\) be a smooth domain and let \(\uplambda _1(D)\) be the first non trivial eigenvalue of the Laplace-Beltrami operator \(-\Delta _{{\mathbb {S}}^{N-1}}\), with zero Neumann condition on \(\partial D\). It holds that

  1. (i)

    if \(\uplambda _1(D) < N-1\), then \(\Omega _0\) is not a local minimizer for I;

  2. (ii)

    if \(\uplambda _1(D) > N-1\), then \(\Omega _0\) is a local minimizer for I.

Proof

To prove (i), let \((w_j)_{j\in {\mathbb {N}}}\) be a \(L^2(D)\)-orthonormal basis of eigenfunctions of the eigenvalue problem

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta _{{\mathbb {S}}^{N-1}} w_j=\uplambda _j w_j &{} \hbox {in D},\\ \quad \quad \quad \frac{\partial w_j}{\partial \nu _{_{\partial D}}} =0 &{} \hbox {on }\partial D,\end{array}\right. } \end{aligned}$$
(5.1)

where \(\nu _{_{\partial D}}\) is the exterior unit co-normal to \(\partial D\), i.e. for any \(q\in \partial D\), \(\nu _{_{\partial D}}(q)\) is the only unit vector in \(T_q{\mathbb {S}}^{N-1}\) such that \(\nu _{_{\partial D}}(q)\perp T_q\partial D\) and \(\nu _{_{\partial D}}(q)\) points outward D. We define the following extension of \(w_j\) to the cone \(\Sigma _D\)

$$\begin{aligned} {\tilde{w}}_j(rq):=\frac{1}{N}r^{\alpha _j} w_j(q) \ \ q\in D, r>0, \end{aligned}$$
(5.2)

where

$$\begin{aligned} \alpha _j:=-\frac{N-2}{2}+\sqrt{\left( \frac{N-2}{2}\right) ^2 +\uplambda _j}. \end{aligned}$$
(5.3)

We claim that \(w={\tilde{w}}_j\big |_{\Omega _0}\) is the unique solution of

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta w = 0 &{} \hbox { in}\ \Omega _0,\\ w = \frac{1}{N} w_j &{} \hbox { on}\ \Gamma _0,\\ \frac{\partial w}{\partial \nu } = 0 &{} \hbox { on}\ \Gamma _{1,0}. \end{array}\right. } \end{aligned}$$
(5.4)

Indeed, writing the Laplace operator in polar coordinates and exploiting (5.1) we easily check that

$$\begin{aligned} \Delta {\tilde{w}}_j= & {} \frac{\partial ^2 {\tilde{w}}_j}{\partial ^2 r} + \frac{N-1}{r} \frac{\partial {\tilde{w}}_j}{\partial r} + \frac{1}{r^2} \Delta _{{\mathbb {S}}^{N-1}} {\tilde{w}}_j \\= & {} \left( \alpha _j(\alpha _j-1)+ \alpha _j(N-1)-\uplambda _j\right) \frac{r^{\alpha _j-2} w_j(q)}{N}=0, \end{aligned}$$

because \(\alpha _j\) satisfies \(\alpha _j^2 + (N-2)\alpha _j -\uplambda _j=0\). Moreover, by definition, we have \({\tilde{w}}_j\big |_D=\frac{1}{N}w_j\) and \(\frac{\partial {\tilde{w}}_j}{\partial \nu } = 0\) on \(\Gamma _{1,0}\).

Now, let us take \(j=1\). It is well known that the first eigenfunction \(w_1\) is smooth and satisfies \(\int _D w_1 \ \mathrm{d}\sigma =0\), i.e. \(w_1\in T_0M\). Computing \(I^{\prime \prime }(0)[w_1,w_1]\), thanks to Proposition 4.5 and taking into account that \({\tilde{w}}_1\big |_{\Omega _0}\) is the solution of (5.4), with \(j=1\), we get that

$$\begin{aligned} \displaystyle I^{\prime \prime }(0)[w_1,w_1]=\displaystyle - \frac{1}{N^2}\int _{D} w_1^2 \ \mathrm{d}\sigma \displaystyle +\frac{1}{N} \int _{D} w_1 \left( \frac{\partial {\tilde{w}}_1}{\partial \nu }\right) \ \mathrm{d}\sigma . \end{aligned}$$
(5.5)

Then, since the \(L^2(D)\)-norm of \(w_1\) is equal to 1, the exterior unit normal \(\nu \) to \(\Gamma _0\) is the radial direction, and

$$\begin{aligned} \frac{\partial {\tilde{w}}_1}{\partial \nu }=\frac{1}{N}\alpha _1 r^{\alpha _1-1} w_1=\frac{\alpha _1}{N} w_1\ \ \hbox {on }D, \end{aligned}$$
(5.6)

from (5.5) we obtain

$$\begin{aligned} \displaystyle I^{\prime \prime }(0)[w_1,w_1]=\displaystyle - \frac{1}{N^2} +\frac{\alpha _1}{N^2}. \end{aligned}$$

Thus we deduce that

$$\begin{aligned} I^{\prime \prime }(0)[w_1,w_1]<0 \ \ \hbox {if and only if} \ \ -1+\alpha _1<0. \end{aligned}$$

Finally, from (5.3) it is immediate to check that \(\alpha _1<1\) is equivalent to \(\uplambda _1(D)<N-1\) and the proof of (i) is complete.

To prove (ii), let \(v\in T_0M\) and assume, without loss of generality, that \(\int _D v^2 \ \mathrm{d}\sigma =1\). Taking \((w_j)_{j\in {\mathbb {N}}}\) as in the proof of (i), since \(v \in T_0 M\) we can write

$$\begin{aligned} v=\sum _{j=1}^\infty (v,w_j)_{L^2(D)} w_j. \end{aligned}$$

Let \({\tilde{w}}_j\) be the harmonic extension of \(w_j\) defined in (5.2). Then, as \({\tilde{w}}_j\big |_{\Omega _0}\) is a solution to (5.4) for any \(j\in {\mathbb {N}}\), we infer that \({\tilde{v}}:=\sum _{j=1}^\infty (v,w_j)_{L^2(D)} {\tilde{w}}_j\) is a solution to

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 0 &{} \hbox { in}\ \Omega _0,\\ u = \frac{1}{N} v &{} \hbox { on}\ \Gamma _0,\\ \frac{\partial u}{\partial \nu } = 0 &{} \hbox { on}\ \Gamma _{1,0}. \end{array}\right. } \end{aligned}$$

Thus, by Proposition 4.5, we get

$$\begin{aligned} \displaystyle I^{\prime \prime }(0)[v,v]=\displaystyle - \frac{1}{N^2}\int _{D} v^2\ \mathrm{d}\sigma \displaystyle +\frac{1}{N} \int _{D} v \ \frac{\partial {\tilde{v}}}{\partial \nu } \ \mathrm{d}\sigma . \end{aligned}$$

As in (5.6) we have that \(\frac{\partial {\tilde{w}}_j}{\partial \nu }=\frac{\alpha _j}{N} w_j\) on D, for any \(j\in {\mathbb {N}}\). Hence, since \(\int _D v^2 \ \mathrm{d}\sigma =1\), we deduce

$$\begin{aligned} \displaystyle I^{\prime \prime }(0)[v,v]=\displaystyle - \frac{1}{N^2} +\frac{1}{N^2} \int _{D} v \sum _{j=1}^\infty \alpha _j (v,w_j)_{L^2(D)} w_j \ \mathrm{d}\sigma =- \frac{1}{N^2} +\frac{1}{N^2} \sum _{j=1}^\infty \alpha _j (v,w_j)_{L^2(D)}^2. \end{aligned}$$

Now, if \(\uplambda _1(D)>N-1\) it follows that \(\alpha _1>1\), and, as \((\alpha _j)_{j\in {\mathbb {N}}}\) is a nondecreasing sequence, we obtain

$$\begin{aligned} \displaystyle I^{\prime \prime }(0)[v,v]>- \frac{1}{N^2} +\frac{1}{N^2} \sum _{j=1}^\infty (v,w_j)_{L^2(D)}^2=0, \end{aligned}$$
(5.7)

having used that \(\sum _{j=1}^\infty (v,w_j)_{L^2(D)}^2=1\), as \(\int _D v^2(q) \ \mathrm{d}\sigma =1\). Hence (ii) holds. \(\square \)

We conclude this section with a useful criterion for checking the property \(\uplambda _1(D)<N-1\). To this end let \({e} \in {\mathbb {S}}^{N-1}\) and let \(u_{{e}}\in C^\infty ({\mathbb {R}}^N)\) be the function defined by

$$\begin{aligned} u_{{e}}(x)=x \varvec{\cdot } {e}, \end{aligned}$$
(5.8)

which satisfies

$$\begin{aligned} -\Delta _{{\mathbb {S}}^{N-1}} u_{{e}}=(N-1) u_{{e}} \ \ \text{ on }\ {\mathbb {S}}^{N-1}. \end{aligned}$$
(5.9)

We have

Proposition 5.2

Let D be a smooth proper domain of \({\mathbb {S}}^{N-1}\) and let \({e}\in {\mathbb {S}}^{N-1}\) satisfy

$$\begin{aligned} \int _{D} u_{{e}}\ \mathrm{d}\sigma =0. \end{aligned}$$

Assume that either one of the following holds:

  1. (i)

    \(\displaystyle \int _{\partial D} u_{{e}} \frac{\partial u_{{e}}}{\partial \nu } \ \mathrm{d}{\hat{\sigma }}<0\);

  2. (ii)

    \(\displaystyle \int _{\partial D} u_{{e}} \frac{\partial u_{{e}}}{\partial \nu } \ \mathrm{d}{\hat{\sigma }}=0\), and \(u_{{e}}\) is not an eigenfunction of \( {\left\{ \begin{array}{ll} -\Delta _{{\mathbb {S}}^{N-1}} w=\uplambda w&{}{} \text{ in }\ D,\\ \quad \quad \quad \quad \frac{\partial w}{\partial \nu } =0 &{}{} \text{ on }\ \partial {D},\end{array}\right. }\)

where \(\mathrm{d}{\hat{\sigma }}\) is the \((N-2)\)-dimensional area element of \(\partial D\) and \(\nu =\nu _{_{\partial D}}\) is the exterior unit co-normal to \(\partial D\). Then \(\uplambda _1(D)<N-1\).

Proof

Taking \(u_{{e}}\) as test function in the variational characterization of the first non-trivial eigenvalue of \(-\Delta _{{\mathbb {S}}^{N-1}}\) with zero Neumann condition on \(\partial D\), applying Green’s identity and exploiting (5.9), we have

$$\begin{aligned} \int _D |\nabla u_{{e}}|^2 \ \mathrm{d}\sigma =\int _{\partial D} u_{{e}} \frac{\partial u_{{e}}}{\partial \nu } \ \mathrm{d}{\hat{\sigma }}- \int _{D} u_{{e}} \Delta u_{{e}} \ \mathrm{d}\sigma =\int _{\partial D} u_{{e}} \frac{\partial u_{{e}}}{\partial \nu } \ \mathrm{d}{\hat{\sigma }}+ (N-1)\int _{D} u^2_{{e}} \ \mathrm{d}\sigma . \end{aligned}$$

Therefore, if (i) holds it follows that

$$\begin{aligned} \displaystyle \frac{\int _D |\nabla u_{{e}}|^2 \ \mathrm{d}\sigma }{\int _{D} u_{{e}}^2 \ \mathrm{d}\sigma }<N-1 \end{aligned}$$
(5.10)

which implies that \(\uplambda _1(D)<N-1\). This completes the proof for the case (i). On the other hand, under the assumption (ii), the equality sign in (5.10) holds, but as \(u_{{e}}\) is not an eigenfunction it follows that \(N-1\) cannot be the smallest non-trivial eigenvalue. \(\square \)

6 Existence of Volume-Constrained Minimizers for the Torsional Energy

Let \(D\subset {\mathbb {S}}^{N-1}\) be a domain of the unit sphere and let \(\Sigma _D\) be the cone generated by D. We will always assume that D is smooth so that \(\Sigma _D\) is smooth exept at the vertex. In Sect. 3 we defined the torsional energy \({\mathcal {E}}(\Omega ;\Sigma _D)\) for smooth domains \(\Omega \subset \Sigma _D\) strictly star-shaped with respect to the vertex of the cone. In this section we study the minimization problem for the torsional energy under a volume constraint in a larger class of sets. Thus we recall some definitions.

Definition 6.1

We say that \(\Omega \subset {\mathbb {R}}^N\) is quasi-open, if for any \(\varepsilon >0\), there exists an open set \(\Lambda _\varepsilon \) such that \(\mathrm {cap}(\Lambda _\varepsilon )\leqq \varepsilon \) and \(\Omega \cup \Lambda _\varepsilon \) is open, where \(\mathrm {cap}(\Lambda _\varepsilon )\) denotes the capacity of \(\Lambda _\varepsilon \) with respect to the \(H^1\)-norm (see [14, Sect. 3.3] or [9, Sect. 2.1]).

For any quasi-open set \(\Omega \subset \Sigma _D\) we consider the Sobolev space:

$$\begin{aligned} H^1_0(\Omega ; \Sigma _D):=\left\{ u \in H^1(\Sigma _D); \ \ u=0 \ \ \hbox {q.e. on} \ \Sigma _D\setminus \Omega \right\} , \end{aligned}$$

where q.e. means quasi-everywhere, i.e. up to sets of zero capacity.

Definition 6.2

We say that u is a (weak) solution of the mixed boundary value problem

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u = 1 &{} \text {in}\ \Omega ,\\ u = 0 &{} \text {on}\ \partial \Omega \cap \Sigma _D,\\ \frac{\partial u}{\partial \nu } = 0 &{} \text {on}\ \partial \Sigma _D, \end{array}\right. } \end{aligned}$$
(6.1)

if \(u\in H^1_0(\Omega ; \Sigma _D)\) and

$$\begin{aligned} \int _{\Sigma _D} \nabla u \varvec{\cdot } \nabla v \ \mathrm{d}x = \int _{\Sigma _D} v \ \mathrm{d}x \ \ \ \ \forall v \in H^1_0(\Omega ; \Sigma _D). \end{aligned}$$

Remark 6.3

As \(\Sigma _D\) is connected and smooth (execpt at the vertex) then \(\Sigma _D\) is uniformly Lipschitz. Thus if \(|\Omega |<+\infty \) the inclusion \(H^1_0(\Omega ; \Sigma _D)\hookrightarrow L^2(\Sigma _D)\) is compact (see [9, Proposition 2.3-(i)]). This implies that the functional

$$\begin{aligned} J(v)=\frac{1}{2}\int _{\Sigma _D} |\nabla v|^2 \ \mathrm{d}x - \int _{\Sigma _D} v \ \mathrm{d}x \end{aligned}$$
(6.2)

has a unique minimizer \(u_\Omega \in H^1_0(\Omega ; \Sigma _D)\) which is the unique (weak) solution to (6.1), which is called energy function or torsion function of \(\Omega \). We also recall that \(\Omega =\{u_\Omega >0\}\) up to a set of zero capacity (see [9, Proposition 2.8-(e)]). Moreover we denote by \(\uplambda _1(\Omega ; \Sigma _D)\) the first eigenvalue of the Laplacian in \(H_0^1(\Omega ; \Sigma _D)\), i.e.

$$\begin{aligned} \uplambda _1(\Omega ; \Sigma _D)=\min _{v \in H_0^1(\Omega ; \Sigma _D)\setminus \{0\}} \frac{\int _{\Sigma _D} |\nabla v|^2 \ \mathrm{d}x}{\int _{\Sigma _D} v^2 \ \mathrm{d}x}. \end{aligned}$$
(6.3)

Then, as before, we define the torsional energy of \(\Omega \) (relative to \(\Sigma _D\)) as:

$$\begin{aligned} {\mathcal {E}}(\Omega ; \Sigma _D)=J(u_\Omega )=-\frac{1}{2} \int _\Omega |\nabla u_\Omega |^2 \ \mathrm{d}x=-\frac{1}{2} \int _\Omega u_\Omega \ \mathrm{d}x. \end{aligned}$$
(6.4)

We want to study the problem of minimizing the functional \({\mathcal {E}}(\Omega ;\Sigma _D)\) among quasi-open sets of uniformly bounded measure. Therefore, fixing \(c>0\) we define

$$\begin{aligned} {\mathcal {O}}_c(\Sigma _D):=\inf \{{\mathcal {E}}(\Omega ;\Sigma _D); \ \ \Omega \ \hbox {quasi-open}, \ \Omega \subset \Sigma _D, \ |\Omega |\leqq c\}. \end{aligned}$$
(6.5)

Our aim is to give a sufficient condition on the cone \(\Sigma _D\) (hence on D) for the infimum in (6.5) to be achieved. We begin by recalling some known properties of the function \(u_\Omega \) that will be used in this section.

