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

\stackMath

Regularity of shape optimizers for some spectral fractional problems

Giorgio Tortone Giorgio Tortone
Dipartimento di Matematica, Università di Pisa
Largo B. Pontecorvo 5, 56127 Pisa - ITALY
[email protected]
Abstract.

This paper is dedicated to the spectral optimization problem

min{λ1s(Ω)++λms(Ω)+Λn(Ω):ΩD s-quasi-open}\mathrm{min}\left\{\lambda_{1}^{s}(\Omega)+\cdots+\lambda_{m}^{s}(\Omega)+\Lambda\mathcal{L}_{n}(\Omega)\colon\Omega\subset D\mbox{ s-quasi-open}\right\}

where Λ>0,Dn\Lambda>0,D\subset\mathbb{R}^{n} is a bounded open set and λis(Ω)\lambda_{i}^{s}(\Omega) is the ii-th eigenvalue of the fractional Laplacian on Ω\Omega with Dirichlet boundary condition on nΩ\mathbb{R}^{n}\setminus\Omega.
We first prove that the first mm eigenfunctions on an optimal set are locally Hölder continuous in the class C0,sC^{0,s} and, as a consequence, that the optimal sets are open sets. Then, via a blow-up analysis based on a Weiss type monotonicity formula, we prove that the topological boundary in DD of a minimizer Ω\Omega is composed of a relatively open regular part and a closed singular part of Hausdorff dimension at most nnn-n^{*}, for some n3n^{*}\geq 3. Finally we use a viscosity approach to prove C1,αC^{1,\alpha}-regularity of the regular part of the boundary.

Key words and phrases:
Shape optimization, Dirichlet eigenvalues, fractional Laplacian, viscosity solution, improvement of flatness.
2010 Mathematics Subject Classification:
49Q10, 35R11, 47A75, 49R05 35R35,
Work partially supported by the ERC project no. 853404 Variational approach to the regularity of the free boundaries - VAREG held by Bozhidar Velichkov.

1. Introduction

Functionals involving the eigenvalues of an elliptic operator and the Lebesgue measure were the object of an intense study in the last decades. The major interest of this researches was to better comprehend the interaction between the geometric features of the optimizers and their spectrum.

One of the most active line of research was the one related to eigenvalues of the Dirichlet Laplacian, that is

min{F(λ1(Ω),,λm(Ω))+Λn(Ω):ΩD quasi-open}\mathrm{min}\left\{F(\lambda_{1}(\Omega),\dots,\lambda_{m}(\Omega))+\Lambda\mathcal{L}_{n}(\Omega)\colon\Omega\subset D\mbox{ quasi-open}\right\}

with Λ>0,Dn\Lambda>0,D\subseteq\mathbb{R}^{n} and λi(Ω)\lambda_{i}(\Omega) the ii-th eigenvalue of the Dirichlet Laplacian on Ω\Omega. Unfortunately, despite of their simple formulation, these problems turn out to be quite challenging and an explicit construction of optimal sets is known only for the cases of the first (the optimizer is the ball) and the second eigenvalue (the optimizer is the union of two equal and disjoint balls) with D=nD=\mathbb{R}^{n}, thanks to the Faber-Krahn and Hong-Krahn-Szegö inequalities.
Indeed, for more general functionals, the regularity of the optimal sets and of the associated eigenfunctions turns out to be a rather difficult topic, usually due to the min-max formulation of the high order eigenvalues. Starting from the seminal contribution of Buttazzo-Dal Maso [9], and the additional existence results of [7, 30], several results were recently obtained both in the context of nondegenerate [6, 8, 31, 39, 40] and degenerate functionals [8, 27, 33] of eigenvalues.

Recently, many authors approached the same class of problems in the context of nonlocal operators, with greater attention on the fractional Laplacian (see for example [4, 5, 25, 28, 34, 38]). As in the local case, the validity of a Faber-Krahn inequality (see [4, Theorem 3.5]) immediately implies the minimality of the first eigenvalue of the fractional Laplacian on the ball, among domains of the same Lebesgue measure. Nevertheless, already in the case of the second eigenvalue the nonlocal attitude of the fractional Laplacian affects the minimization problem: indeed in [5] the authors proved a Hong-Krahn-Szegö type inequality which implies that the infimum of the second eigenvalue, among sets of the same Lebesgue measure, is reached by a sequence of two disjoint balls with same volume whose mutual distance tends to infinity. Focusing on this specific phenomena, in [34] the authors studied the behaviour of minimizing sequences for nonlocal shape functionals wondering how the mutual positions of connected components can impact the existence of minimizers. Exploiting a nonlocal version of a concentration-compactness type principle, they prove that either an optimal shape exists or there exists a minimizing sequence consisting of two components whose mutual distance tends to infinity.

In this paper we are interested in shape optimization problems depending on the first mm eigenfunctions of the fractional Laplacian. Precisely, for s(0,1)s\in(0,1), we consider the following nondegenerate problem

(1.1) min{λ1s(Ω)++λms(Ω)+Λn(Ω):ΩD s-quasi-open};\mathrm{min}\left\{\lambda_{1}^{s}(\Omega)+\cdots+\lambda_{m}^{s}(\Omega)+\Lambda\mathcal{L}_{n}(\Omega)\colon\Omega\subset D\text{ $s$-quasi-open}\right\};

where Λ>0,Dn\Lambda>0,D\subset\mathbb{R}^{n} is a bounded open set and λis(Ω)\lambda_{i}^{s}(\Omega) denotes the ii-th eigenvalue of the fractional Laplacian on Ω\Omega with Dirichlet boundary condition on nΩ\mathbb{R}^{n}\setminus\Omega, counted with the due multiplicity (see (2.2) for a detailed min-max definition).
Comparing the functional in (1.1) with the ones studied in the local case, if m=1m=1 our result extends to the fractional setting the minimization problem for the first eigenvalue first addressed in [6], while for m>1m>1 it can be seen as the fractional counterpart of the recent works of [26, 31].

Main results and organization of the paper.

Once we summarize few notions concerning the fractional Laplacian, its spectrum and the min-max formulation of high order eingevalues, we start by proving the existence of shape optimizers of (1.1), for every Λ>0,Dn\Lambda>0,D\subset\mathbb{R}^{n} open and bounded.
Afterwards, in order to translate the regularity issues of the minima Ω\Omega of (1.1) in terms of its first mm eigenfunctions, we prove the validity of an almost-minimality conditions for the vector V=(v1,,vm)V=(v^{1},\dots,v^{m}) of the first mm normalized eigenfunctions on Ω\Omega.
More precisely, in Proposition 2.1 we prove the existence of constant K,ε>0K,\varepsilon>0 such that

[V]Hs(n)2+Λn({|V|>0})(1+KV~VL1(n))[V~]Hs(n)2+Λn({|V~|>0})[V]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{\big{\lvert}{V}\big{\rvert}>0\})\leq\left(1+K\|{\tilde{V}-V}\|_{L^{1}(\mathbb{R}^{n})}\right)[\tilde{V}]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{|\tilde{V}|>0\})

for every V~H0s(D;m)\tilde{V}\in H^{s}_{0}(D;\mathbb{R}^{m}) such that V~L(n)VL(n)\|{\tilde{V}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\|{V}\|_{L^{\infty}(\mathbb{R}^{n})} and VV~L1(n)ε\|{V-\tilde{V}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon.
Nevertheless, the presence of the Gagliardo–Slobodeckiĭ seminorm makes unworkable the application of some basic tool such as the harmonic replacement and the blow-up analysis as well. Thus, in the same spirit of [12, 13, 14, 15, 18], we overcome this difficulty by considering the local realization of the fractional Laplacian in terms of the Caffarelli-Silvestre extension [10] (see Section 2 for more details).
Indeed, given the vector G=(g1,,gm)HΩ1,a(n+1;m)G=(g^{1},\dots,g^{m})\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}), with gig^{i} the extension in n+1\mathbb{R}^{n+1} of the ii-th normalized eigenfunctions on Ω\Omega, in Theorem 2.2 we introduce an almost-minimality condition in terms of the functional

𝒥(G)=n+1|y|a|G|2dX+Λ~n({|G|>0}n)with Λ~=2Λds,\mathcal{J}(G)=\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathbb{R}^{n})\quad\mbox{with }\tilde{\Lambda}=\frac{2\Lambda}{d_{s}},

that is, there exist K,ε>0K,\varepsilon>0 such that

𝒥(G)(1+KG~GL1(n))𝒥(G~)\mathcal{J}(G)\leq\left(1+K\|{\tilde{G}-G}\|_{L^{1}(\mathbb{R}^{n})}\right)\mathcal{J}(\tilde{G})

for every GG~L1(n)ε\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon (see Section 2 for the precise formulation). Moreover, exploiting the localization of the energies in 𝒥\mathcal{J}, we prove a localized almost-minimality condition, namely there exist σ>0,r0>0\sigma>0,r_{0}>0 such that

𝒥(G,Br(X0))(1+σrn)𝒥(G~,Br(X0))+σ2dsi=1mλis(Ω)rn,\mathcal{J}(G,B_{r}(X_{0}))\leq\left(1+\sigma r^{n}\right)\mathcal{J}(\tilde{G},B_{r}(X_{0}))+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)r^{n},

for every X0n,r(0,r0]X_{0}\in\mathbb{R}^{n},r\in(0,r_{0}] and GG~H01,a(Br(X0);m)G-\tilde{G}\in H^{1,a}_{0}(B_{r}(X_{0});\mathbb{R}^{m}) (see also (3.7) and (3.8) for two alternative almost-minimality conditions). By using this new formulations, we can prove the following regularity result for the first mm normalized eigenfunctions.