Proposition 6.4

Let \(c>0\). There exists a positive constant C depending only on N, \(\Sigma _D\) and c such that, for any quasi-open subset \(\Omega \) of \(\Sigma _D\) with \(|\Omega | \leqq c\), it holds that

  1. (i)

    \(u_\Omega \) is bounded and \(\Vert u_\Omega \Vert _{L^\infty (\Sigma _D)}\leqq C |\Omega |^{2/N}\);

  2. (ii)

    \(\int _{\Sigma _D} |\nabla u_\Omega |^2\ \mathrm{d}x \leqq C |\Omega |^{\frac{N+2}{N}}\);

  3. (iii)

    \(\int _{\Sigma _D} u_\Omega ^2\ \mathrm{d}x \leqq C |\Omega |^{\frac{N+4}{N}}\).

Proof

Let us fix \(c>0\). Since the cone \(\Sigma _D\) is a uniformly Lipschitz connected open set of \({\mathbb {R}}^N\) then we can apply [9, Lemma 2.5]. Hence, for any quasi-open subset \(\Omega \subset \Sigma _D\), with \(|\Omega |\leqq c\), fixing \(p \in ]N/2, +\infty [\) and taking \(f=\chi _\Omega \), where \(\chi _\Omega \) denotes the characteristic function of \(\Omega \), we obtain from [9, Lemma 2.5] that there exists a positive constant \({\tilde{C}}\) depending on N, p, \(\Sigma _D\) and c only such that

$$\begin{aligned} \Vert u_\Omega \Vert _{L^\infty (\Sigma _D)} \leqq {\tilde{C}} \Vert f\Vert _{L^p(\Sigma _D)} |\Omega |^{2/N - 1/p}= {\tilde{C}} |\Omega |^{1/p} |\Omega |^{2/N - 1/p} = {\tilde{C}}|\Omega |^{2/N}, \end{aligned}$$

which gives (i).

Next, taking \(u_\Omega \) as test function in the weak formulation of (6.1) we get

$$\begin{aligned} \int _\Omega |\nabla u_\Omega |^2 \ \mathrm{d}x = \int _\Omega u_\Omega \ \mathrm{d}x, \end{aligned}$$

and, by (i), we obtain

$$\begin{aligned} \int _\Omega |\nabla u_\Omega |^2 \ \mathrm{d}x \leqq C |\Omega |^{2/N} |\Omega |= C |\Omega |^{\frac{N+2}{N}}, \end{aligned}$$

i.e. (ii). Finally (iii) is a trivial consequence of (i) since

$$\begin{aligned} \int _{\Omega } u_\Omega ^2 \ \mathrm{d}x\leqq \Vert u_\Omega \Vert _{L^\infty (\Sigma _D)}^2 |\Omega |\leqq C^2 |\Omega |^{4/N}|\Omega |=C^2 |\Omega |^{\frac{N+4}{N}}. \end{aligned}$$

\(\square \)

Notice that as a straightforward consequence of the previous result it holds that \(\mathcal {O}_c(\Sigma _D)>-\infty \).

Remark 6.5

As remarked in [23, Remark 4.2] there is a natural invariance by scaling in our problem, which, in particular, allows to claim that the infimum as in (6.5), but with volume bounded by another constant \(\uplambda >0\), can be easily computed from \(\mathcal {O}_c(\Sigma _D)\). Namely we have

$$\begin{aligned} \uplambda ^{-\frac{N+2}{N}}\mathcal {O}_\uplambda (\Sigma _D)=c^{-\frac{N+2}{N}}\mathcal {O}_c(\Sigma _D)=\mathcal {O}_1(\Sigma _D). \end{aligned}$$
(6.6)

Indeed, for any quasi-open \(\Omega \subset \Sigma _D\), for any \(t>0\) it holds that \(t\Omega \subset \Sigma _D\), \(|t\Omega |=t^N|\Omega |\), and it is easy to check that \(u_{t\Omega }(x)=t^2 u_\Omega \left( \frac{x}{t}\right) \) and

$$\begin{aligned} {\mathcal {E}}(t\Omega ; \Sigma _D)=t^{N+2} {\mathcal {E}}(\Omega ; \Sigma _D). \end{aligned}$$
(6.7)

In particular \(\mathcal {O}_c(\Sigma _D)\) can be defined by taking \(|\Omega |=c\) in (6.5) and either a minimizer exists for any fixed volume or there are no minimizers whatever bound for the volume is chosen.

Among the quasi-open sets in \(\Sigma _D\) we can consider the spherical sectors

$$\begin{aligned} \Omega _{D,R}:=\Sigma _D\cap B_R(0). \end{aligned}$$
(6.8)

In this case the solution of (6.1) is radial and explicitly given by

$$\begin{aligned} u_{\Omega _{D,R}}(x)={\left\{ \begin{array}{ll} \frac{R^2-|x|^2}{2N} &{} \hbox {if} x\in \Omega _{D,R},\\ 0 &{} \hbox {if} x\in \Sigma _D\setminus {\Omega _{D,R}}, \end{array}\right. } \end{aligned}$$
(6.9)

and its energy is

$$\begin{aligned} {\mathcal {E}}(\Omega _{D,R}; \Sigma _D)=-\frac{1}{2N^2(N+2)} R^{N+2} {\mathcal {H}}_{N-1}(D). \end{aligned}$$
(6.10)

Therefore, by (6.5), we have

$$\begin{aligned} \mathcal {O}_c(\Sigma _D) \leqq {\mathcal {E}}(\Omega _{D,R_c}; \Sigma _D)<0, \end{aligned}$$
(6.11)

where \(R_c=R_c(D)>0\) is such that \(|\Omega _{D,R_c}|=R_c^{N} {\mathcal {H}}_{N-1}(D)=c\), namely \(R_c(D)=\left( \frac{c}{ {\mathcal {H}}_{N-1}(D)}\right) ^{\frac{1}{N}}\).

Remark 6.6

Notice that for any \(c>0\) it holds

$$\begin{aligned} {\mathcal {E}}(\Omega _{D,R_c(D)}; \Sigma _D)=-\frac{1}{2N^2(N+2)} c^{\frac{N+2}{N}} \left[ {\mathcal {H}}_{N-1}(D)\right] ^{-\frac{2}{N}} \end{aligned}$$
(6.12)

which means that \({\mathcal {E}}(\Omega _{D,R_c}; \Sigma _D)\) is monotone increasing with respect to \({\mathcal {H}}_{N-1}(D)\).

Remark 6.7

When D is a hemisphere, let us say for convenience the upper hemisphere, denoted by \({\mathbb {S}}^{N-1}_+={\mathbb {S}}^{N-1}\cap \{(x_1,\ldots ,x_N)\in {\mathbb {R}}^N; \ x_N>0\}\), then the cone \(\Sigma _{{\mathbb {S}}^{N-1}_+}\) coincides with the half-space \({\mathbb {R}}^N_+\). In this case it is well known, for example by symmetrization, that \(\mathcal {O}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) \) is achieved by any half-ball of measure c and

$$\begin{aligned} \mathcal {O}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right)= & {} {\mathcal {E}}\left( \Omega _{{\mathbb {S}}^{N-1}_+,R_c({\mathbb {S}}^{N-1}_+)}; \Sigma _{{\mathbb {S}}^{N-1}_+}\right) \nonumber \\= & {} -\frac{{\omega }_N}{4N(N+2)} \left( \frac{2c}{ N\omega _N}\right) ^{\frac{N+2}{N}}. \end{aligned}$$
(6.13)

In the general case, using the smoothness of the cone, we prove in Proposition 6.10 that it always holds

$$\begin{aligned} \mathcal {O}_c\left( \Sigma _D\right) \leqq \mathcal {O}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) . \end{aligned}$$
(6.14)

The main result of this section is to show that if the strict inequality holds in (6.14) then the infimum is achieved. Indeed we have

Theorem 6.8

Let \(c>0\) and assume that

$$\begin{aligned} \mathcal {O}_c\left( \Sigma _D\right) < \mathcal {O}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) . \end{aligned}$$
(6.15)

Then \(\mathcal {O}_c(\Sigma _D)\) is achieved.

Proof

Let \((\Omega _n)_n \subset \Sigma _D\) be a minimizing sequence for \(\mathcal {O}_c(\Sigma _D)\) and consider the corresponding energy functions \(u_{\Omega _n}\in H^1_0(\Omega _n;\Sigma _D)\) for any \(n\in {\mathbb {N}}\). By definition, we have

$$\begin{aligned} {\mathcal {E}}(\Omega _n; \Sigma _D)=-\frac{1}{2}\int _{\Omega _n} u_{\Omega _n} \ \mathrm {d}x \rightarrow {\mathcal {O}}_c(\Sigma _D), \ \text{ as }\ n\rightarrow +\infty . \end{aligned}$$

Setting \(u_n:=u_{\Omega _n}\), since \(|\Omega _n|\leqq c\), for any \(n\in {\mathbb {N}}\), by Proposition 6.4 we find a positive constant \(C_1\) independent of n such that

$$\begin{aligned} \Vert u_n\Vert _{H^1(\Sigma _D)} \leqq C_1 \ \ \forall n \in {\mathbb {N}}. \end{aligned}$$
(6.16)

In particular, up to a subsequence (still denoted by \((u_n)_n\)), we have \(\Vert u_n\Vert _{L^2(\Sigma _D)}^2\rightarrow \uplambda \), for some \(\uplambda \geqq 0\). We first observe that \(\uplambda >0\). Otherwise, if \(\uplambda =0\), by Hölder’s inequality and exploiting the uniform bound \(|\Omega _n|\leqq c\), we would have

$$\begin{aligned} \Vert u_n\Vert _{L^1(\Sigma _D)}\rightarrow 0, \ \text{ as }\ n\rightarrow +\infty , \end{aligned}$$
(6.17)

which implies that \({\mathcal {E}}(\Omega _n; \Sigma _D)\rightarrow 0\), as \(n\rightarrow +\infty \), contradicting \(\mathcal {O}_c(\Sigma _D)<0\) (see (6.11)). Therefore, as \((u_n)_n\) is bounded in \(H^1(\Sigma _D)\) and

$$\begin{aligned} \Vert u_n\Vert ^2_{L^2(\Sigma _D)} \rightarrow \uplambda ,\ \ \text{ for }\ n\rightarrow +\infty , \end{aligned}$$
(6.18)

for some \(\uplambda >0\), we can apply, with small modifications in the proof, the concentration-compactness principle of P. L. Lions (see [17, Lemma III.1]). Hence, there exists a subsequence \((u_{n_k})_k\) satisfying one of three following possibilities:

  1. (i)

    there exists \((y_{n_k})_k \subset \overline{\Sigma _D}\) satisfying

    $$\begin{aligned} \forall \varepsilon>0 \ \exists R>0 \ \ \hbox {such that} \ \ \int _{B_R(y_{n_k})\cap \Sigma _D} u_{n_k}^2 \ \mathrm{d}x \geqq \uplambda - \varepsilon \ \ \forall k\in {\mathbb {N}}; \end{aligned}$$
  2. (ii)

    \(\displaystyle \lim _{k\rightarrow +\infty } \sup _{y \in \Sigma _D} \int _{B_R(y)\cap \Sigma _D} u_{n_k}^2 \ \mathrm{d}x=0\), for all \(R>0\);

  3. (iii)

    there exists \(\alpha \in ]0,\uplambda [\) such that for all \(\varepsilon >0\), there exist \(k_0\geqq 1\) and two sequences \((u_{1,k})_k\), \((u_{2,k})_k\) bounded in \(H^1(\Sigma _D)\) satisfying, for \(k\geqq k_0\)

    $$\begin{aligned}&\Vert u_{n_k} - u_{1,k}-u_{2,k}\Vert _{L^2(\Sigma _D)} \leqq 4\varepsilon ,\\&\left| \int _{\Sigma _D} u_{1,k}^2 \ \mathrm {d}x -\alpha \right| \leqq \varepsilon , \ \ \left| \int _{\Sigma _D} u_{2,k}^2 \ \mathrm {d}x -(\uplambda -\alpha ) \right| \leqq \varepsilon ,\\&\text{ dist }(\text{ supp }(u_{1,k}), \text{ supp }(u_{2,k})) \rightarrow +\infty , \ \text{ as } k\rightarrow +\infty ,\\&\liminf _{k\rightarrow +\infty } \int _{\Sigma _D} |\nabla u_{n_k}|^2 - |\nabla u_{1,k}|^2 - |\nabla u_{2,k}|^2 \ \mathrm {d}x \geqq 0. \end{aligned}$$

We now divide the proof in some steps. We begin by showing that the “vanishing” case (ii) cannot occur.

Step 1: (ii) cannot happen.

Assume by contradiction that (ii) holds. The idea is to show that \(u_{n_k}\rightarrow 0\) strongly in \(L^2(\Sigma _D)\), as \(k\rightarrow +\infty \), contradicting (6.18). To prove this we invoke [28, Lemma 1.21] (whose proof can be easily adapted for functions in \(H^1(\Sigma _D)\)), which claims that (ii) and (6.16) imply \(u_{n_k}\rightarrow 0\) in \(L^p(\Sigma _D)\), for any \(2<p<2^*\), as \(k\rightarrow +\infty \), where \(2^*=\frac{2N}{N-2}\) is the critical Sobolev exponent. Then, exploiting that \(u_{n_k}\in H_0^1(\Omega _{n_k}; \Sigma _D)\) and \(|\Omega _{n_k}|\leqq c\), by Hölder’s inequality we readily conclude that \(u_{n_k}\rightarrow 0\) in \(L^2(\Sigma _D)\), as \(k\rightarrow +\infty \).

In the next step we prove that the “dichotomy” case (iii) cannot occur.

Step 2: (iii) cannot happen.

Assume by contradiction that (iii) holds. We claim that, up to a further subsequence, there exists another minimizing sequence \(({{\widetilde{\Omega }}}_{n_k})_k\subset \Sigma _D\), with \({{\widetilde{\Omega }}}_{n_k} \subset \Omega _{n_k}\), for any k, satisfying:

  • \({{\widetilde{\Omega }}}_{n_k} = \Omega _{1,k} \cup \Omega _{2,k}\), for some quasi-open subsets \(\Omega _{1,k}\), \(\Omega _{2,k}\) of \(\Omega _{n_k}\);

  • \(\mathrm {dist}( \Omega _{1,k}, \Omega _{2,k})\rightarrow +\infty \), as \(k\rightarrow +\infty \);

  • \(c_i:=\liminf _{k\rightarrow +\infty } |\Omega _{i,k}|>0\), for \(i=1,2\).

Indeed, by (iii) and a diagonal argument, we find bounded subsequences \((u_{1,k})_k\), \((u_{2,k})_k\) in \(H^1(\Sigma _D)\) (still indexed by k) satisfying

$$\begin{aligned} \begin{array}{lll} &{}{}&{}{}\displaystyle \Vert u_{n_k} - u_{1,k}-u_{2,k}\Vert _{L^2(\Sigma _D)} \rightarrow 0,\\ &{}{}&{}{}\displaystyle \int _{\Sigma _D} u_{1,k}^2 \ \mathrm {d}x \rightarrow \alpha , \ \ \int _{\Sigma _D} u_{2,k}^2 \ \mathrm {d}x \rightarrow (\uplambda -\alpha ), \ \text{ as }\ k\rightarrow +\infty ,\\ &{}{}&{}{}\displaystyle \text{ dist }(\text{ supp }(u_{1,k}), \text{ supp }(u_{2,k})) \rightarrow +\infty , \ \text{ as }\ k\rightarrow +\infty ,\\ &{}{}&{}{}\displaystyle \liminf _{k\rightarrow +\infty } \int _{\Sigma _D} |\nabla u_{n_k}|^2 - |\nabla u_{1,k}|^2 - |\nabla u_{2,k}|^2 \ \mathrm {d}x \geqq 0. \end{array} \end{aligned}$$
(6.19)

By the proof of [17, Lemma III.1] we see that \(u_{1,k}\), \(u_{2,k}\) can be chosen to be non-negative and in addition, since \(u_{n_k} \in H^1_0(\Omega _{n_k};\Sigma _D)\), we also have that \(u_{1,k}, u_{2,k} \in H^1_0(\Omega _{n_k};\Sigma _D)\) for any k. In particular, as \(u_{1,k}, u_{2,k} \in H^1(\Sigma _D)\), setting \(\Omega _{1,k}:=\{u_{1,k}>0\}\), \(\Omega _{2,k}:=\{u_{2,k}>0\}\) it follows that \(\Omega _{1,k}\), \(\Omega _{2,k}\) are quasi-open subsets of \(\Sigma _D\). Therefore, \({{\widetilde{\Omega }}}_{k}:= \Omega _{1,k} \cup \Omega _{2,k}\) is a quasi-open set contained in \(\Omega _{n_k}\) and denoting by \({\tilde{u}}_{n_k}:=u_{{{\widetilde{\Omega }}}_{n_k}}\) the torsion function of \({{\widetilde{\Omega }}}_{n_k}\) and arguing as in [6, Sect. 3.3] (with obvious small modifications), we infer that

$$\begin{aligned} \Vert u_{n_k} - {\tilde{u}}_{n_k}\Vert _{H^1(\Sigma _D)}\rightarrow 0, \ \ \text{ as }\ k\rightarrow +\infty . \end{aligned}$$
(6.20)

From (6.20) it follows that \(({{\widetilde{\Omega }}}_{n_k})_k\) is a minimizing sequence for \({\mathcal {O}}_c(D)\). Moreover, by construction and (6.19) we readily deduce that \(\mathrm {dist}( \Omega _{1,k}, \Omega _{2,k})\rightarrow +\infty \), as \(k\rightarrow +\infty \). Finally, setting

$$\begin{aligned} c_i:=\liminf _{n\rightarrow +\infty } |\Omega _{i,k}|, \ \ i=1,2, \end{aligned}$$
(6.21)

it holds that \(c_i>0\) for \(i=1,2\). Indeed, assuming by contradiction, for instance, that \(c_1=0\), by Hölder’s inequality and Sobolev’s inequality (note that \(\Sigma _D\) satisfies the cone condition) we get

$$\begin{aligned} \int _{\Sigma _D} u_{1,k}^2 \ \mathrm{d}x \leqq \left( \int _{\Sigma _D} |u_{1,k}|^{2^*} \ \mathrm{d}x \right) ^{\frac{2}{2^*}} |\Omega _{1,k}|^{\frac{2}{N}} \leqq C(N,\Sigma _D) \int _{\Sigma _D} |\nabla u_{1,k}|^2 \ \mathrm{d}x \ |\Omega _{1,k}|^{\frac{2}{N}}. \end{aligned}$$

Now, recalling that \((u_{1,k})_k\) is a bounded sequence in \(H^1(\Sigma _D)\), from the previous inequality and since we are assuming \(c_1=0\) we deduce that

$$\begin{aligned} \liminf _{k\rightarrow +\infty } \int _{\Sigma _D} u_{1,k}^2 \ \mathrm{d}x=0, \end{aligned}$$

which contradicts (6.19). Hence \(c_1>0\), and by the same argument we infer that \(c_2>0\). The proof of the claim is complete.