Theorem 1.1.

Let DnD\subset\mathbb{R}^{n} be an open bounded set and Λ>0\Lambda>0. Then, the shape optimization problem (1.1) admits a solution Ω\Omega. Moreover, the vector GH1,a(n+1;m)G\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) of the first mm normalized eigenfunctions is C0,sC^{0,s}-Hölder continuous in n+1\mathbb{R}^{n+1}. Thus, the shape optimizer Ω\Omega is an open set and

n(Ω)=n({vi0})=n({gi0}n)\mathcal{L}_{n}(\Omega)=\mathcal{L}_{n}(\{v^{i}\neq 0\})=\mathcal{L}_{n}(\{g^{i}\neq 0\}\cap\mathbb{R}^{n})

for every i=1,,mi=1,\dots,m.

The first part of the Theorem reflects the regularity result obtained in the local case in [8], where the authors proved Lipschitz continuity of eigenfunctions on optimal domains of (1.1) for s=1s=1.
The novelty of the second part of the Theorem lies in the validity of the unique continuation principle for eigenfunctions inside the optimal domains Ω\Omega even if disconnected. This is a purely nonlocal feature: indeed it is not possible to prove that the optimizer is connected, as in the local case [31, Corollary 4.3], but on the other hand the support of the first mm normalized eigenfunction coincides with Ω\Omega, up to a (n1)(n-1)-dimensional set.

Afterwards, with the aim to rely on a free boundary approach to the regularity of the shape optimizer, we denote through the paper

(1.2) {|G|>0}n=ΩandF(G)=ΩD,\{|G|>0\}\cap\mathbb{R}^{n}=\Omega\quad\mbox{and}\quad F(G)=\partial\Omega\cap D,

in such a way as to express the regularity issues of the thin domain Ωn\Omega\subset\mathbb{R}^{n} in terms of the “extended” vector GG. Thus, suitably modifying many ideas of the regularity theory with thin free boundary [12, 17, 18, 19], we finally prove the following result on the topological boundary of Ω\Omega.

Theorem 1.2.

Let Ω\Omega be a shape optimizer of (1.1) for some Λ>0,Dn\Lambda>0,D\subset\mathbb{R}^{n} open and bounded. Then, Ω\Omega has locally finite perimeter in DD and its topological boundary ΩD\partial\Omega\cap D in DD is a disjoint union of a regular part Reg(ΩD)\mathrm{Reg}(\partial\Omega\cap D) and one-phase singular set Sing(ΩD)\mathrm{Sing}(\partial\Omega\cap D):

  1. 1.

    Reg(ΩD)\mathrm{Reg}(\partial\Omega\cap D) is an open subset of Ω\partial\Omega and is locally the graph of a C1,αC^{1,\alpha} function.

  2. 2.

    Sing(ΩD)\mathrm{Sing}(\partial\Omega\cap D) consists only of points in which the Lebesgue density of Ω\Omega is strictly between 1/21/2 and 11. Moreover, there is n3n^{*}\geq 3 such that:

    • if n<nn<n^{*}, Sing(ΩD)\mathrm{Sing}(\partial\Omega\cap D) is empty;

    • if n=nn=n^{*}, Sing(ΩD)\mathrm{Sing}(\partial\Omega\cap D) contains at most a finite number of isolated points;

    • if n>nn>n^{*}, the (nn)(n-n^{*})-dimensional Hausdorff measure of Sing(ΩD)\mathrm{Sing}(\partial\Omega\cap D) is locally finite in DD.

Notice that the presence of a dimensional threshold n3n^{*}\geq 3 (see the precise definition (5.12)) is deeply related to the existence of ss-homogeneous global minimizer of the thin one-phase problem (see [14, 21]).
The analysis of the free boundary F(G)=ΩDF(G)=\partial\Omega\cap D is carried out by combining a classification of blow-up limits via a Weiss type monotonicity formula and a viscosity formulation of the problem near free boundary points.

More precisely, we prove that the vector of the first mm normalized eigenfunctions G=(g1,,gm)G=(g^{1},\dots,g^{m}) is a viscosity solutions (see Section 6 for the precise definition) of the problem

{LaG=0in n+1{y=0};yaG=λGin {|G|>0}n;ts|G|=ΛΓ(1+s)on F(G)with λ=(λ1s(Ω),,λms(Ω)),\begin{cases}-L_{a}G=0&\text{in $\mathbb{R}^{n+1}\setminus\{y=0\}$;}\\ -\partial^{a}_{y}G=\lambda G&\text{in $\{|G|>0\}\cap\mathbb{R}^{n}$;}\\ \frac{\partial}{\partial t^{s}}|G|=\frac{\sqrt{\Lambda}}{\Gamma(1+s)}&\text{on $F(G)$}\end{cases}\quad\text{with $\lambda=(\lambda_{1}^{s}(\Omega),\dots,\lambda_{m}^{s}(\Omega))$,}

where

ts|G|(X0):=limt0+|G|(x0+tν(x0),0)ts,X0=(x0,0)F(G),\frac{\partial}{\partial t^{s}}|G|(X_{0}):=\displaystyle\lim_{t\rightarrow 0^{+}}\frac{|G|(x_{0}+t\nu(x_{0}),0)}{t^{s}},\quad\textrm{$X_{0}=(x_{0},0)\in F(G)$},

with ν(x0)\nu(x_{0}) the unit normal to F(G)F(G) at X0X_{0} pointing toward {|G|>0}n\{|G|>0\}\cap\mathbb{R}^{n}.
Thus, the analysis of the regular part of the free boundary is addressed by slighlty modifying the viscosity methods developed in [17, 19]. More precisely, we prove an Harnack type inequality strictly relying on the behaviour of the first eigenfunction g1g^{1} and of the norm |G||G| near free boundary points. Lastly, the regularity of Reg(Ω)\mathrm{Reg}(\partial\Omega) is achieved by proving C1,αC^{1,\alpha} regularity result for flat free boundaries for a general class of linear problems (see Subsection 6.2 for more details).

Remark 1.1.

The main difference with the results of the local analogue [31], besides the issue of connection and unique continuation, lies in the density estimates on Ω\Omega. As pointed out in Remark 3.9, our proof only relies on the different local regularity of fractional eigenfunctions near their zero set depending on whether or not they change sign. This peculiarity has been already observed in other similar problem, and it resembles the same differences that appear between the classical two-phase problem of Alt-Caffarelli-Friedman [2] and its fractional counterpart [1] (see also [20, Theorem 1.1] where the authors described a similar phenomenon in the thin-counterpart of [32]).
Therefore, we think that this feature can be exploited in the analysis of branching points in shape optimization problems involving degenerate functional of fractional eigenvalues, that is

min{λk1s(Ω)++λkms(Ω)+Λn(Ω):ΩD s-quasi-open}with 1<k1km,\mathrm{min}\left\{\lambda_{k_{1}}^{s}(\Omega)+\cdots+\lambda_{k_{m}}^{s}(\Omega)+\Lambda\mathcal{L}_{n}(\Omega)\colon\Omega\subset D\text{ $s$-quasi-open}\right\}\quad\mbox{with }1<k_{1}\leq\cdots\leq k_{m},

where the differences with the local case could be more evident.

The paper is organized as follows: in Section 2 we prove the existence of shape optimizer for (1.1) and the almost-minimality of the eigenfunctions for a more general free boundary problem. At the end of the Section, exploiting the Caffarelli-Silvestre extension, we introduce a more useful formulation of these almost-minimality conditions. In Section 3 we study the local behavior of eigenfunctions answering the classical questions of optimal regularity, non-degeneracy and density estimates for local minimizers. Moreover, we prove the validity of a unique continuation result for the eigenfunctions of optimal domain. Then, in Section 4 we derive a Weiss-type formula which will allow to characterize the blow-up limits in Section 5. Finally, the blow-up analysis of Section 5 will lead to the definition of Reg(F(G))\mathrm{Reg}(F(G)) and Sing(F(G))\mathrm{Sing}(F(G)) and to some classical estimates of the Hausdorff dimension of the singular set. Finally, in Section 6 we introduce the notion of viscosity solution near F(G)F(G) and in Section 7.3 we use a viscosity approach to obtain C1,αC^{1,\alpha} smoothness of the regular part Reg(F(G))\mathrm{Reg}(F(G)). More precisely, we develop an Harnack type inequality for flat solutions, which will be the basic tool for an improvement of flatness result, from which the C1,αC^{1,\alpha} regularity of a flat free boundary follows by classic arguments.

Acknowledgements.

We are grateful to Dario Mazzoleni for useful discussions and suggestions.

Notations.

Let s(0,1)s\in(0,1), we say that Ω\Omega is a ss-quasi-open set if there exists a decreasing sequence {Ωk}k\{\Omega_{k}\}_{k\in\mathbb{N}} of open subsets of n\mathbb{R}^{n} such that ΩΩk\Omega\cup\Omega_{k} is open and caps(Ωk)0+\mathrm{cap}_{s}(\Omega_{k})\to 0^{+} as kk\to\infty. We refer to [3, 34] for more details on ss-capacity and ss-quasi-open set.
Let Ωn\Omega\subset\mathbb{R}^{n} be bounded and open, we introduce the space H0s(Ω)H^{s}_{0}(\Omega) as the completion of Cc(Ω)C^{\infty}_{c}(\Omega) with respect to the norm

uHs(n)=(nu2dx+[u]Hs(n)2)1/2\big{\|}{u}\big{\|}_{H^{s}(\mathbb{R}^{n})}=\left(\int_{\mathbb{R}^{n}}{u^{2}\mathrm{d}x}+[u]^{2}_{H^{s}(\mathbb{R}^{n})}\right)^{1/2}