In order to conclude the proof of Step 2 we show that the previous claim leads to a contradiction. To this end we begin observing that by invariance by dilatation (see Remark 6.5) it follows that our minimization problem is equivalent to

$$\begin{aligned} {\mathcal {M}}(\Sigma _D):=\inf \left\{ \frac{{\mathcal {E}}(\Omega ; \Sigma _D)}{|\Omega |^{\frac{N+2}{N}}}; \ \ \Omega \ \hbox {quasi-open}, \ \Omega \subset \Sigma _D,\ |\Omega |>0\right\} . \end{aligned}$$

In particular by (6.6) and a straightforward computation we check that

$$\begin{aligned} \mathcal {O}_c(\Sigma _D)=c^{\frac{N+2}{N}}\mathcal {O}_1(\Sigma _D)=c^{\frac{N+2}{N}} {\mathcal {M}}(\Sigma _D), \end{aligned}$$

and a minimizing sequence for \({\mathcal {M}}(\Sigma _D)\) is given by \(\Lambda _k:=|{{\widetilde{\Omega }}}_{n_k} |^{-\frac{1}{N}}{{\widetilde{\Omega }}}_{n_k}\). Then, setting \(\Lambda _{i,k}:=|{{\widetilde{\Omega }}}_{n_k} |^{-\frac{1}{N}}\Omega _{i,k}\), \(i=1,2\), and in view of the previous claim, up to a subsequence, we have \(\Lambda _k= \Lambda _{1,k} \cup \Lambda _{2,k}\), where \(|\Lambda _{i,k}|\rightarrow \frac{c_i}{c_1+c_2}>0\), as \(k\rightarrow +\infty \), \(i=1,2\), and \(\Lambda _{1,k} \cap \Lambda _{2,k}=\emptyset \) for all sufficiently large k. Now, as \(\Lambda _{i,k} \subset \Sigma _D\) are quasi-open subsets with positive measure, then by definition of \({\mathcal {M}}(\Sigma _D)\), for any \(i=1,2\), we have

$$\begin{aligned} \frac{{\mathcal {E}}(\Lambda _{i,k}; \Sigma _D)}{|\Lambda _{i,k}|^{\frac{N+2}{N}}}\geqq {\mathcal {M}}(\Sigma _D). \end{aligned}$$
(6.22)

In addition, assuming without loss of generality that \(|\Lambda _{1,k}|\geqq |\Lambda _{2,k}|\), by an elementary computation, we deduce that

$$\begin{aligned} \displaystyle \left( |\Lambda _{1,k}|+|\Lambda _{2,k}|\right) ^{\frac{N+2}{N}}= & {} \displaystyle |\Lambda _{1,k}|^{\frac{N+2}{N}}+\frac{N+2}{N} |\Lambda _{1,k}|^{\frac{2}{N}}|\Lambda _{2,k}| \nonumber \\&\quad +\frac{N+2}{N^2} (|\Lambda _{1,k}|+\xi _k)^{\frac{2-N}{N}} |\Lambda _{2,k}|^2 \nonumber \\\geqq & {} \displaystyle |\Lambda _{1,k}|^{\frac{N+2}{N}}+|\Lambda _{2,k}|^{\frac{N+2}{N}}+\frac{N+2}{N^2} (|\Lambda _{1,k}|+\xi _k)^{\frac{2-N}{N}} |\Lambda _{2,k}|^2,\nonumber \\ \end{aligned}$$
(6.23)

where \(\xi _k \in [0,|\Lambda _{2,k}|]\). Then, setting \({\mathcal {M}}_k(\Sigma _D):=\frac{{\mathcal {E}}(\Lambda _k; \Sigma _D)}{|\Lambda _k|^{\frac{N+2}{N}}}<0\), recalling that \((\Lambda _k)_k\) is minimizing for \({\mathcal {M}}(\Sigma _D)\), exploiting the properties of \(\Lambda _k\) and taking into account (6.23), (6.22), we infer that for all sufficiently large k it holds

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathcal {E}}(\Lambda _{k}; \Sigma _D)&{}{}=&{}{} \displaystyle {\mathcal {M}}_k(\Sigma _D)|\Lambda _k|^{\frac{N+2}{N}}\\ &{}{}=&{}{} \displaystyle \left( {\mathcal {M}}(\Sigma _D)+o(1)\right) \left( |\Lambda _{1,k}|+|\Lambda _{2,k}|\right) ^{\frac{N+2}{N}}\\ &{}{}\leqq &{}{} \displaystyle \left( {\mathcal {M}}(\Sigma _D)+o(1)\right) \left( |\Lambda _{1,k}|^{\frac{N+2}{N}}+|\Lambda _{2,k}|^{\frac{N+2}{N}}+\frac{N+2}{N^2} \xi _k^{\frac{2-N}{N}} |\Lambda _{2,k}|^2\right) \\ &{}{}=&{}{} \displaystyle {\mathcal {M}}(\Sigma _D) |\Lambda _{1,k}|^{\frac{N+2}{N}}\\ &{}{}&{}{}\ + {\mathcal {M}}(\Sigma _D)|\Lambda _{2,k}|^{\frac{N+2}{N}}+{\mathcal {M}}(\Sigma _D)\frac{N+2}{N^2}(|\Lambda _{1,k}|+\xi _k)^{\frac{2-N}{N}} |\Lambda _{2,k}|^2 + o(1)\\ &{}{}\leqq &{}{} \displaystyle {\mathcal {E}}(\Lambda _{1,k}; \Sigma _D)+ {\mathcal {E}}(\Lambda _{2,k}; \Sigma _D)\\ &{}{}&{}{} +{\mathcal {M}}(\Sigma _D)\frac{N+2}{N^2}(|\Lambda _{1,k}|+\xi _k)^{\frac{2-N}{N}} |\Lambda _{2,k}|^2 + o(1)\\ &{}{}=&{}{} \displaystyle {\mathcal {E}}(\Lambda _{k}; \Sigma _D)+{\mathcal {M}}(\Sigma _D)\frac{N+2}{N^2}(|\Lambda _{1,k}|+\xi _k)^{\frac{2-N}{N}} |\Lambda _{2,k}|^2 + o(1)\\ &{}{}<&{}{}\displaystyle {\mathcal {E}}(\Lambda _{k}; \Sigma _D) \end{array}\nonumber \\ \end{aligned}$$
(6.24)

where the last inequality is strict because \(|\Lambda _{i,k}|\rightarrow \frac{c_i}{c_1+c_2}>0\), for \(i=1,2\), \(k\rightarrow +\infty \), and \({\mathcal {M}}(\Sigma _D)<0\). Clearly (6.24) is contradictory and this concludes the proof of Step 2.

From Step 1 and Step 2 we know that the only admissible case is (i). Roughly speaking (i) states that, there exists a sequence \((y_{n_k})_k \subset \overline{\Sigma _D}\) such that for a sufficiently large ball \(B_R(y_{n_k})\), the mass of \(u_{n_k}\) is concentrated in \(B_R(y_{n_k})\cap \Omega _{n_k}\), while the part in the complement \(B_R^\complement (y_k)\cap \Omega _{n_k}\) is negligible. With (i) at hand we can show that the same happens for the energy, in particular the possible tails of \(\Omega _{n_k}\) do not play a role. This is the content of the next technical step.

Step 3: For any fixed \(\varepsilon >0\) there exist \({\bar{R}}>1\) and \({\bar{k}} \in {\mathbb {N}}\), both depending only on \(\varepsilon \), such that

$$\begin{aligned} {\mathcal {E}}(\Omega _{n_k}; \Sigma _D)\geqq {\mathcal {E}}(B_{2 R}(y_{n_k})\cap \Omega _{n_k}; \Sigma _D) - 2c\sqrt{2\varepsilon }, \ \ \forall k\geqq {\bar{k}} \ \forall R\geqq {\bar{R}}.\nonumber \\ \end{aligned}$$
(6.25)

Let us fix \(\varepsilon >0\) and let \(R>0\) be tha radius given by (i). Let \(\varphi \in C^\infty _c({\mathbb {R}}^N)\) such that \(0\leqq \varphi \leqq 1\), \(\varphi \equiv 1\) in \(B_{R}(0)\), \(\varphi \equiv 0\) in \(B_{2R}^\complement (0)\) and \(|\nabla \varphi |\leqq \frac{C_0}{R}\) in \({\mathbb {R}}^N\), where \(C_0>0\) is a constant independent of R. We set \(\varphi _k(x):=\varphi (x-y_{n_k})\) and observe that

$$\begin{aligned} \begin{array}{lll} \displaystyle \int _{\Sigma _D} |\nabla u_{n_k}|^2 \ \mathrm {d}x &{}{}\geqq &{}{}\displaystyle \int _{\Sigma _D} |\nabla u_{n_k}|^2\varphi _k^2 \ \mathrm {d}x\\ &{}{}=&{}{}\displaystyle \int _{\Sigma _D} |\nabla (u_{n_k}\varphi _k)|^2 \ \mathrm {d}x\\ &{}{}&{}{}\displaystyle - \underbrace{2 \int _{\Sigma _D} u_{n_k} \varphi _k \nabla u_{n_k} \mathbf {\cdot } \nabla \varphi _k \ \mathrm {d}x}_\mathbf {(I) } - \underbrace{\int _{\Sigma _D} u_{n_k}^2 |\nabla \varphi _k |^2 \ \mathrm {d}x}_\mathbf {(II) }. \end{array} \end{aligned}$$
(6.26)

Then, exploiting the properties of \(\varphi _k\), Hölder’s inequality and (6.16) we infer that

$$\begin{aligned} |\mathbf (I) | \leqq \frac{C_0}{R} \Vert \nabla u_{n_k}\Vert _{L^2(\Sigma _D)} \Vert u_{n_k}\Vert _{L^2(\Sigma _D)} \leqq \frac{C_0C_1^2}{R}, \end{aligned}$$

where \(C_0\), \(C_1\) are both independent of k and R. Similarly, for \(|\mathbf (II) |\) we have

$$\begin{aligned} |\mathbf (II) | \leqq \frac{C_0^2C_1}{R^2}, \end{aligned}$$

and thus by (6.26) and assuming without loss of generality that \(R>1\) we obtain that

$$\begin{aligned} \int _{\Sigma _D} |\nabla u_{n_k}|^2 \ \mathrm{d}x \geqq \int _{\Sigma _D} |\nabla (u_{n_k}\varphi _k)|^2 \ \mathrm{d}x- \frac{C_2}{R}, \end{aligned}$$
(6.27)

where \(C_2>0\) is independent of k and R. In addition, let us write

$$\begin{aligned} -\int _{\Sigma _D} u_{n_k} \ \mathrm{d}x = -\int _{\Sigma _D} u_{n_k}\varphi _k \ \mathrm{d}x - \underbrace{\int _{\Sigma _D\cap B_R(y_k)} u_{n_k}(1-\varphi _k) \ \mathrm{d}x}_\mathbf{(III) } - \underbrace{\int _{\Sigma _D\cap B_R^\complement (y_k)} u_{n_k}(1-\varphi _k) \ \mathrm{d}x}_\mathbf{(IV) }. \end{aligned}$$

We first observe that \(\mathbf (III) =0\), because \(\varphi _k\equiv 1\) in \(B_R(y_k)\), while for (IV), applying Hölder’s inequality, taking into account that \(u_{n_k}=0\) q.e. on \(\Sigma \setminus \Omega _{n_k}\) and the properties of \(\varphi _k\), we get that

$$\begin{aligned} |\mathbf (IV) | \leqq \left( \int _{\Sigma _D\cap B_R^\complement (y_k)} u_{n_k}^2 \ \mathrm{d}x\right) ^{\frac{1}{2}} |B_{2R}(y_k)\cap \Omega _{n_k}|. \end{aligned}$$

Now, thanks to (i) and (6.18) it follows that \(\Vert u_{n_k}\Vert _{L^2(\Sigma _D\cap B_R^\complement (y_k))} \leqq \sqrt{2\varepsilon }\) for all sufficiently large k, and thus, as \(|\Omega _{n_k}|\leqq c\), we deduce that

$$\begin{aligned} |\mathbf (IV) | \leqq \sqrt{2\varepsilon } c. \end{aligned}$$

Summing up, we have proved that

$$\begin{aligned} -\int _{\Sigma _D} u_{n_k} \ \mathrm{d}x \geqq -\int _{\Sigma _D} u_{n_k}\varphi _k \ \mathrm{d}x - \sqrt{2\varepsilon } c. \end{aligned}$$
(6.28)

Hence, combining (6.27), (6.28) and recalling the definition of the functional J (see (6.2)) we obtain

$$\begin{aligned} J(u_{n_k})\geqq J(u_{n_k}\varphi _k) - \frac{C_2}{2R}-\sqrt{2\varepsilon } c. \end{aligned}$$

Since \(\varepsilon \) is fixed and \(C_2\) is independent of R and k, up to taking a larger R, we can assume that \(\frac{C_2}{2R}<\sqrt{2\varepsilon } c\). Then, observing that \(u_{n_k}\varphi _k \in H_0^1(B_{2R}(y_{n_k})\cap \Omega _{n_k}; \Sigma _D)\) we finally get

$$\begin{aligned} {\mathcal {E}}(\Omega _{n_k}; \Sigma _D)=J(u_{n_k})\geqq J(u_{n_k}\varphi _k) - 2\sqrt{2\varepsilon } c \geqq {\mathcal {E}}(B_{2R}(y_{n_k})\cap \Omega _{n_k}; \Sigma _D) - 2\sqrt{2\varepsilon } c, \end{aligned}$$

so that Step 3 is proved.

In the next step we prove that the sequence of points \((y_{n_k})_k\subset {\overline{\Sigma }}_D\) provided by (i) is bounded.

Step 4: The sequence \((y_{n_k})_k\subset {\overline{\Sigma }}_D\) existing by (i) is bounded.