where

(1.3) [u]Hs(n)2=C(n,s)2nn|u(x)u(z)|2|xz|n+2sdxdzwith C(n,s)=22ssΓ(n2+s)πn/2Γ(1s).[u]_{H^{s}(\mathbb{R}^{n})}^{2}=\frac{C(n,s)}{2}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}{\frac{\big{\lvert}{u(x)-u(z)}\big{\rvert}^{2}}{\big{\lvert}{x-z}\big{\rvert}^{n+2s}}\mathrm{d}x\mathrm{d}z}\quad\text{with }C(n,s)=\frac{2^{2s}s\Gamma(\frac{n}{2}+s)}{\pi^{n/2}\Gamma(1-s)}.

More generally, for a bounded ss-quasi-open set Ωn\Omega\subset\mathbb{R}^{n} we denote with H0s(Ω)H^{s}_{0}(\Omega) the Sobolev space defined as the set of functions in Hs(n)H^{s}(\mathbb{R}^{n}) which, up to a set of zero ss-capacity, vanish outside of Ω\Omega.
From now on, we denote by {ei}i=1,,n\{e_{i}\}_{i=1,\ldots,n} and {fi}i=1,,m\{f^{i}\}_{i=1,\ldots,m} canonical basis in n\mathbb{R}^{n} and m\mathbb{R}^{m} respectively. Unit directions in n\mathbb{R}^{n} and m\mathbb{R}^{m} will be typically denoted by ee and ff. The Euclidean norm in either space is denoted by |||\cdot| while the dot product is denoted by ,\langle\cdot,\cdot\rangle.
A point Xn+1X\in\mathbb{R}^{n+1} will be denoted by X=(x,y)n×X=(x,y)\in\mathbb{R}^{n}\times\mathbb{R} and we will use the notation x=(x,xn)x=(x^{\prime},x_{n}). Moreover, a ball in n+1\mathbb{R}^{n+1} with radius r>0r>0 centered at XX is denoted by Br(X)B_{r}(X) and for simplicity Br=Br(0)B_{r}=B_{r}(0). Also, we use r=Br{y=0}\mathcal{B}_{r}=B_{r}\cap\{y=0\} to denote the nn-dimensional ball in n×{y=0}\mathbb{R}^{n}\times\{y=0\}.
We will often consider the following sets: let gg be a continuous non-negative function in BrB_{r}, then

Br+(g)\displaystyle B^{+}_{r}(g) :=Br{(x,0):g(x,0)=0}n+1\displaystyle:=B_{r}\setminus\{(x,0)\colon g(x,0)=0\}\subset\mathbb{R}^{n+1}
r+(g)\displaystyle\mathcal{B}^{+}_{r}(g) :=Br+(g){y=0}n.\displaystyle:=B^{+}_{r}(g)\cap\{y=0\}\subset\mathbb{R}^{n}.

By abuse of notation, if G=(g1,,gm)G=(g^{1},\ldots,g^{m}) is a vector valued continuous function, we use Br+(G),r+(G)B_{r}^{+}(G),\mathcal{B}_{r}^{+}(G) in place of Br+(|G|),r+(|G|)B^{+}_{r}(|G|),\mathcal{B}^{+}_{r}(|G|) respectively. Also, we will denote with PP and LL respectively the half-hyperplane P:={Xn+1:xn0,y=0}P:=\{X\in\mathbb{R}^{n+1}\colon x_{n}\leq 0,y=0\} and L:={Xn+1:xn=0,y=0}L:=\{X\in\mathbb{R}^{n+1}\colon x_{n}=0,y=0\}.

In regard to the thin one-phase problem (6.1), we remark that if F(g)F(g) is C2C^{2} then any function gg which is harmonic in B1+(g)B^{+}_{1}(g) has an asymptotic expansion at a point x0F(g),x_{0}\in F(g),

g(x,y)=αU(xx0,ν(x0),y)+o(|xx0|s+|y|s).g(x,y)=\alpha U(\langle x-x_{0},\nu(x_{0})\rangle,y)+o(|x-x_{0}|^{s}+|y|^{s}).

Here U(t,z)U(t,z) is the prototype of regular one-plane solution of the thin one-phase problem (6.1), which is given by

(1.4) U(t,z)=(t2+z2+t2)s=rscos2s(θ2),U(t,z)=\left(\frac{\sqrt{t^{2}+z^{2}}+t}{2}\right)^{s}=r^{s}\cos^{2s}\left(\frac{\theta}{2}\right),

where in the second formulation we used the polar coordinates

t=rcosθ,z=rsinθ,r0,πθπ.t=r\cos\theta,\quad z=r\sin\theta,\quad r\geq 0,\quad-\pi\leq\theta\leq\pi.