Assume by contradiction that there exists a subsequence (still indexed by k) such that

$$\begin{aligned} \lim _{k\rightarrow +\infty }|y_{n_k}|= +\infty . \end{aligned}$$

Thanks to the assumption (6.15) we can fix \(\varepsilon >0\) sufficiently small so that

$$\begin{aligned} {\mathcal {O}}_c(\Sigma _D)+2c\sqrt{2\varepsilon }<{\mathcal {O}}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) , \end{aligned}$$
(6.29)

and by Step 3 we find R sufficiently large depending only on \(\varepsilon \), such that for all sufficiently large k

$$\begin{aligned} {\mathcal {E}}(\Omega _{n_k}; \Sigma _D)\geqq {\mathcal {E}}(B_{2R}(y_{n_k})\cap \Omega _{n_k}; \Sigma _D) - 2c\sqrt{2\varepsilon }. \end{aligned}$$
(6.30)

We observe that \(B_{2R}(y_{n_k})\cap \Omega _{n_k}\) intersects the boundary of \(\Sigma _D\). More precisely, for all sufficiently large k, it holds that

$$\begin{aligned} {\mathcal {H}}_{N-1}\left( \overline{(B_{2R}(y_{n_k})\cap \Omega _{n_k})}\cap \partial \Sigma _D\right) >0. \end{aligned}$$
(6.31)

Indeed, on the contrary, setting for convenience \(\Theta _{R,k}:=B_{2R}(y_{n_k})\cap \Omega _{n_k}\) there exists a subsequence (still indexed by k) such that \({\mathcal {H}}_{N-1}(\overline{\Theta _{R,k}}\cap \partial \Sigma _D)=0\) for all \(k\in {\mathbb {N}}\), and by the same argument of [9, Remark 4.3] we conclude that \(H_0^1(\Theta _{R,k};\Sigma _D)= H_0^1(\Theta _{R,k})\), and thus

$$\begin{aligned} {\mathcal {E}}( \Theta _{R,k}; \Sigma _D)={\mathcal {E}}(\Theta _{R,k}; {\mathbb {R}}^N), \end{aligned}$$
(6.32)

where \({\mathcal {E}}(\Theta _{R,k}; {\mathbb {R}}^N)\) denotes the “free” energy of \(\Theta _{R,k}\), under a homogeneous Dirichlet condition, namely, \({\mathcal {E}}(\Theta _{R,k}; {\mathbb {R}}^N)\) is the minimizer in \(H_0^1(\Theta _{R,k})\) of the functional \(J(v)=\frac{1}{2}\int _{{\mathbb {R}}^N} |\nabla v|^2 \ \mathrm{d}x - \int _{{\mathbb {R}}^N} v \ \mathrm{d}x\). Then, by considering the Schwartz symmetrization \(u_{\Theta _{R,k}}^*\) of the energy function \(u_{\Theta _{R,k}}\) associated to \(\Theta _{R,k}\), and thanks to the classical Pólya-Szegö inequality, we infer that

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathcal {E}}( \Theta _{R,k}; {\mathbb {R}}^N)&{}{}=&{}{}\displaystyle \frac{1}{2}\int _{\Theta _{R,k}} |\nabla u_{\Theta _{R,k}}|^2 \ \mathrm {d}x - \int _{ \Theta _{R,k}} u_{\Theta _{R,k}} \ \mathrm {d}x\\ &{}{} \geqq &{}{}\displaystyle \frac{1}{2}\int _{\Theta _{R,k}^*} |\nabla u_{\Theta _{R,k}}^*|^2 \ \mathrm {d}x - \int _{ \Theta _{R,k}^*} u_{\Theta _{R,k}}^* \ \mathrm {d}x\\ &{}{} \geqq &{}{}\displaystyle {\mathcal {E}}(\Theta _{R,k}^*; {\mathbb {R}}^N), \end{array} \end{aligned}$$
(6.33)

Hence, as \(\Theta _{R,k}^*\) is a ball, with \(c_k:=|\Theta _{R,k}^*|=|\Theta _{R,k}|\leqq c\), then from (6.32),(6.33), taking into account Remark 6.6 and (6.13) (noticing that \({\mathcal {E}}(\Theta _{R,k}^*; {\mathbb {R}}^N)={\mathcal {E}}(\Omega _{{\mathbb {S}}^{N-1},R_{c_k}({\mathbb {S}}^{N-1})}; \Sigma _{{\mathbb {S}}^{N-1}})\), where \(\Omega _{{\mathbb {S}}^{N-1},R_{c_k}({\mathbb {S}}^{N-1})}\) is the ball centred at the origin of radius \(R_{c_k}({\mathbb {S}}^{N-1})\) with \(|\Omega _{{\mathbb {S}}^{N-1},R_{c_k}({\mathbb {S}}^{N-1})}|=c_k\), see (6.8), (6.12)), we deduce that

$$\begin{aligned} {\mathcal {E}}(\Theta _{R,k}; \Sigma _D) \geqq {\mathcal {E}}(\Theta _{R,k}^*; {\mathbb {R}}^N)> \mathcal {O}_{c_k}\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) \geqq {\mathcal {O}}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) . \end{aligned}$$
(6.34)

Finally, recalling that \(\Theta _{R,k}=B_{2R}(y_{n_k})\cap \Omega _{n_k}\), then from (6.25) and (6.34) we have, for large k,

$$\begin{aligned} {\mathcal {E}}(\Omega _{n_k}; \Sigma _D)\geqq {\mathcal {O}}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) - 2c\sqrt{2\varepsilon }. \end{aligned}$$

Hence passing to the limit as \(k\rightarrow +\infty \) we conclude that

$$\begin{aligned} \mathcal {O}_c(\Sigma _D)\geqq {\mathcal {O}}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) - 2c\sqrt{2\varepsilon }, \end{aligned}$$

which contradicts (6.29).

Then, by (6.31), there exists \(k_0\in {\mathbb {N}}\) such that \(\mathrm {dist}(y_{n_k}, \partial \Sigma _D)\leqq 2R\) for all \(k\geqq k_0\) and we can find a sequence of points \((z_k)_k \subset \partial \Sigma _D\setminus \{0\}\) such that \(z_k \in (\overline{B_{2R}(y_{n_k})\cap \Omega _{n_k}})\cap \partial \Sigma _D\) and \(B_{2R}(y_{n_k})\subset B_{4R}(z_k)\), for all \(k\geqq k_0\). Then, by monotonicity of the torsional energy \({\mathcal {E}}\) with respect to the set inclusion, noticing that \(H_0^1(B_{2R}(y_{n_k})\cap \Omega _{n_k}; \Sigma _D)\subset H_0^1(B_{4R}(z_k)\cap \Omega _{n_k}; \Sigma _D)\), we have

$$\begin{aligned} {\mathcal {E}}(B_{2R}(y_{n_k})\cap \Omega _{n_k}; \Sigma _D)\geqq {\mathcal {E}}(B_{4R}(z_k)\cap \Omega _{n_k}; \Sigma _D), \end{aligned}$$
(6.35)

for all \(k\geqq k_0\). Clearly, by construction, \(|B_{4R}(z_k)\cap \Omega _{n_k}| \leqq c\) for all \(k\geqq k_0\) and \(|z_k|\rightarrow +\infty \), as \(k\rightarrow +\infty \). We claim that, up to a further subsequence (still indexed by k) it holds

$$\begin{aligned} {\mathcal {E}}(B_{4R}(z_k)\cap \Omega _{n_k}; \Sigma _D) \geqq \mathcal {O}_c\left( \Sigma _{{\mathbb {S}}_+^{N-1}}\right) +o(1), \end{aligned}$$
(6.36)

for all sufficiently large k, where \(o(1)\rightarrow 0\) as \(k\rightarrow +\infty \).

Notice that if Claim (6.36) holds then the proof of Step 4 is complete. Indeed combining (6.30), (6.35) and (6.36), up to a subsequence, we have

$$\begin{aligned} {\mathcal {E}}(\Omega _{n_k}; \Sigma _D)\geqq \mathcal {O}_c(\Sigma _{{\mathbb {S}}^{N-1}_+}) +o(1)-2c\sqrt{2\varepsilon }, \end{aligned}$$

for all sufficiently large k. Then, passing to the limit as \(k\rightarrow +\infty \) we get

$$\begin{aligned} {\mathcal {O}}_c(\Sigma _D)\geqq {\mathcal {O}}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) -2c\sqrt{2\varepsilon }, \end{aligned}$$

but this contradicts (6.29).

Proof of Claim (6.36): We first observe that since \(\partial \Sigma _D\setminus \{0\}\) is a smooth hypersurface, then, for any fixed \(q\in \partial D \subset \partial \Sigma _D\setminus \{0\}\) there exists an open neighborhood V of q in \(\partial \Sigma _D\setminus \{0\}\) such that \(V-q\) is the graph over \(T_q\partial \Sigma _D\) of a smooth function \(g: U\rightarrow {\mathbb {R}}\), with \(g(0)=0\), where U is an open neighborhood of the origin in \(T_q\partial \Sigma _D\) (without loss of generality we can assume that U is a ball and g is smooth in \({\overline{U}}\)). Namely, fixing a orthonormal base \({\mathcal {B}}^\prime :=\{v_1,\ldots ,v_{N-1}\}\) of \(T_q\partial \Sigma _D\) and choosing \({\mathcal {B}}:=\{v_1,\ldots ,v_{N-1},-\nu (q)\}\) as orthonormal base of \({\mathbb {R}}^N\), where \(-\nu (q)\) is the inner unit normal of \(\partial \Sigma _D\) at q, denoting by \(x^\prime =(x_1^\prime ,\ldots ,x^\prime _{N-1})\) the coordinates of the points in \(T_q\partial \Sigma _D\) with respect to \({\mathcal {B}}^\prime \) and by \(x=(x^\prime ,x_N)\) the coordinates in \({\mathbb {R}}^N\) with respect to \({\mathcal {B}}\), we can identify

$$\begin{aligned} V-q=\{(x^\prime ,x_N) \in {\mathbb {R}}^N; \ x^\prime =(x_1^\prime ,\ldots ,x_{N-1}^\prime )\in U, \ x_N=g(x_1^\prime ,\ldots , x_{N-1}^\prime )\}, \end{aligned}$$

where U is the orthogonal projection of \(V-q\) onto \(T_q\partial \Sigma _D\). To be precise, if \(\varphi \) is a local parametrization centered at q, i.e. \(\varphi (0)=q\), by writing \(\varphi -q=\sum _{i=1}^{N-1} [(\varphi -q) \varvec{\cdot } v_i] v_i +(\varphi -q) \varvec{\cdot } (-\nu (q))\), and since \(\sum _{i=1}^{N-1} [(\varphi -q) \varvec{\cdot } v_i] v_i\) is a locally invertible map from an open neighborhood of the origin in \({\mathbb {R}}^{N-1}\) to an open neighborhood of the origin in \(T_q\partial \Sigma _D\cong {\mathbb {R}}^{N-1}\), then, denoting by G its local inverse and taking \(g(x^\prime ):=[(\varphi -q)\circ G(x^\prime )] \varvec{\cdot } (-\nu (q))\) we obtain the desired map. In particular, notice that since \(\frac{\partial [(\varphi -q)\circ G]}{\partial x^\prime _i}(0)\in T_q\partial \Sigma _D\) it follows that \(\frac{\partial g}{\partial x_i^\prime }(0)=0\), for any \(i=1,\ldots ,N-1\).

Now, since \(\partial \Sigma _D\) is a cone it follows that for any \(t>0\), \(T_{tq}\partial \Sigma _D=T_q\partial \Sigma _D\), \(\nu (tq)=\nu (q)\) and \(tV-tq\) is the graph over \(T_q\partial \Sigma _D\) of the map \(g_t:tU\rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} g_t(x^\prime ):=tg\left( \frac{x^\prime }{t}\right) ,\ x^\prime \in tU. \end{aligned}$$
(6.37)

For any \(x^\prime \in tU\), for any \(i=1,\ldots ,N-1\), we have

$$\begin{aligned} \frac{\partial g_t}{\partial x^\prime _i}(x^\prime )= \frac{\partial g}{\partial x^\prime _i}\left( \frac{x^\prime }{t}\right) =\frac{\partial g}{\partial x^\prime _i}\left( 0\right) + \frac{1}{t} \nabla _{x^\prime } \left[ \frac{\partial g}{\partial x^\prime _i}\right] \left( \frac{\xi }{t}\right) \varvec{\cdot } x^\prime , \end{aligned}$$
(6.38)

where \(\xi =\xi (x^\prime ,t)\) belongs to the segment joining 0 and \(\frac{x^\prime }{t}\), and \(\nabla _{x^\prime }\) denotes the gradient with respect to \(x_1^\prime ,\ldots ,x_{N-1}^\prime \). Hence, for any fixed ball \(B_{R_1}(0)\subset T_q\partial \Sigma _D\), for all \(t>0\) sufficiently large such that \(B_{R_1}(0) \subset tU\), recalling that \(\frac{\partial g}{\partial x^\prime _i}\left( 0\right) =0\), we have

$$\begin{aligned} \max _{x^\prime \in \overline{B_{R_1}(0)}}{|\nabla _{x^\prime } g_t(x^\prime )|}\leqq \frac{1}{t} \max _{x^\prime \in {\overline{U}}} \sqrt{\sum _{i=1}^{N-1}\left| \nabla _{x^\prime } \left[ \frac{\partial g}{\partial x^\prime _i}\right] (x^\prime )\right| ^2}R_1\leqq \frac{C}{t}, \end{aligned}$$
(6.39)

where C is independent of t, and, in particular,

$$\begin{aligned} \lim _{t\rightarrow +\infty } \max _{x^\prime \in \overline{B_{R_1}(0)}}{|\nabla _{x^\prime } g_t(x^\prime )|}=0. \end{aligned}$$
(6.40)

Let \(C^+_{B_{R_1}(0)}\), \(E^+\left( g_t\big |_{B_{R_1}(0)}\right) \) be, respectively, the upper cylinder generated by \(B_{R_1}(0)\) and the epigraph associated to \(g_t\big |_{B_{R_1}(0)}\), namely

$$\begin{aligned} \begin{array}{lll} \displaystyle C^+_{B_{R_1}(0)}&{}:=&{}\displaystyle \{(x^\prime ,x_N)\in {\mathbb {R}}^N; \ x^\prime =(x_1^\prime ,\ldots ,x_{N-1}^\prime )\in B_{R_1}(0),\ x_N>0\},\\ \displaystyle E^+\left( g_t\big |_{B_{R_1}(0)}\right) &{}:=&{}\displaystyle \{(x^\prime ,x_N)\in {\mathbb {R}}^N; \ x^\prime =(x_1^\prime ,\ldots ,x_{N-1}^\prime )\in B_{R_1}(0),\\ &{}&{}\quad x_N>g_t(x_1^\prime ,\ldots ,x_{N-1}^\prime )\}. \end{array} \end{aligned}$$
(6.41)

Then, the map \(F_t: \overline{C^+_{B_{R_1}(0)}} \rightarrow \overline{E^+\left( g_t\big |_{B_{R_1}(0)}\right) }\) defined by

$$\begin{aligned} F_t(x^\prime ,x_N)=(x^\prime , x_N+g_t(x^\prime )), \ (x^\prime ,x_N)\in \overline{C^+_{B_{R_1}(0)}}, \end{aligned}$$
(6.42)

is a diffeomorphism whose Jacobian matrix is of the form

$$\begin{aligned} \mathrm {Jac}(F_t)(x^\prime ,x_N)=\left[ \begin{array}{cc}{\mathbb {I}}_{N-1}&{}0_{N-1}^T\\ \nabla _{x^\prime } g_t(x^\prime )&{}1 \end{array}\right] , \end{aligned}$$
(6.43)

where \({\mathbb {I}}_{N-1}\) is the identity matrix of order \(N-1\), \(0_{N-1}\) is the null vector in \({\mathbb {R}}^{N-1}\), T is the transposition. Notice also that \(\mathrm {Jac}(F_t)\) is independent of \(x_N\) and in view of (6.40) it holds that

$$\begin{aligned} \lim _{t\rightarrow +\infty }\Vert \mathrm {Jac}(F_t)-{\mathbb {I}}_{N} \Vert _{C^0\left( \overline{C^+_{B_{R_1}(0)}}\right) }= 0. \end{aligned}$$
(6.44)

Now, let us consider the sequence \((q_k)_k\subset \partial D\), where \(q_k:=\frac{z_k}{|z_k|}\), and \((z_k)_k\subset \partial \Sigma _D\setminus \{0\}\) is the sequence appearing in Claim (6.36). Since \(\partial D\) is a compact subset of \({\mathbb {S}}^{N-1}\) we deduce that, up to a subsequence (still indexed by k) it holds that \(\text {dist}_{{\mathbb {S}}^{N-1}}(q_k, {\bar{q}})\rightarrow 0\), as \(k\rightarrow +\infty \), for some \({\bar{q}}\in \partial D\), where \(\mathrm {dist}_{{\mathbb {S}}^{N-1}}\) denotes the geodesic distance in \({\mathbb {S}}^{N-1}\). Then, from the previous discussion there exist an open neighborhood \(V_1\) of \({\bar{q}}\) in \(\partial \Sigma _D\setminus \{0\}\), a convex open neighborhood \(U_1\) of the origin in \(T_{{\bar{q}}}\partial \Sigma _D\) and a smooth function \(g_1:\overline{U_1}\rightarrow {\mathbb {R}}\) such that \(g_1(0)=0\), \(\nabla _{x^\prime }g_1(0)=0\), and \(V_1-{\bar{q}}\) is the graph over \(T_{{\bar{q}}}\partial \Sigma _D\) associated to \(g_1\big |_{U_1}\), where \(x^\prime =(x_1^\prime ,\ldots ,x_{N-1}^\prime )\) are the coordinates with respect to fixed orthonormal base \(\{{\bar{v}}_1,\ldots ,{\bar{v}}_{N-1}\}\) of \(T_{{\bar{q}}}\partial \Sigma _D\). Since \(\text {dist}_{{\mathbb {S}}^{N-1}}(q_k, {\bar{q}})\rightarrow 0\), as \(k\rightarrow +\infty \), then definitely \(q_k \in V_1\), \(\Pi _{T_{{\bar{q}}}\partial \Sigma _D}(q_k-{\bar{q}})\in U_1\), where \(\Pi _{T_{{\bar{q}}}\partial \Sigma _D}:{\mathbb {R}}^N\rightarrow T_{{\bar{q}}}\partial \Sigma _D\) is the orthogonal projection onto \(T_{{\bar{q}}}\partial \Sigma _D\) and \(\Pi _{T_{{\bar{q}}}\partial \Sigma _D}(q_k-{\bar{q}})\rightarrow 0\), as \(k\rightarrow +\infty \). Let \({\bar{x}}_k^\prime :=({\bar{x}}_{1,k}^\prime ,\ldots ,{\bar{x}}_{N-1,k}^\prime )\) be the coordinates of \(\Pi _{T_{{\bar{q}}}\partial \Sigma _D}(q_k-{\bar{q}})\) and set

$$\begin{aligned} U_{1,k}:= & {} U_1-{\bar{x}}_k^\prime ,\\ g_{1,k}(x^{\prime }):= & {} g_1(x^{\prime }+{\bar{x}}^\prime _k)-g_1({\bar{x}}_{k}^\prime ), \ x^\prime \in U_{1,k}. \end{aligned}$$

Then we readily check that \(V_1-q_k\) is a cartesian graph over \(T_{{\bar{q}}}\partial \Sigma _D\), associated to \(g_{1,k}:U_{1,k}\rightarrow {\mathbb {R}}\). Notice that, since \({\bar{x}}_k^\prime \rightarrow 0\), as \(k\rightarrow +\infty \), then there exists a ball \(B_{{\bar{R}}}(0)\) in \(T_{{\bar{q}}}\partial \Sigma _D\) such that \(B_{{\bar{R}}}(0)\subset U_{1,k}\) for all sufficiently large k. In particular, setting \(t_k:=|z_k|\), recalling that \(|z_k|\rightarrow +\infty \), as \(k\rightarrow +\infty \), then, \( t_kU_{1,k}\) invades \(T_{{\bar{q}}}\partial \Sigma _D\). As in (6.37) we consider the rescaled map \(h_{t_k}:t_kU_{1,k}\rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} h_{t_k}(x^{\prime })=t_k g_{1,k}\left( \frac{x^{\prime }}{t_k}\right) =t_k\left[ g_1\left( \frac{x^{\prime }}{t_k}+{\bar{x}}_k^\prime \right) -g_1({\bar{x}}_{k}^\prime )\right] , \ x^{\prime }\in t_k U_{1,k}, \end{aligned}$$

and arguing as in (6.38) we have the expansion

$$\begin{aligned} \frac{\partial h_{t_k}}{\partial x^{\prime }_i}(x^{\prime })=\frac{\partial g_1}{\partial x^\prime _i}\left( \frac{x^{\prime }}{t_k}+{\bar{x}}_k^\prime \right) =\frac{\partial g_1}{\partial x^\prime _i}\left( {\bar{x}}_k^\prime \right) + \frac{1}{t_k} \left[ \nabla _{x^\prime } \frac{\partial g_1}{\partial x^\prime _i}\left( \frac{\xi _k}{t_k}+{\bar{x}}_k^\prime \right) \right] \varvec{\cdot } x^{\prime },\nonumber \\ \end{aligned}$$
(6.45)

where \(\xi _k\) belongs to the segment joining \({\bar{x}}_k^\prime \) and \(\frac{x^{\prime }}{t_k}\). Let us fix a ball \(B_{R_1}(0)\) in \(T_{{\bar{q}}}\partial \Sigma _D\), with \(R_1\) to be chosen later and independently on k, and observe that \(B_{R_1}(0)\subset t_kU_{1,k}\) for all sufficiently large k. Since \({\bar{x}}_k^\prime \rightarrow 0\), as \(k\rightarrow +\infty \), and \(\nabla _{x^\prime } g_1(0)=0\) we get that the first term in the right-hand side of (6.45) goes to zero as \(k\rightarrow +\infty \), and arguing as in (6.39) for the second term, we conclude that

$$\begin{aligned} \lim _{k\rightarrow +\infty } \max _{x^{\prime } \in \overline{B_{R_1}(0)}}{|\nabla _{x^{\prime }} h_{t_k}(x^{\prime })|} = 0. \end{aligned}$$
(6.46)

Let \(F_{t_k}: \overline{C^+_{B_{R_1}(0)}} \rightarrow \overline{E^+\left( h_{t_k}\big |_{B_{R_1}(0)}\right) }\) be the diffeomorphism defined by

$$\begin{aligned} F_{t_k}(x^\prime ,x_N)=(x^\prime ,x_N+h_{t_k}(x^\prime )), \ (x^\prime ,x_N)\in \overline{C^+_{B_{R_1}(0)}}, \end{aligned}$$

where \(E^+\left( h_{t_k}\big |_{B_{R_1}(0)}\right) \) is the epigraph associated to \(h_{t_k}\big |_{B_{R_1}(0)}\) (see (6.41)) and where we recall that \(x=(x^\prime ,x_N)\) are the coordinates with respect to the orthogonal base \(\{{\bar{v}}_1,\ldots ,{\bar{v}}_{N-1},-\nu ({\bar{q}})\}\).

Now, let us consider the set \(B_{4R}(z_k)\cap \Omega _{n_k}\) appearing in (6.36). Recalling that \(z_k=t_k q_k\), since \(t_k\rightarrow +\infty \), as \(k\rightarrow +\infty \), and since R is independent of k, then, for all sufficiently large k it holds

$$\begin{aligned} \overline{(B_{4R}(z_k)\cap \Omega _{n_k})} \cap \partial \Sigma _D-z_k\subset t_k(V_1-q_k). \end{aligned}$$

We observe that for any k we have

$$\begin{aligned} {\mathcal {E}}(B_{4R}(z_k)\cap \Omega _{n_k}; \Sigma _D)={\mathcal {E}}(B_{4R}(z_k)\cap \Omega _{n_k}-z_k; \Sigma _D-z_k) \end{aligned}$$
(6.47)

and we set for brevity

$$\begin{aligned} {{\widetilde{\Omega }}}_{R,k}:=(B_{4R}(z_k)\cap \Omega _{n_k})-z_k. \end{aligned}$$
(6.48)

Notice that since \({{{\widetilde{\Omega }}}_{R,k}}\) is a uniformly bounded subset of \({\mathbb {R}}^N\) and \(t_k(V_1-q_k)\) is the cartesian graph of \(h_{t_k}:t_kU_{1,k}\rightarrow {\mathbb {R}}\) then we can choose \(R_1>0\) (independent of k) in such a way that

$$\begin{aligned} F_{t_k}^{-1}\left( \ \overline{{{\widetilde{\Omega }}}_{R,k}}\ \right) \subset B_{R_1}(0)\times [0,+\infty [, \end{aligned}$$
(6.49)

for all large k. Let us denote by \({\widetilde{u}}_{k}:= u_{{{\widetilde{\Omega }}}_{R,k}} \in H_0^1({{\widetilde{\Omega }}}_{R,k}; \Sigma _D-z_k)\) the energy function of \({{\widetilde{\Omega }}}_{R,k}\) and let \({\widetilde{U}}_k:={\widetilde{u}}_k\circ F_{t_k}\). Notice that by construction and thanks to (6.49), it follows that \({\tilde{U}}_k\) extends to a function \({\tilde{U}}_k \in H_0^1(F_{t_k}^{-1}({\tilde{\Omega }}_{R,k}); {\mathbb {R}}^N_+)\).

Thanks to (6.43) we have that \(\text {det}(\text {Jac}(F_{t_k}))\equiv 1\) and thus

$$\begin{aligned} \int _{{{\widetilde{\Omega }}}_{R,k}} {\widetilde{u}}_k \ \mathrm{d}y=\int _{F_{t_k}^{-1}({{\widetilde{\Omega }}}_{R,k})} {\widetilde{u}}_k\circ F_{t_k}\ \left| \text {det}(\text {Jac}(F_{t_k}))\right| \ \mathrm{d}x =\int _{F_{t_k}^{-1}({{\widetilde{\Omega }}}_{R,k})} {\widetilde{U}}_k \ \mathrm{d}x,\nonumber \\ \end{aligned}$$
(6.50)

and

$$\begin{aligned} |F_{t_k}^{-1}({{\widetilde{\Omega }}}_{R,k})|=|{{\widetilde{\Omega }}}_{R,k}|\leqq c. \end{aligned}$$
(6.51)

Moreover, setting

$$\begin{aligned} M_{1,k}:=\max _{(x^\prime ,x_N)\in \overline{C^+_{B_{R_1}(0)}}} \max _{|\eta |\leqq 1} |[\text {Jac}(F_{t_k})(x^\prime ,x_N)] \eta |, \end{aligned}$$

which is well defined (see (6.43)) and taking into account that

$$\begin{aligned} \nabla _x {\widetilde{U}}_k(x^\prime ,x_N)=(\text{ Jac }(F_{t_k})(x^\prime ,x_N))^T \nabla _{y} {\widetilde{u}}_k(F_{t_k}(x^\prime ,x_N)) \ \ \text{ for } \text{ a.e. }\ (x^\prime ,x_N) \in F_{t_k}^{-1}({{\widetilde{\Omega }}}_{R,k}), \end{aligned}$$

we obtain that

$$\begin{aligned} \begin{array}{lll} \displaystyle \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm{d}x&{}\leqq &{}\displaystyle M_{1,k}^2 \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})}|\nabla _{y} {\widetilde{u}}_k\circ F_{t_k}|^2 \ \mathrm{d}x\\ &{}=&{}\displaystyle M_{1,k}^2 \int _{{\tilde{\Omega }}_{R,k}}|\nabla _{y} {\widetilde{u}}_k|^2 \ \mathrm{d}y. \end{array} \end{aligned}$$
(6.52)

In view of (6.46) and recalling (6.43), (6.44) we deduce that \(M_{1,k}\rightarrow 1\) as \(k\rightarrow +\infty \). Therefore, recalling that \({\widetilde{u}}_k\) is the energy function of \({{\widetilde{\Omega }}}_{R,k}=B_{4R}(z_k)\cap \Omega _{n_k}-z_k\) and thanks to (6.52) we get that \( \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm{d}x\) is bounded by a uniform positive constant. Moreover, by definition and thanks to (6.50), (6.52) we have

$$\begin{aligned} \begin{array}{lll} \displaystyle {\mathcal {E}}({\tilde{\Omega }}_{R,k}; \Sigma _D-z_k)&{}{}=&{}{}\displaystyle \frac{1}{2}\int _{{\tilde{\Omega }}_{R,k}}|\nabla _{y} {\widetilde{u}}_k|^2 \ \mathrm {d}y -\int _{{{\widetilde{\Omega }}}_{R,k}} {\widetilde{u}}_k \ \mathrm {d}y\\ &{}{}\geqq &{}{}\displaystyle \frac{1}{2 } \frac{1}{M_{1,k}^2} \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm {d}x -\int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} {\tilde{U}}_k \ \mathrm {d}x\\ &{}{}=&{}{}\displaystyle J( {\tilde{U}}_k) + \displaystyle \frac{1}{2}\left( \frac{1}{M_{1,k}^2}-1\right) \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm {d}x. \end{array}\nonumber \\ \end{aligned}$$
(6.53)

Hence, as \({\tilde{U}}_k \in H_0^1(F_{t_k}^{-1}({\tilde{\Omega }}_{R,k}); {\mathbb {R}}^N_+)\), denoting by \(W_k:=u_{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} \in H_0^1(F_{t_k}^{-1} ({\tilde{\Omega }}_{R,k}); {\mathbb {R}}^N_+)\) the energy function of \(F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})\) then by reflection, symmetrization and taking into account (6.51) and (6.13) we infer that

$$\begin{aligned} \begin{array}{lll} {\mathcal {E}}({\tilde{\Omega }}_{R,k}; \Sigma _D-z_k)&{}{}\geqq &{}{}\displaystyle J(W_k) + \displaystyle \frac{1}{2}\left( \frac{1}{M_{1,k}^2}-1\right) \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm {d}x\\ &{}{}\geqq &{}{}\displaystyle \mathcal {O}_{c_k}(\Sigma _{{\mathbb {S}}^{N-1}_+}) + \displaystyle \frac{1}{2}\left( \frac{1}{M_{1,k}^2}-1\right) \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm {d}x\\ &{}{}\geqq &{}{}\displaystyle \mathcal {O}_{c}(\Sigma _{{\mathbb {S}}^{N-1}_+}) + \displaystyle \frac{1}{2}\left( \frac{1}{M_{1,k}^2}-1\right) \int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm {d}x. \end{array}\nonumber \\ \end{aligned}$$
(6.54)

Finally, since \(\int _{F_{t_k}^{-1}({\tilde{\Omega }}_{R,k})} |\nabla _{x} {\tilde{U}}_k|^2 \ \mathrm{d}x\) is uniformly bonded and \(M_{1,k}\rightarrow 1\), as \(k\rightarrow +\infty \), then from (6.47), (6.48) and (6.54), we get

$$\begin{aligned} {\mathcal {E}}(B_{4R}(z_k)\cap \Omega _{n_k}; \Sigma _D)\geqq \mathcal {O}_c(\Sigma _{{\mathbb {S}}^{N-1}_+}) +o(1), \end{aligned}$$

for all sufficiently large k, and this proves Claim (6.36).

In the next step, we prove the pre-compactness of the sequence \((u_{n_k})_k\) in \(L^2(\Sigma _D)\).

Step 5: The sequence \((u_{n_k})_k\) admits a subsequence which strongly converges in \(L^2(\Sigma _D)\).

We first show that

$$\begin{aligned} \lim _{R\rightarrow +\infty } \sup _{k \in {\mathbb {N}}} \int _{B_R^\complement (0)\cap \Sigma _D} u_{n_k}^2 \ \mathrm{d}x=0 \end{aligned}$$
(6.55)

Indeed, if (6.55) is not true there exist \(\varepsilon ^\prime >0\), a sequence \((R_m)_m\subset {\mathbb {R}}^+\) such that \(R_m\rightarrow +\infty \), as \(m\rightarrow +\infty \), and we find a subsequence \((n_{k_m})_m\) such that for all \(m\in {\mathbb {N}}\)

$$\begin{aligned} \int _{B_{R_m}^\complement (0)\cap \Sigma _D} u_{n_{k_m}}^2 \ \mathrm{d}x\geqq \frac{\varepsilon ^\prime }{2}. \end{aligned}$$
(6.56)

On the other hand, taking \(\varepsilon =\frac{\varepsilon ^\prime }{4}\) in (i) we find \(R^\prime >0\) depending only on \(\varepsilon ^\prime \) such that for all \(k \in {\mathbb {N}}\)

$$\begin{aligned} \int _{B_{R^\prime }(y_k)\cap \Sigma _D} u_{n_{k}}^2 \ \mathrm{d}x\geqq \uplambda -\frac{\varepsilon ^\prime }{4}. \end{aligned}$$
(6.57)

Now, in view of Step 4 we know that \((y_{n_k})_k\) is bounded, and thus there exists \(R^{\prime \prime }>0\) independent of k such that \(B_{R^\prime }(y_k)\subset B_{R^{\prime \prime }}(0)\) for all k. Hence, from (6.57) we get that

$$\begin{aligned} \int _{B_{R^{\prime \prime }(0)}\cap \Sigma _D} u_{n_{k}}^2 \ \mathrm{d}x \geqq \int _{B_{R^\prime }(y_k)\cap \Sigma _D} u_{n_{k}}^2 \ \mathrm{d}x\geqq \uplambda -\frac{\varepsilon ^\prime }{4}, \end{aligned}$$
(6.58)

for all sufficiently large k. Finally, by writing

$$\begin{aligned} \int _{\Sigma _D} u_{n_{k_m}}^2 \ \mathrm{d}x = \int _{B_{R_m}(0)\cap \Sigma _D} u_{n_{k_m}}^2 \ \mathrm{d}x+ \int _{B_{R_m}^\complement (0)\cap \Sigma _D} u_{n_{k_m}}^2 \ \mathrm{d}x, \end{aligned}$$

and recalling that \(R_m\rightarrow +\infty \), then, we have \(R_m>R^{\prime \prime }\), for all sufficiently large m, and thus from (6.56), (6.58) we deduce that

$$\begin{aligned} \int _{\Sigma _D} u_{n_{k_m}}^2 \ \mathrm{d}x \geqq \uplambda -\frac{\varepsilon ^\prime }{4} + \frac{\varepsilon ^\prime }{2} = \uplambda +\frac{\varepsilon ^\prime }{4}, \end{aligned}$$

for all sufficiently large m, but this contradicts (6.18) and (6.55) is thus proved.