2. Properties of the eigenfunctions on the optimal sets

In this Section we start with some general properties about the spectrum of the fractional Laplacian and we then prove existence of shape optimizers for the problem (1.1) among the class of ss-quasi-open sets. Afterwards, we introduce an almost-minimality condition satisfied by the eigenfunctions associated to shape optimizer, in terms of their Gagliardo–Slobodeckiĭ seminorm.
Since this condition is not convenient for our analysis, we conclude the Section with a reformulation of the almost-minimality by exploiting the local realization of the fractional Laplacian with the extension techinique.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded open set and s(0,1)s\in(0,1), consider the Dirichlet-type eigenvalue problem

(2.1) {(Δ)sv=λvin Ωv=0in nΩ.\begin{cases}(-\Delta)^{s}v=\lambda v&\mbox{in }\Omega\\ v=0&\mbox{in }\mathbb{R}^{n}\setminus\Omega.\end{cases}

where the fractional Laplacian is defined by

(Δ)sv(x)=C(n,s) P.V.nv(x)v(z)|xz|n+2sdzwith C(n,s)=22ssΓ(n2+s)πn/2Γ(1s).(-\Delta)^{s}v(x)=C(n,s)\mbox{ P.V.}\int_{\mathbb{R}^{n}}{\frac{v(x)-v(z)}{\big{\lvert}{x-z}\big{\rvert}^{n+2s}}\mathrm{d}z}\quad\text{with }C(n,s)=\frac{2^{2s}s\Gamma(\frac{n}{2}+s)}{\pi^{n/2}\Gamma(1-s)}.

In analogy with the local case, if λ\lambda\in\mathbb{R} is such that (2.1) admits a nontrivial solution vv, then we say that λ\lambda is an eigenvalue of Ω\Omega with vv an associated eigenfunction (see [4, 5, 29] and reference therein for more result and generalisation of the concept of fractional eigenvalues).
Moreover, the spectrum of the Dirichlet fractional Laplacian is a closed set of strictly positive real numbers defined as an increasing sequence

0<λ1s(Ω)<λ2s(Ω)λ3s(Ω)λms(Ω)0<\lambda_{1}^{s}(\Omega)<\lambda_{2}^{s}(\Omega)\leq\lambda_{3}^{s}(\Omega)\leq\cdots\leq\lambda_{m}^{s}(\Omega)\leq\cdots

where the strict inequality λ1s(Ω)<λ2s(Ω)\lambda_{1}^{s}(\Omega)<\lambda_{2}^{s}(\Omega) is a consequence of the nonlocal attitude of the fractional Laplacian (see [5, Theorem 2.8] and Subsection 3.3). Thus, from now on we will denote by viv^{i} the eigenfunction corresponding to the eigenvalue λis(Ω)\lambda_{i}^{s}(\Omega), normalized with respect to L2(Ω)L^{2}(\Omega)-norm. We recall that the family of eigenfunctions {vi}i\{v^{i}\}_{i\in\mathbb{N}} form a complete orthonormal system in L2(Ω)L^{2}(\Omega) and the Dirichlet eigenvalues λis(Ω)\lambda_{i}^{s}(\Omega) of the fractional Laplacian on Ω\Omega can be variationally characterized by the following min-max formulation

(2.2) λ1s(Ω)=min{[u]Hs(N)2uL2(N)2:uH0s(Ω){0}},λis(Ω)=minEiH0s(Ω)maxuEi{0}[u]Hs(N)2uL2(N)2,if i2,\displaystyle\begin{aligned} \lambda_{1}^{s}(\Omega)&=\min\left\{\frac{[u]^{2}_{H^{s}(\mathbb{R}^{N})}}{\|u\|^{2}_{L^{2}(\mathbb{R}^{N})}}:u\in H^{s}_{0}(\Omega)\setminus\{0\}\right\},\\ \lambda_{i}^{s}(\Omega)&=\min_{E_{i}\subset H^{s}_{0}(\Omega)}\max_{u\in E_{i}\setminus\{0\}}\frac{[u]^{2}_{H^{s}(\mathbb{R}^{N})}}{\|u\|^{2}_{L^{2}(\mathbb{R}^{N})}},\quad\text{if $i\geq 2$},\end{aligned}

where in the second case the infimum is taken over all ii-dimensional linear subspaces EiE_{i} of H0s(Ω)H^{s}_{0}(\Omega).
As pointed out in [5, Theorem 3.1](see also [4, Theorem 3.1]), the supremum of an eigenfunction viv^{i} on a set Ω\Omega can be estimated in terms of a power of the corresponding eigenvalue