In order to prove the relative compactness of the sequence \((u_{n_k})_k\) in \(L^2(\Sigma _D)\), it suffices to find, for any given \(\varepsilon >0\), a relative compact sequence \((v_k)_k\) in \(L^2(\Sigma _D)\), depending on \(\varepsilon \), with the property that

$$\begin{aligned} \Vert u_{n_k}-v_k\Vert _{L^2(\Sigma _D)}< \varepsilon \qquad \forall k \in {\mathbb {N}}. \end{aligned}$$
(6.59)

Indeed, the latter property readily implies that the set \(\{u_{n_k};\ k \in {\mathbb {N}}\}\) is totally bounded in \(L^2(\Sigma _D)\), and therefore it is relative compact since \(L^2(\Sigma _D)\) is a Banach space. So let \(\varepsilon >0\). By (6.55), there exists \(R>0\) with

$$\begin{aligned} \int _{B_R^\complement (0)\cap \Sigma _D} u_{n_k}^2 \ \mathrm{d}x< \varepsilon \qquad \forall k \in {\mathbb {N}}. \end{aligned}$$

Hence (6.59) holds with \(v_k: = \chi _{B_R(0)}u_{n_k}\) for \(k \in {\mathbb {N}}\), where \(\chi _{B_R(0)}\) denotes the characteristic function of the ball \(B_R(0)\). Moreover, by (6.16) and the compactness of the embedding \(H^1(B_R(0)\cap \Sigma _D) \hookrightarrow L^2(B_R(0)\cap \Sigma _D)\) the sequence of functions \(u_{n_k}\big |_{B_R(0)}\), \(k \in {\mathbb {N}}\) is relatively compact in \(L^2(B_R(0)\cap \Sigma _D)\), which obviously implies that the sequence \((v_k)_k\) is relatively compact in \(L^2(\Sigma _D)\), as required. We have thus established the relative compactness of the sequence \((u_{n_k})_k\) in \(L^2(\Sigma _D)\), as claimed.

Step 6: Existence of a minimizer for \({\mathcal {O}}_c(\Sigma _D)\).

In the previous steps we proved that the sequence of energy functions \((u_n)_n\), associated to a minimizing sequence \((\Omega _n)_n \subset \Sigma _D\) for \({\mathcal {O}}_c(\Sigma _D)\), is bounded in \(H^1(\Sigma _D)\) (see (6.16)) and possesses a subsequence which strongly converges in \(L^2(\Sigma _D)\). Hence, up to a subsequence (still indexed by n for convenience), we have \(u_n \rightharpoonup {\bar{u}}\) in \(H^1(\Sigma _D)\), for some \({\bar{u}} \in H^1(\Sigma _D)\), and \(u_n \rightarrow {\bar{u}}\) in \(L^2(\Sigma _D)\), as \(n\rightarrow +\infty \).

We set \(\Omega :=\{{\bar{u}}>0\} \subset \Sigma _D\). Since \({\bar{u}}\in H^1(\Sigma _D)\) then \(\Omega \) is a quasi-open subset of \(\Sigma _D\), in addition, arguing as in [8, Proof of Lemma 5.2], namely using that \(u_n \rightarrow {\bar{u}}\) in \(L^2(\Sigma _D)\), as \(n\rightarrow +\infty \), and applying Fatou’s Lemma, we infer that

$$\begin{aligned} |\Omega |=\int _{\Sigma _D} \chi _{\{{\bar{u}}>0\}} \ \mathrm{d}x \leqq \liminf _{n\rightarrow +\infty } \int _{\Sigma _D} \chi _{\{ u_n>0\}} \ \mathrm{d}x = \liminf _{n\rightarrow +\infty } |\Omega _n| \leqq c. \end{aligned}$$

We claim that \(\Omega \) is a minimizer for \({\mathcal {O}}_c(\Sigma _D)\) and that \({\bar{u}}\) is the torsion function of \(u_\Omega \). To prove this we first observe that as \(u_n \rightarrow {\bar{u}}\) in \(L^2(\Sigma _D)\) and since \(|\Omega _n|\leqq c\), \(|\Omega |\leqq c\) it follows that \(u_n \rightarrow {\bar{u}}\) in \(L^1(\Sigma _D)\). Indeed by construction we have \(u_n\in H_0^1(\Omega _n;\Sigma _D)\), \({\bar{u}} \in H_0^1(\Omega ;\Sigma _D)\), and by Hölder’s inequality we deduce that

$$\begin{aligned} \int _{\Sigma _D}|u_n-{\bar{u}}| \ \mathrm{d}x \leqq \left( \int _{\Omega _n \cup \Omega }|u_n-{\bar{u}}|^2 \ \mathrm{d}x\right) ^{\frac{1}{2}} |\Omega _n\cup \Omega |^{\frac{1}{2}}\leqq \Vert u_n-{\bar{u}}\Vert _{L^2(\Sigma _D)} \sqrt{2c}. \end{aligned}$$

Now, as \(u_n \rightharpoonup {\bar{u}}\) in \(H^1(\Sigma _D)\), we have

$$\begin{aligned} \Vert {\bar{u}}\Vert _{H^1(\Sigma _D)}^2 \leqq \liminf _{n\rightarrow +\infty } \Vert u_n\Vert _{H^1(\Sigma _D)}^2, \end{aligned}$$

and thus, since \(u_n \rightarrow {\bar{u}}\) in \(L^2(\Sigma _D)\), we readily get that

$$\begin{aligned} \int _{\Sigma _D} |\nabla {\bar{u}}|^2 \ \mathrm{d}x \leqq \liminf _{n\rightarrow +\infty } \int _{\Sigma _D} |\nabla u_n|^2 \ \mathrm{d}x \end{aligned}$$

Then, recalling the definition of the functional J (see (6.2)), exploiting that \(u_n \rightarrow {\bar{u}}\) in \(L^1(\Sigma _D)\) and since \(u_n\) are the energy functions associated to \(\Omega _n\), we obtain that

$$\begin{aligned} J({\bar{u}})\leqq \liminf _{n\rightarrow +\infty } \left( \frac{1}{2} \int _{\Sigma _D} |\nabla u_n|^2 \ \mathrm{d}x - \int _{\Sigma _D} u_n \ \mathrm{d}x \right) =\liminf _{n\rightarrow +\infty } {\mathcal {E}}(\Omega _n; \Sigma _D) ={\mathcal {O}}_c(\Sigma _D).\nonumber \\ \end{aligned}$$
(6.60)

Finally, considering the energy function \(u_{\Omega }\) associated to \(\Omega \), i.e. the minimizer of J in \(H_0^1(\Omega ;\Sigma _D)\), then, by the minimality of \(u_{\Omega }\), since \({\bar{u}} \in H_0^1(\Omega ;\Sigma _D)\) and thanks to (6.60) we have

$$\begin{aligned} {\mathcal {E}}(\Omega ; \Sigma _D)=J(u_{\Omega })\leqq J({\bar{u}}) \leqq {\mathcal {O}}_c(\Sigma _D). \end{aligned}$$
(6.61)

Therefore \({\mathcal {E}}(\Omega ; \Sigma _D)={\mathcal {O}}_c(\Sigma _D)\), and (6.61) implies that \(J(u_\Omega )=J({\bar{u}})\). Hence \(\Omega \) is a minimizer for \({\mathcal {O}}_c(\Sigma _D)\) and \({\bar{u}}=u_\Omega \) in \(H^1(\Sigma _D)\). \(\square \)

Corollary 6.9

If \(D\subset {\mathbb {S}}^{N-1}\) is a smooth domain such that

$$\begin{aligned} {\mathcal {H}}_{N-1}(D) < {\mathcal {H}}_{N-1}({\mathbb {S}}^{N-1}_+) \end{aligned}$$
(6.62)

then \(\mathcal {O}_c(\Sigma _D)\) is achieved, for any \(c>0\).

Proof

By (6.9)–(6.13) we readily check that (6.62) implies the condition (6.15), and by Theorem 6.8 we conclude. \(\square \)

We conclude this section with

Proposition 6.10

Let \(D\subset {\mathbb {S}}^{N-1}\) be a smooth domain and let \(c>0\). Then

$$\begin{aligned} \mathcal {O}_c\left( \Sigma _D\right) \leqq \mathcal {O}_c\left( \Sigma _{{\mathbb {S}}^{N-1}_+}\right) . \end{aligned}$$
(6.63)

Proof

Let us fix \(q\in \partial D\subset \partial \Sigma _D\setminus \{0\}\) and let \(\{v_1,\ldots ,v_{N-1}\}\) be an orthonormal basis of \(T_q\partial \Sigma _D\). We denote by \(x=(x^\prime ,x_N)\) the coordinates of points in \({\mathbb {R}}^N\) with respect to \(\{v_1,\ldots ,v_{N-1},-\nu (q)\}\), where \(-\nu (q)\) is the inner unit normal to \(\partial \Sigma _D\) at q. As seen in the proof of Claim (6.36) there exist an open neighborhood V of q in \(\partial \Sigma _D\setminus \{0\}\), an open neighborhood U of the origin in \(T_q\partial \Sigma _D\), and a smooth map \(g:{\overline{U}}\rightarrow {\mathbb {R}}\), \(g=g(x^\prime )\) such that \(V-q\) is the graph over \(T_q\partial \Sigma _D\) of \(g\big |_{U}\).

Let \(B^+_{R}(0)\subset {\mathbb {R}}^N_+\) be a N-dimensional half-ball such that \(|B^+_{R}(0)|=c\), i.e. \(B^+_{R}(0)\) is a half-ball of volume c contained in the upper half-space delimited by \(T_q\partial \Sigma _D\). Let \(u_{B^+_{R}(0)}\in H_0^1(B^+_{R}(0); {\mathbb {R}}^N_+)\) be the energy function of \(B^+_{R}(0)\). Then, by definition and recalling Remark 6.7, we have

$$\begin{aligned} \mathcal {O}_c(\Sigma _{{\mathbb {S}}^{N-1}_+})={\mathcal {E}}(B^+_{R}(0); {\mathbb {R}}^N_+)=J(u_{B^+_{R}(0)}). \end{aligned}$$
(6.64)

Let \(B_{R_1}(0)\) be a ball in \(T_q\partial \Sigma _D\), with \(R_1>R\). Clearly

$$\begin{aligned} \overline{B^+_R(0)} \subset B_{R_1}(0)\times [0,+\infty [. \end{aligned}$$
(6.65)

Let \((t_k)_k\subset {\mathbb {R}}^+\) be a sequence such that \(t_k\rightarrow +\infty \), as \(k\rightarrow +\infty \), then, setting \(z_k:=t_k q\) we obtain a diverging sequence of points on \(\partial \Sigma _D\setminus \{0\}\). We consider the rescaled map \(g_{t_k}:t_k U\rightarrow {\mathbb {R}}\) defined by (6.37) and the associated diffeomorphism \(F_{t_k}: \overline{C^+_{B_{R_1}(0)}} \rightarrow \overline{E^+\left( g_{t_k}\big |_{B_{R_1}(0)}\right) }\) given by (6.42), where \(E^+\left( g_{t_k}\big |_{B_{R_1}(0)}\right) \), \(C^+_{B_{R_1}(0)}\) are defined by (6.41). The inverse diffeomorphism \(F_{t_k}^{-1}\) is given by

$$\begin{aligned} F_{t_k}^{-1}(x^\prime ,x_N)=(x^\prime , x_N-g_{t_k}(x^\prime )), \ (x^\prime ,x_N)\in \overline{E^+\left( g_{t_k}\big |_{B_{R_1}(0)}\right) }, \end{aligned}$$
(6.66)

and as done in (6.43), (6.44) we readily check that

$$\begin{aligned} \mathrm {Det}(Jac(F_{t_k}^{-1}))\equiv 1,\ \lim _{k\rightarrow +\infty }\Vert \mathrm {Jac}(F^{-1}_{t_k})-{\mathbb {I}}_{N} \Vert _{C^0\left( \overline{E^+\left( g_{t_k}\big |_{B_{R_1}(0)}\right) }\right) }= 0.\nonumber \\ \end{aligned}$$
(6.67)

Moreover, setting \(U_k:=u_{B^+_R(0)}\circ F_{t_k}^{-1}\) we notice that since \(u_{B^+_R(0)}\in H_0^1(B^+_{R}(0); {\mathbb {R}}^N_+)\) (actually \(u_{B^+_R(0)}= 0\) in \({\mathbb {R}}^N_+\setminus B^+_{R}(0)\), see (6.9) with \(D={\mathbb {S}}^{N-1}_+\)), then, by construction, taking into account (6.65), it follows that \(U_k\) extends to a function \(U_k\in H_0^1(F_{t_k}\left( B^+_{R}(0)\right) ; \Sigma _D-z_k)\). Arguing as in (6.50)–(6.52), taking into account (6.67), we infer that \(|F_{t_k}\left( B^+_{R}(0)\right) |=|B^+_{R}(0)|=c\),

$$\begin{aligned} \int _{F_{t_k}\left( B^+_{R}(0)\right) } U_k \ \mathrm{d}x=\int _{B^+_{R}(0)} u_{B^+_{R}(0)} \ \mathrm{d}y, \end{aligned}$$
(6.68)
$$\begin{aligned} \displaystyle \int _{F_{t_k}\left( B^+_{R}(0)\right) } |\nabla _{x} U_k|^2 \ \mathrm{d}x\leqq \displaystyle M_{2,k}^2 \int _{B^+_{R}(0)}|\nabla _{y} u_{B^+_R(0)}|^2 \ \mathrm{d}y, \end{aligned}$$
(6.69)

where

$$\begin{aligned} M_{2,k}:=\max _{(x^\prime ,x_N)\in \overline{E^+\left( g_{t_k}\big |_{B_{R_1}(0)}\right) }} \max _{|\eta |\leqq 1} |[\text {Jac}(F^{-1}_{t_k})(x^\prime ,x_N)] \eta |, \end{aligned}$$

and \(M_{2,k}\rightarrow 1\), as \(k\rightarrow +\infty \). Hence, combining (6.64), (6.68) and (6.69) we deduce

$$\begin{aligned} \begin{array}{lll} \mathcal {O}_c(\Sigma _{{\mathbb {S}}^{N-1}_+})&{}{}=&{}{}\displaystyle \frac{1}{2}\int _{B^+_{R}(0)}|\nabla _{y} u_{B^+_R(0)}|^2 \ \mathrm {d}y -\int _{B^+_{R}(0)} u_{B^+_R(0)} \ \mathrm {d}y\\ &{}{}\geqq &{}{}\displaystyle \frac{1}{2}\frac{1}{M_{2,k}^2} \int _{F_{t_k}\left( B^+_{R}(0)\right) } |\nabla _{x} U_k|^2 \ \mathrm {d}x -\int _{F_{t_k}\left( B^+_{R}(0)\right) } U_k \ \mathrm {d}x \\ &{}{}=&{}{}\displaystyle J(U_k) + \displaystyle \frac{1}{2}\left( \frac{1}{M_{2,k}^2}-1\right) \int _{F_{t_k}\left( B^+_{R}(0)\right) } |\nabla _{x} U_k|^2 \ \mathrm {d}x \\ &{}{}\geqq &{}{}\displaystyle {\mathcal {E}}\left( F_{t_k}\left( B^+_{R}(0)\right) \right) + o(1) \end{array} \end{aligned}$$
(6.70)

where in the last inequality we used that \(U_k\in H_0^1(F_{t_k}\left( B^+_{R}(0)\right) ; \Sigma _D-z_k)\), the definition of torsional energy of \(F_{t_k}\left( B^+_{R}(0)\right) \), \(M_{2,k}\rightarrow 1\), as \(k\rightarrow +\infty \) and that \(\int _{F_{t_k}\left( B^+_{R}(0)\right) } |\nabla _{x} U_k|^2 \ \mathrm{d}x\) is uniformly bounded. Summing up, from (6.70) and the definition of \(\mathcal {O}_c(\Sigma _D)\) we finally have

$$\begin{aligned}&\mathcal {O}_c(\Sigma _{{\mathbb {S}}^{N-1}_+}) \geqq {\mathcal {E}}(F_{t_k}\left( B^+_{R}(0)\right) ; \Sigma _D-z_k) + o(1)= {\mathcal {E}}(F_{t_k}\left( B^+_{R}(0)\right) \\&\quad +z_k; \Sigma _D)+o(1)\geqq \mathcal {O}_c(\Sigma _D) + o(1), \end{aligned}$$

and passing to the limit as \(k\rightarrow +\infty \) we obtain (6.63). \(\square \)

7 Properties of Minimizers and Proof of Theorem 1.2

In this section we show some qualitative properties of the minimizers of the torsional energy functional with fixed volume (we refer to Sect. 6 for the notations). In view of the scaling invariance of our problem (see Remark 6.5) it suffices to focus on the case \(\mathcal {O}_1(\Sigma _D)\). We begin by proving that any minimizer for \(\mathcal {O}_1(\Sigma _D)\) is bounded.

Proposition 7.1

If \(\Omega \) is a minimizer for \(\mathcal {O}_1(\Sigma _D)\) then \(\Omega \) is bounded.

Proof

We argue as in [3, Sect. 2.1.2] with slightly changes. Let \(\Omega \) be a minimizer for \(\mathcal {O}_1(\Sigma _D)\). In view of [7, Theorem 1], in order to prove that \(\Omega \) is bounded it is sufficient to show that \(\Omega \) is a local shape subsolution for the energy \({{\mathcal {T}}}_D\), which means that, there exist \(\delta >0\) and \(\Lambda >0\) such that for any quasi-open subset \({{\widetilde{\Omega }}}\subset \Omega \) with \(\Vert u_{\Omega }-u_{{{\widetilde{\Omega }}}}\Vert _{L^2{\Sigma _D}}< \delta \), it holds that

$$\begin{aligned} {\mathcal {E}}(\Omega ; \Sigma _D)+\Lambda |\Omega | \leqq {\mathcal {E}}({{\widetilde{\Omega }}}; \Sigma _D)+\Lambda |{{\widetilde{\Omega }}}|. \end{aligned}$$

Let us assume, by contradiction, that there exist a sequence \((\Lambda _n)_n\subset {\mathbb {R}}^+\) with \(\Lambda _n\rightarrow 0\), as \(n\rightarrow +\infty \), and an increasing sequence \(({{\widetilde{\Omega }}}_n)_n\subset \Omega \) of quasi-open subsets such that

$$\begin{aligned} {\mathcal {E}}(\Omega ; \Sigma _D)+\Lambda _n |\Omega | > {\mathcal {E}}({{\widetilde{\Omega }}}_n; \Sigma _D)+\Lambda _n |{{\widetilde{\Omega }}}_n|, \end{aligned}$$
(7.1)

and \(\Vert u_{\Omega }-u_{{{\widetilde{\Omega }}}_n}\Vert _{L^2{\Sigma _D}}\rightarrow 0\), as \(n\rightarrow +\infty \). Then, let us fix \(t_n>1\) such that

$$\begin{aligned} |t_n {{\widetilde{\Omega }}}_n|=t_n^N|{{\widetilde{\Omega }}}_n|=1= |\Omega |. \end{aligned}$$

Obviously, \(t_n\rightarrow 1^+\), as \(n\rightarrow +\infty \), and by the minimality of \(\Omega \) we have

$$\begin{aligned} t_n^{N+2}{\mathcal {E}}({{\widetilde{\Omega }}}_n; \Sigma _D)= {\mathcal {E}}(t_n{{\widetilde{\Omega }}}_n; \Sigma _D)\geqq {\mathcal {E}}(\Omega ; \Sigma _D). \end{aligned}$$

Thus, from (7.1) we obtain

$$\begin{aligned} \Lambda _n (|\Omega |-|{{\widetilde{\Omega }}}_n|) > {\mathcal {E}}({{\widetilde{\Omega }}}_n; \Sigma _D) - {\mathcal {E}}(\Omega ; \Sigma _D) \geqq \frac{(t_n^{N+2}-1)}{t_n^{N+2}} (-{\mathcal {E}}(\Omega ; \Sigma _D)), \end{aligned}$$

and dividing by \(t_n^N-1=\frac{|\Omega |}{|{{\widetilde{\Omega }}}_n|}-1\) we get

$$\begin{aligned} \frac{-{\mathcal {E}}(\Omega ; \Sigma _D)}{t_n^{N+2}} \frac{(t_n^{N+2}-1)}{t_n^{N}-1} < \Lambda _n \frac{(|\Omega |-|{{\widetilde{\Omega }}}_n|)}{t_n^N-1}= \Lambda _n |{{\widetilde{\Omega }}}_n|. \end{aligned}$$
(7.2)

This gives a contradiction because, as \(n\rightarrow +\infty \), the left-hand side of (7.2) converges to \(-\frac{N+2}{N}{\mathcal {E}}(\Omega ; \Sigma _D)\), while the right-hand side converges to zero. \(\square \)

Proposition 7.2

If \(\Omega \) is a minimizer for \(\mathcal {O}_1(\Sigma _D)\) then the torsion function \(u_\Omega \in H_0^1(\Omega ; \Sigma _D)\) is Lipschitz continuous in any Lipschitz domain \(\omega \subset \Sigma _D\) such that \({\overline{\omega }}\subset \Sigma _D\) and \(\Omega =\{u_\Omega >0\}\) is an open subset of \(\Sigma _D\).

Proof

The result follows essentially as in the case of Dirichlet boundary conditions, which was addressed in the work [5]. The extension to the case of mixed boundary conditions was done in [21, Theorem 2.14] for the problem of minimizing the first eigenvalue. In our situation the proof would be similar. \(\square \)

Next, we prove that any minimizer is connected.

Proposition 7.3

If \(\Omega \) is a minimizer for \(\mathcal {O}_1(\Sigma _D)\) then \(\Omega \) is a connected subset of \(\Sigma _D\).

Proof

As seen in the proof of Theorem 6.8, Step 2, we notice that, as \(|\Omega |=1\), then \(\Omega \) is also a minimizer for

$$\begin{aligned} {\mathcal {M}}(\Sigma _D):=\inf \left\{ \frac{{\mathcal {E}}(\Omega ; \Sigma _D)}{|\Omega |^{\frac{N+2}{N}}}; \ \ \Omega \ \hbox {quasi-open}, \ \Omega \subset \Sigma _D,\ |\Omega |>0\right\} . \end{aligned}$$

Assume by contradiction that \(\Omega \) is not connected. Then there exist two open subsets \(\Omega _1\), \(\Omega _2\) of \(\Sigma _D\), with \(\Omega _1, \Omega _2\ne \emptyset \), such that \(\Omega _1\cap \Omega _2=\emptyset \) and

$$\begin{aligned} \Omega = \Omega _1\cup \Omega _2. \end{aligned}$$

In addition, since \(\Omega _i\) is an open nonempty subset of \(\Sigma _D\) then \(|\Omega _i|>0\), for \(i=1,2\), and by construction we have

$$\begin{aligned}&\frac{{\mathcal {E}}(\Omega _i; \Sigma _D)}{|\Omega _i|^{\frac{N+2}{N}}} \geqq {\mathcal {M}}(\Sigma _D), \ \text{ for }\ i=1,2,\ \text{ and }\\&{\mathcal {E}}(\Omega ; \Sigma _D)={\mathcal {E}}(\Omega _1; \Sigma _D) + {\mathcal {E}}(\Omega _2; \Sigma _D), \ \ |\Omega |=|\Omega _1|+|\Omega _2|. \end{aligned}$$

Then, since \(\frac{N+2}{N}>1\), by the convexity of \(t\mapsto t^{\frac{N+2}{N}}\), taking into account that \(|\Omega _1|>0\), \(|\Omega _2|>0\) and \({\mathcal {M}}(\Sigma _D)<0\), we deduce that

$$\begin{aligned} {\mathcal {M}}(\Sigma _D) |\Omega |^{\frac{N+2}{N}}= & {} {\mathcal {M}}(\Sigma _D) \left( |\Omega _1|+|\Omega _2|\right) ^{\frac{N+2}{N}}\\< & {} {\mathcal {M}}(\Sigma _D) \left( |\Omega _1|^{\frac{N+2}{N}} + |\Omega _2|^{\frac{N+2}{N}}\right) \\\leqq & {} {\mathcal {E}}(\Omega _1; \Sigma _D)+ {\mathcal {E}}(\Omega _2; \Sigma _D)={\mathcal {E}}(\Omega ; \Sigma _D), \end{aligned}$$

which is a contradiction. \(\square \)

Concerning the regularity of the minimizing set, by the theory of free boundary problems we have

Proposition 7.4

Let \(\Omega \subset \Sigma _D\) be a minimizer for \(\mathcal {O}_1(\Sigma _D)\) and let \(\Gamma =\partial \Omega \cap \Sigma _D\) be its relative boundary. Then there exists a critical dimension \(d^*\) which can be either 5,6 or 7, such that

  1. (i)

    \(\Gamma \) is smooth if \(N<d^*\);

  2. (ii)

    \(\Gamma \) can have countable isolated singularities if \(N=d^*\);

  3. (iii)

    \(\Gamma \) can have a singular set of dimension \(N-d^*\), if \(N>d^*\).

Moreover on the regular part of \(\Gamma \) the normal derivative \(\frac{\partial u_\Omega }{\partial \nu }\) is constant, namely \(\frac{\partial u_\Omega }{\partial \nu }\equiv - \sqrt{\frac{2(N+2)}{N} |{\mathcal {O}}_1(\Sigma _D)|}\), where \(u_\Omega \) is the torsion function of \(\Omega \).

Proof

The points (i)–(iii) follows from the results of [11, 15] and [27]. Let us prove the last statement.

Let \(\Gamma _{reg}\) be the regular part of \(\Gamma \) (which is a relative open set of \(\Gamma \)), let \(x_0\in \Gamma _{reg}\), and let \(B_r(x_0)\) be a small ball such that \({B_{2r}(x_0)}\subset \Sigma _D\) and \(\Gamma \cap {B_{2r}(x_0)} \subset \Gamma _{reg}\). Moreover, let \(\psi \in C^\infty _c(B_r(x_0))\) and consider the vector field V given by \(V=\psi {\bar{\nu }}\), where \({\bar{\nu }}\) is a smooth extension of the normal versor \(\left. \nu \right| _{\Gamma \cap {B_{2r}(x_0)}}\) of \(\Gamma \cap {B_{2r}(x_0)}\) to a smooth vector field defined in \(\overline{B_r(x_0)}\). Hence, by construction, we have that \(V:{\mathbb {R}}^N\rightarrow {\mathbb {R}}^N\) is a smooth vector field with compact support in \(B_r(x_0)\), and in particular it holds that \(V(0)=0\) and \(V(x)=0\in T_x\partial \Sigma _D\) for all \(x\in \partial \Sigma _D\setminus \{0\}\). This means that the associated flow \(\xi :(-t_0,t_0)\times \overline{\Sigma _D}\rightarrow \overline{\Sigma _D}\), for some \(t_0>0\), preserves the boundary \(\partial \Sigma _D\) and we consider the induced deformation of \(\Omega \), \((\Omega _t)_{t\in (-t_0,t_0)}\), where \(\Omega _t:=\xi (t,\Omega )\). Actually, since \(\mathrm {supp}(\psi ) \subset B_r(x_0)\) we infer that \(\xi (t,x)=x\) for \(x \in B_{\frac{3}{2} r}^\complement (x_0)\), \(t\in (-t_0,t_0)\).

Let \(u_{\Omega _t}\in H^1_0(\Omega _t;\Sigma _D)\) be the torsion function of \(\Omega _t\), for \(t\in (-t_0,t_0)\) and let us set \(u_t:=u_{\Omega _t}\). Arguing as in the proof of [23, Proposition 4.3] we can prove that the map from \((-t_0,t_0)\) to \(H^1(\Sigma _D)\), \(t\mapsto u_t\) is differentiable. In particular the function \(f:(-t_0,t_0)\rightarrow L^1(\Sigma _D)\), given by \(f(t)=|\nabla u_t|^2\) is differentiable. We also notice that, since \(u_\Omega \) is a weak solution to (6.1), then by standard elliptic regularity theory \(u_\Omega \in W^{2,2}(\Omega \cap B_r(x_0))\). In particular it holds that \(V f(0)=\psi {\bar{\nu }}|\nabla u_\Omega |^2\in W^{1,1}(\Sigma _D,{\mathbb {R}}^N)\) and by easy modifications to the proof of [14, Theorem 5.2.2] we infer that the function \(t\mapsto {\mathcal {E}}(\Omega _t;\Sigma _D)= -\frac{1}{2}\int _{\Omega _t}|\nabla u_t|^2 \ \mathrm{d}x\) is differentiable at \(t=0\) and

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}t}\left. \left( {\mathcal {E}}(\Omega _t;\Sigma _D)\right) \right| _{t=0}= -\int _\Omega \nabla u_\Omega \varvec{\cdot } \nabla u^\prime \ \mathrm{d}x - \frac{1}{2}\int _\Omega \mathrm {div}(V |\nabla u_\Omega |^2) \ \mathrm{d}x, \end{aligned}$$

where \(u^\prime =\frac{\mathrm{d}}{\mathrm{d}t}\left. \left( u_t\right) \right| _{t=0}\) is a solution to (see the proof of [23, Proposition 4.3])

$$\begin{aligned} {\left\{ \begin{array}{ll} -\Delta u^\prime = 0 &{} \text {in } \Omega ,\\ u^\prime = -\frac{\partial u_{\Omega _\varphi }}{\partial \nu } \langle V,\nu \rangle &{} \text {on } \Gamma ,\\ \frac{\partial u^\prime }{\partial \nu } = 0 &{} \text {on } \Gamma _{1}\setminus \{0\}. \end{array}\right. } \end{aligned}$$
(7.3)

We point out that since the flow \(\xi \) leaves invariant \(B^\complement _{\frac{3}{2}r}(x_0)\) for all \(t\in (-t_0,t_0)\), we have \(u^\prime \equiv 0\) in \(\Omega \cap B^\complement _{\frac{3}{2}r}(x_0)\) and thus

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}t}\left. \left( {\mathcal {E}}(\Omega _t;\Sigma _D)\right) \right| _{t=0}= \underbrace{-\int _{\Omega \cap B_{\frac{3}{2}r}(x_0)} \nabla u_\Omega \varvec{\cdot } \nabla u^\prime \ \mathrm{d}x}_\mathbf{(I) } - \underbrace{\frac{1}{2}\int _\Omega \mathrm {div}(V |\nabla u_\Omega |^2) \ \mathrm{d}x}_\mathbf{(II) }.\nonumber \\ \end{aligned}$$
(7.4)

Let us analyse \(\mathbf (I) \). We first observe that as \(\Gamma \cap B_{2r}(x_0)\subset \Gamma _{reg}\) and \(u^\prime \) is a solution to (7.3), then by standard elliptic regularity theory, it follows that \(u^\prime \in W^{2,2}\left( \Omega \cap B_{\frac{3}{2}r}(x_0)\right) \) and it is smooth inside \(\Omega \). Hence, applying the Green’s formula, taking into account that \(\Delta u^\prime =0\) in \(\Omega \cap B_{\frac{3}{2}r}(x_0)\), \(\frac{\partial u^\prime }{\partial \nu }=0\) on \(\Omega \cap \partial B_{\frac{3}{2}r}(x_0)\) (because \(u^\prime \equiv 0\) in \(B_{\frac{3}{2} r}^\complement (x_0)\) and \(u^\prime \) is smooth inside \(\Omega \)) and \(u_\Omega =0\) on \(\Gamma \cap \overline{B_{\frac{3}{2}r}(x_0)}\) we get that \(\mathbf (I) =0\). For \(\mathbf (II) \), applying the divergence theorem, and recalling the definition of V, in the end, we obtain

$$\begin{aligned} \begin{array}{lll} \displaystyle \frac{\mathrm{d}}{\mathrm{d}t}\left. \left( {\mathcal {E}}(\Omega _t;\Sigma _D)\right) \right| _{t=0}= & {} \displaystyle - \frac{1}{2}\int _{\Gamma \cap \overline{B_r(x_0)}} |\nabla u_\Omega |^2 \psi \ \mathrm{d}\sigma . \end{array} \end{aligned}$$
(7.5)

On the other hand, for the volume, we have

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}t}\left. \left( |\Omega _t|\right) \right| _{t=0} =\int _\Gamma \langle V,\nu \rangle \ \mathrm{d}\sigma =\int _{\Gamma \cap \overline{B_r(x_0)}} \psi \ \mathrm{d}\sigma . \end{aligned}$$
(7.6)

Now, since \(\Omega \) is a minimizer for \(\mathcal {O}_1(\Sigma _D)\), then, recalling Remark 6.5 and as observed in the proof of Step 2 of Theorem 6.8 we get that \(\Omega \) is also a minimizer for \(\frac{{\mathcal {E}}(\Omega ;\Sigma _D)}{|\Omega |^{\frac{N+2}{N}}}\). Thus from (7.5), (7.6), since \(|\Omega |=1\) and \({\mathcal {E}}(\Omega ;\Sigma _D)= {\mathcal {O}}_1(\Sigma _D)<0\), we readily obtain that

$$\begin{aligned} \frac{\mathrm{d}}{\mathrm{d}t}\left. \left( \frac{{\mathcal {E}}(\Omega _t;\Sigma _D)}{|\Omega _t|^{\frac{N+2}{N}}} \right) \right| _{t=0}=- \frac{1}{2}\int _{\Gamma \cap \overline{B_r(x_0)}} \left( |\nabla u_\Omega |^2-\frac{2(N+2)}{N} |{\mathcal {O}}_1(\Sigma _D)|\right) \psi \ \mathrm{d}\sigma =0. \end{aligned}$$

Finally, from the arbitrariness of \(\psi \) and \(x_0\) we conclude that \(|\nabla u_\Omega |^2\equiv \frac{2(N+2)}{N} |{\mathcal {O}}_1(\Sigma _D)|\) on \(\Gamma _{reg}\), and as \(u_\Omega =0\) on \(\Gamma _{reg}\) then by Hopf’s lemma it follows that \(\frac{\partial u_\Omega }{\partial \nu }\equiv - \sqrt{\frac{2(N+2)}{N} |{\mathcal {O}}_1(\Sigma _D)|}\) on \(\Gamma _{reg}\). \(\square \)

We conclude this section with the following:

Proof of Theorem 1.2

It follows from Theorem 5.1, Theorem 6.8, Corollary 6.9 and Proposition 7.4. \(\square \)

8 The Isoperimetric Problem and Proof of Theorem 1.3

In this section we study the isoperimetric problem in the class of strictly star-shaped domains in cones, i.e. domains in \(\Sigma _D\) whose relative boundary is a radial graph.