(2.3) viL(n)[C~n,sλis(Ω)]n4s,where C~n,s=T2,s(nn2s)n2s2s\big{\|}{v^{i}}\big{\|}_{L^{\infty}(\mathbb{R}^{n})}\leq\left[\tilde{C}_{n,s}\lambda_{i}^{s}(\Omega)\right]^{\frac{n}{4s}},\quad\mbox{where }\tilde{C}_{n,s}=T_{2,s}\left(\frac{n}{n-2s}\right)^{\frac{n-2s}{2s}}

with T2,sT_{2,s} the sharp Sobolev constant.
As a consequence of the min-max formulation, we have the following variational formulation for the sum of the first mm Dirichlet eigenfunctions on a ss-quasi-open set Ω\Omega

(2.4) i=1mλis(Ω)=min{i=1m[ui]Hs(n)2:U=(u1,,um)H0s(Ω;m),Ωuiujdx=δij},\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)=\text{min}\left\{\sum_{i=1}^{m}[u^{i}]^{2}_{H^{s}(\mathbb{R}^{n})}\colon U=(u^{1},\dots,u^{m})\in H^{s}_{0}(\Omega;\mathbb{R}^{m}),\int_{\Omega}u^{i}u^{j}\mathrm{d}x=\delta_{ij}\right\},

where V=(v1,,vm)V=(v^{1},\dots,v^{m}) is the vector of the first mm-eigenfunctions normalized with respect to the L2L^{2}-norm. Finally, from this last formulation we can deduce the existence of a minimum for (1.1).

Proposition 2.1.

Let Λ>0\Lambda>0 and DnD\subset\mathbb{R}^{n} be a bounded open set. Then the shape optimization problem

min{i=1mλis(Ω~)+Λn(Ω~):Ω~D s-quasi-open}\mathrm{min}\Bigg{\{}\sum_{i=1}^{m}\lambda_{i}^{s}(\tilde{\Omega})+\Lambda\mathcal{L}_{n}(\tilde{\Omega})\colon\tilde{\Omega}\subset D\mbox{ s-quasi-open}\Bigg{\}}

admits a solution.

Proof.

Let (Ωk)k(\Omega_{k})_{k} be a minimizing sequence of ss-quasi-open sets to the problem (1.1).
Set Vk=(vk1,,vkm)V_{k}=(v_{k}^{1},\dots,v_{k}^{m}) as the vector of the first mm eigenfunctions on Ωk\Omega_{k}, then

[Vk]Hs(n)2=i=1m[vki]Hs(n)2=i=1mλis(Ωn)\left[V_{k}\right]^{2}_{H^{s}(\mathbb{R}^{n})}=\sum_{i=1}^{m}[v_{k}^{i}]^{2}_{H^{s}(\mathbb{R}^{n})}=\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega_{n})

which implies that VkV_{k} is uniformly bounded in Hs(n;m)H^{s}(\mathbb{R}^{n};\mathbb{R}^{m}). Thus, up to a subsequence, (Vk)k(V_{k})_{k} weakly converges in H0s(D;m)H^{s}_{0}(D;\mathbb{R}^{m}) and strongly in L2(D;m)L^{2}(D;\mathbb{R}^{m}) to some VH0s(D;m)V_{\infty}\in H^{s}_{0}(D;\mathbb{R}^{m}).
Since VV_{\infty} is an orthonormal vector, if we set Ω={|V|>0}\Omega_{\infty}=\{|V_{\infty}|>0\}, by the weak convergence in HsH^{s} and the lower semi-continuity of the Lebesgue measure, we get

i=1mλis(Ω)+Λ|Ω|\displaystyle\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega_{\infty})+\Lambda|\Omega_{\infty}| i=1m[vi]Hs(n)2+Λ|Ω|\displaystyle\leq\sum_{i=1}^{m}[v_{\infty}^{i}]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda|\Omega_{\infty}|
lim infk(i=1m[vki]Hs(n)2+Λ|Ωk|)\displaystyle\leq\liminf_{k\to\infty}\left(\sum_{i=1}^{m}[v_{k}^{i}]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda|\Omega_{k}|\right)
lim infk(i=1mλis(Ωk)+Λ|Ωk|),\displaystyle\leq\liminf_{k\to\infty}\left(\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega_{k})+\Lambda|\Omega_{k}|\right),

as we claimed. ∎

Hence, in view of the variational characterization (2.4) of the sum of the first mm Dirichlet eigenvalues, given Ω\Omega a shape optimizer of (1.1), the vector of the first mm normalized eigenfunctions V=(v1,,vm)V=(v^{1},\dots,v^{m}) on Ω\Omega is a solution to the problem

(2.5) min{[V]Hs(n)2+Λn({|V|>0}):VH0s(D;m),Dvivjdx=δij},\text{min}\left\{[V]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{\big{\lvert}{V}\big{\rvert}>0\})\colon V\in H^{s}_{0}(D;\mathbb{R}^{m}),\int_{D}v^{i}v^{j}\mathrm{d}x=\delta_{ij}\right\},

where Ω={|V|>0}\Omega=\{|V|>0\}. Moreover, by (2.3) we get

(2.6) |V|L(n)m[C~n,sλms(Ω)]n4s,where C~n,s=T2,s(nn2s)n2s2s\big{\|}{|V|}\big{\|}_{L^{\infty}(\mathbb{R}^{n})}\leq\sqrt{m}\left[\tilde{C}_{n,s}\lambda^{s}_{m}(\Omega)\right]^{\frac{n}{4s}},\quad\mbox{where }\tilde{C}_{n,s}=T_{2,s}\left(\frac{n}{n-2s}\right)^{\frac{n-2s}{2s}}

with T2,sT_{2,s} the sharp Sobolev constant. In the same spirit of [31, 39], by using the Gram-Schmidt procedure we remove the orthogonality constraint.