Using the same notations of the previous sections, if \(\varphi \in C^2({{\overline{D}}}, {\mathbb {R}})\) and \(\Omega _\varphi \), \(\Gamma _\varphi \) are, respectively, the associated star-shaped domain (see (3.3)), and the associated radial graph (see Definition 2.1), the (relative) perimeter of \(\Omega _\varphi \) in \(\Sigma _D\) is given by

$$\begin{aligned} {\mathcal {P}}(\Omega _\varphi ; \Sigma _D)= {\mathcal {H}}_{N-1}(\Gamma _\varphi )= \int _D e^{(N-1)\varphi } \sqrt{1+|\nabla \varphi |^2} \ \mathrm{d}\sigma , \end{aligned}$$

where \(\mathrm{d}\sigma \) is the \((N-1)\)-dimensional area element of \({\mathbb {S}}^{N-1}\). Let us observe that \({\mathcal {P}}\) can be seen as a functional on the space \(C^2({{\overline{D}}}, {\mathbb {R}})\) and, contrarily to the torsional energy functional of Sect. 3, its expression does not involve the associated domain \(\Omega _\varphi \). Thus we set for brevity \({\mathcal {P}}(\varphi )={\mathcal {P}}(\Omega _\varphi ; \Sigma _D)\).

The derivative of \({\mathcal {P}}\) along a variation \(v\in C^2({{\overline{D}}}, {\mathbb {R}})\) is given by

$$\begin{aligned} {\mathcal {P}}^\prime (\varphi )[v]= & {} \int _D e^{(N-1)\varphi } \left( (N-1) \sqrt{1+|\nabla \varphi |^2} v + \frac{\nabla \varphi \varvec{\cdot } \nabla v}{ \sqrt{1+|\nabla \varphi |^2}}\right) \ \mathrm{d}\sigma . \end{aligned}$$

Let \({\mathcal {V}}\) be the volume functional (see (4.1)). We are concerned with critical points \(\varphi \) of \({\mathcal {P}}\) subject to the volume constraint \(\{{\mathcal {V}}=c\}\). Namely we consider the manifold M defined by (4.4) and the restriction \({\mathcal {I}}:=\left. {\mathcal {P}}\right| _M\). A critical point \(\varphi \in M\) for \({\mathcal {I}}\) satisfies

$$\begin{aligned} {\mathcal {P}}^\prime (\varphi )=\uplambda {\mathcal {V}}^\prime (\varphi ), \end{aligned}$$
(8.1)

with a Lagrangian multiplier \(\uplambda \in {\mathbb {R}}\). In the next two propositions we prove that the radial graph \(\Gamma _\varphi \) associated to a critical point \(\varphi \in M\) is a CMC hypersurface which intersects orthogonally \(\partial \Sigma _D\setminus \{0\}\).

This is well known if the variations of the domains are taken in the whole class of subsets of \(\Sigma _D\) of finite relative perimeter (see [25]) but not obvious in our case.

Proposition 8.1

If \(\varphi \in C^2({{\overline{D}}},{\mathbb {R}})\) is a volume-constrained critical point for \({\mathcal {P}}\), then the associated radial graph \(\Gamma _\varphi \) has constant mean curvature \(H\equiv \frac{\uplambda }{N-1}\).

Proof

Let \(v \in C^{1}_c({D})\) be a variation with compact support. By definition (see (8.1)) there exists \(\uplambda \in {\mathbb {R}}\) such that

$$\begin{aligned} \int _D e^{(N-1)\varphi } \left( (N-1) \sqrt{1+|\nabla \varphi |^2} v+ \frac{\nabla \varphi \varvec{\cdot } \nabla v}{ \sqrt{1+|\nabla \varphi |^2}}\right) \ \mathrm{d}\sigma = \uplambda \int _D e^{N\varphi }v \ \mathrm{d}\sigma .\nonumber \\ \end{aligned}$$
(8.2)

Let us observe that

$$\begin{aligned}&\int _D e^{(N-1)\varphi } \frac{\nabla \varphi \varvec{\cdot } \nabla v}{ \sqrt{1+|\nabla \varphi |^2}} \ \mathrm{d}\sigma = \int _D \frac{\nabla \varphi \varvec{\cdot } \nabla (v e^{(N-1)\varphi })}{ \sqrt{1+|\nabla \varphi |^2}} \ \mathrm{d}\sigma \\&\quad - \int _D (N-1) e^{(N-1)\varphi } \frac{|\nabla \varphi |^2}{ \sqrt{1+|\nabla \varphi |^2}} v \ \mathrm{d}\sigma . \end{aligned}$$

Hence we can rewrite (8.2) as

$$\begin{aligned} \int _D \frac{\nabla \varphi \varvec{\cdot } \nabla (v e^{(N-1)\varphi })}{ \sqrt{1+|\nabla \varphi |^2}} \ \mathrm{d}\sigma + \int _D e^{(N-1)\varphi } \frac{N-1}{\sqrt{1+|\nabla \varphi |^2}} v \ \mathrm{d}\sigma = \uplambda \int _D e^{N\varphi }v \ \mathrm{d}\sigma . \end{aligned}$$

Now, since \(\varphi \) is smooth and v has compact support in D, integrating by parts, we get

$$\begin{aligned} \int _D \frac{\nabla \varphi \varvec{\cdot } \nabla (v e^{N\varphi })}{ \sqrt{1+|\nabla \varphi |^2}} \ \mathrm{d}\sigma = - \int _D \mathrm {div}_{{\mathbb {S}}^{N-1}}\left( \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}\right) e^{N\varphi } v \ \mathrm{d}\sigma , \end{aligned}$$
(8.3)

and thus we deduce

$$\begin{aligned} \int _D \left( -\mathrm {div}_{{\mathbb {S}}^{N-1}}\left( \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}\right) e^{(N-1)\varphi } + \frac{N}{\sqrt{1+|\nabla \varphi |^2}} e^{(N-1)\varphi }\right) v \ \mathrm{d}\sigma = \int _D \uplambda e^{N\varphi } v \ \mathrm{d}\sigma . \end{aligned}$$

Therefore, as v is arbitrary, we obtain

$$\begin{aligned} -\mathrm {div}_{{\mathbb {S}}^{N-1}}\left( \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}\right) e^{(N-1)\varphi } + \frac{N-1}{\sqrt{1+|\nabla \varphi |^2}} e^{(N-1)\varphi } = \uplambda e^{N\varphi }\ \ \ \text{ in }\ D, \end{aligned}$$

i.e.

$$\begin{aligned} -\mathrm {div}_{{\mathbb {S}}^{N-1}}\left( \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}\right) + \frac{N-1}{\sqrt{1+|\nabla \varphi |^2}} = \uplambda e^{\varphi } \ \ \text{ in }\ D. \end{aligned}$$
(8.4)

Comparing (8.4) with (2.10) it follows that the mean curvature of \(\Gamma _\varphi \) is constant and it is equal to \(\frac{\uplambda }{N-1}\). \(\square \)

Proposition 8.2

If \(\varphi \) is as in the statement of Proposition 8.1 then \(\Gamma _\varphi \) intersects orthogonally \(\partial \Sigma _D\setminus \{0\}\).

Proof

Let \(\nu _{_{\partial \Sigma _D}}\) be the exterior unit normal to \(\partial \Sigma _D\setminus \{0\}\), and let \(\nu _{_{\Gamma _\varphi }}\) be the exterior unit normal to \(\Gamma _\varphi \). By (2.7) we have

$$\begin{aligned} \nu _{_{\Gamma _\varphi }}({\mathcal {Y}}(q))= \frac{q- \nabla \varphi }{(1+|\nabla \varphi |^2)^{1/2}}, \end{aligned}$$
(8.5)

where \({\mathcal {Y}}\) is the standard parametrization of \(\Gamma _\varphi \) defined by (2.5). Notice that, since by assumption \(\varphi \) is smooth up to the boundary, then (8.5) is well defined on \({{\overline{D}}}\). If \(p \in (\partial \Sigma _D\setminus \{0\}) \cap \overline{\Gamma _\varphi }\), then by definition the intersection is orthogonal at p if and only if \(\nu _{_{\partial \Sigma _D}}(p) \varvec{\cdot } \nu _{_{\Gamma _\varphi }}(p) = 0\). Therefore, writing \(p={\mathcal {Y}}(q)\) this is equivalent to

$$\begin{aligned} \nu _{_{\partial \Sigma _D}}({\mathcal {Y}}(q)) \varvec{\cdot } (q- \nabla \varphi (q)) = 0. \end{aligned}$$

Since \(p={\mathcal {Y}}(q) \in \partial \Sigma _D\setminus \{0\}\) and \(\partial \Sigma _D\setminus \{0\}\) is the boundary of a cone we have \(\nu _{_{\partial \Sigma _D}}({\mathcal {Y}}(q)) \varvec{\cdot } q = 0\) and thus the intersection between \(\partial \Sigma _D\setminus \{0\}\) and \(\Gamma _\varphi \) is orthogonal if and only if

$$\begin{aligned} \nu _{_{\partial \Sigma _D}}({\mathcal {Y}}(q)) \varvec{\cdot } \nabla \varphi (q)= 0 \ \ \ \forall q \in \partial D. \end{aligned}$$
(8.6)

Exploiting again that \(\partial \Sigma _D\) is a cone, we have \(\nu _{_{\partial \Sigma _D}}(p)=\nu _{_{\partial \Sigma _D}}(t p)\) for any \(p\in \partial \Sigma _D\setminus \{0\}\), \(t>0\). Hence, since \({\mathcal {Y}}(q) \in \partial \Sigma _D\setminus \{0\}\), we have \(\nu _{_{\partial \Sigma _D}}({\mathcal {Y}}(q))=\nu _{_{\partial \Sigma _D}}(q)=\nu _{_{\partial D}}(q)\) for any \(q \in \partial D\), where \(\nu _{_{\partial D}}\) is the exterior unit co-normal to \(\partial D\), and thus (8.6) is equivalent to

$$\begin{aligned} \frac{\partial \varphi }{\partial \nu _{_{\partial D}}} = 0 \ \ \hbox {on} \partial D. \end{aligned}$$
(8.7)

To prove this, we argue as in the proof of Proposition 8.1. Taking a variation \(v \in C^1({{\overline{D}}}, {\mathbb {R}})\) and integrating by parts we have

$$\begin{aligned} \int _D \frac{\nabla \varphi \mathbf {\cdot } \nabla (v e^{(N-1)\varphi })}{ \sqrt{1+|\nabla \varphi |^2}} \ \mathrm {d}\sigma= & {} \int _{\partial D} e^{(N-1)\varphi } v \ \left\langle \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}, \nu _{_{\partial D}} \right\rangle \ \mathrm {d}{\hat{\sigma }} \\&- \int _D \mathrm {div}_{{\mathbb {S}}^{N-1}}\left( \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}\right) e^{(N-1)\varphi } v \ \mathrm {d}\sigma . \end{aligned}$$

Using this and arguing as in the proof Proposition 8.1, since \(\varphi \) satisfies the equation (8.4) we obtain

$$\begin{aligned} \int _{\partial D} e^{(N-1)\varphi } v \ \left\langle \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}, \nu _{_{\partial D}} \right\rangle \ \mathrm{d}{\hat{\sigma }}=0. \end{aligned}$$

Since \(v \in C^1({{\overline{D}}}, {\mathbb {R}})\) is arbitrary we can choose v such that \(v= \ \left\langle \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}, \nu _{_{\partial D}} \right\rangle \) on \(\partial D\) and thus

$$\begin{aligned} \int _{\partial D} e^{(N-1)\varphi } \left| \left\langle \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}, \nu _{_{\partial D}} \right\rangle \right| ^2 \ \mathrm{d}{\hat{\sigma }}=0, \end{aligned}$$

which gives \(\left\langle \frac{\nabla \varphi }{ \sqrt{1+|\nabla \varphi |^2}}, \nu _{_{\partial D}} \right\rangle \equiv 0\) on \(\partial D\), and thus (8.7) is proved. \(\square \)

Analogously to Lemma 4.3, if \(\varphi \in M\) is a critical point for \({\mathcal {I}}\) then

$$\begin{aligned} {\mathcal {I}}^{\prime \prime }(\varphi )={\mathcal {P}}^{\prime \prime }(\varphi )-\uplambda {\mathcal {V}}^{\prime \prime }(\varphi ). \end{aligned}$$

Choosing \(c=|\Omega _0|=|\Sigma _D\cap B_1(0)|\) in (4.4) we observe that the function \(\varphi \equiv 0\) belongs to M and it is a critical point for \({\mathcal {I}}\). In particular (8.1) yields \(\uplambda =N-1\). Moreover, for any \(v,w \in T_0M\), since

$$\begin{aligned} {\mathcal {P}}^{\prime \prime }(0)[v,w]= & {} \int _D \left( (N-1)^2 vw + {\nabla v \varvec{\cdot } \nabla w}\right) \ \mathrm{d}\sigma , \end{aligned}$$

and recalling (4.3), it follows that

$$\begin{aligned} {\mathcal {I}}^{\prime \prime }(0)[v,w]={\mathcal {P}}^{\prime \prime }(0)[v,w]-(N-1) {\mathcal {V}}^{\prime \prime }[v,w]= \int _D \left( \nabla v \varvec{\cdot } \nabla w - (N-1) vw\right) \ \mathrm{d}\sigma .\nonumber \\ \end{aligned}$$
(8.8)

From (8.8) we easily have the analogue of Theorem 5.1 for the perimeter functional \({\mathcal {I}}\).

Theorem 8.3

Let \(\uplambda _1(D)\) be the first nontrivial eigenvalue of \(-\Delta _{{\mathbb {S}}^{N-1}}\) on the domain D with zero Neumann condition on \(\partial D\). Then

  1. (i)

    if \(\uplambda _1(D) < N-1\) then \(\varphi \equiv 0\) is not a local minimizer for \({\mathcal {I}}\);

  2. (ii)

    if \(\uplambda _1(D) > N-1\) then \(\varphi \equiv 0\) is a local minimizer for \({\mathcal {I}}\).

Proof

Since \(T_0M\) is made by functions with zero mean value (see (4.5)), considering the \(L^2\)-normalized eigenfunction \(w_1\) corresponding to the eigenvalue \(\uplambda _1(D)\), from (8.8) we get \({\mathcal {I}}^{\prime \prime }(0)[w_1,w_1]<0\) whenever \(\uplambda _1(D)<N-1\). This proves (i).

Viceversa, if \(\uplambda _1(D)>N-1\), from (8.8) and the variational characterization of \(\uplambda _1(D)\) we get that \({\mathcal {I}}^{\prime \prime }[v,v]>0\) for all \(v\in T_0M\) with \(v\ne 0\), and hence (ii) holds. \(\square \)

To find examples of domains \(D\subset {\mathbb {S}}^{N-1}\) satisfying \(\uplambda _1(D)<N-1\) we can use the function \(u_e\in C^\infty ({\mathbb {S}}^{N-1})\) introduced in (5.8) and Proposition 5.2. Hence for the nonconvex domains constructed in the “Appendix”, the spherical sectors are not the minimizers of \({\mathcal {I}}\).

Concerning the existence of a minimizer for the relative perimeter \({\mathcal {P}}(E;\Sigma _D)\) in the whole class of finite perimeter subsets of \(\Sigma _D\), with a fixed volume, we summarise in the following the results stated in [25].

Theorem 8.4

Let \(D\subset {\mathbb {S}}^{N-1}\) be a domain such that \({\mathcal {H}}_{N-1}(D)\leqq {\mathcal {H}}_{N-1}({\mathbb {S}}^{N-1}_+)\). Then, there exists a set of finite perimeter \(E^*\) inside \(\Sigma _D\) which minimizes the relative perimeter under a volume constraint, for any value of the volume. Moreover any minimizer of the relative perimeter, with fixed volume, is a bounded set.

Proof

It follows from Proposition 3.5 and Proposition 3.7 in [25]. \(\square \)

We conclude this section with the following:

Proof of Theorem 1.3

The existence of a set of finite perimeter \(E^*\) inside \(\Sigma _D\) which minimizes the relative perimeter under a volume constraint, and its boundedness, follows from Theorem 8.4. From Theorem 8.3 we infer that \(E^*\) cannot be a spherical sector, while the properties (i)-(iii) of \(\Gamma _{E^*}\) derive from classical results for isoperimetric problems (see e.g. [25, Sect. 2] and the references therein). \(\square \)