Lemma 2.2.

Let ΩD\Omega\subset D be a ss-quasi-open set and V=(v1,,vm)V=(v^{1},\dots,v^{m}) be the vector of the first mm normalized eigenfunctions on Ω\Omega. Let δ>0\delta>0 be fixed and set V~=(v~1,,v~m)H0s(D;m)\tilde{V}=(\tilde{v}^{1},\dots,\tilde{v}^{m})\in H^{s}_{0}(D;\mathbb{R}^{m}) be such that

εm:=i=1mv~iviL1(D)1andsupi=1,,m{viL(D)+v~iL(D)}δ.\varepsilon_{m}:=\sum_{i=1}^{m}\big{\|}{\tilde{v}^{i}-v^{i}}\big{\|}_{L^{1}(D)}\leq 1\quad\text{and}\quad\sup_{i=1,\dots,m}\left\{\big{\|}{v^{i}}\big{\|}_{L^{\infty}(D)}+\big{\|}{\tilde{v}^{i}}\big{\|}_{L^{\infty}(D)}\right\}\leq\delta.

Define W=(w1,,wm)H0s(D;m)W=(w^{1},\dots,w^{m})\in H^{s}_{0}(D;\mathbb{R}^{m}) as the vector obtained orthonormalizing V~\tilde{V} by the Gram-Schimdt procedure associated to the norm L2(n)L^{2}(\mathbb{R}^{n}), that is

wi=w~iw~iL2withw~i={v~1if i=1,v~ij=1i1(nv~iwjdx)wjif j>1.w^{i}=\frac{\tilde{w}^{i}}{\big{\|}{\tilde{w}^{i}}\big{\|}_{L^{2}}}\quad\mbox{with}\quad\tilde{w}^{i}=\begin{cases}\tilde{v}^{1}&\mbox{if }i=1,\\ \tilde{v}^{i}-\sum_{j=1}^{i-1}\left(\int_{\mathbb{R}^{n}}\tilde{v}^{i}w^{j}\mathrm{d}x\right)w^{j}&\mbox{if }j>1\end{cases}.

Then, there exist constants 1ε¯m>01\geq\overline{\varepsilon}_{m}>0 and C¯m>0\overline{C}_{m}>0, depending only on n,m,δn,m,\delta and Ω\Omega, such that the estimate

[W]Hs(n)2(1+C¯mεm)[V~]Hs(n)2[W]^{2}_{H^{s}(\mathbb{R}^{n})}\leq\left(1+\overline{C}_{m}\varepsilon_{m}\right)[\tilde{V}]^{2}_{H^{s}(\mathbb{R}^{n})}

holds for every V~\tilde{V} as above, with εε¯m\varepsilon\leq\overline{\varepsilon}_{m}.

Proof.

In terms of the components, we aim to prove

i=1m[wi]Hs(n)2(1+C¯mεk)i=1m[v~i]Hs(n)2.\sum_{i=1}^{m}\big{[}w^{i}\big{]}^{2}_{H^{s}(\mathbb{R}^{n})}\leq\left(1+\overline{C}_{m}\varepsilon_{k}\right)\sum_{i=1}^{m}\big{[}\tilde{v}^{i}\big{]}^{2}_{H^{s}(\mathbb{R}^{n})}.

Hence, the first step is to prove that there exists ε¯m,C¯m>0\overline{\varepsilon}_{m},\overline{C}_{m}>0 such that the following estimates holds whenever εmε¯m\varepsilon_{m}\leq\overline{\varepsilon}_{m}

i=1mwiviL1(r)Cmεm,supi=1,,mwiLCm\sum_{i=1}^{m}\big{\|}{w^{i}-v^{i}}\big{\|}_{L^{1}(\mathcal{B}_{r})}\leq C_{m}\varepsilon_{m},\quad\sup_{i=1,\dots,m}\big{\|}{w^{i}}\big{\|}_{L^{\infty}}\leq C_{m}

where ε¯m,Cm\overline{\varepsilon}_{m},C_{m} depend only on d,m,δd,m,\delta and Ω\Omega. Since it follows the same lines of the proof of [31, Lemma 2.5], we just sketch the proof. In particular we get

v~1w1L1\displaystyle\big{\|}{\tilde{v}^{1}-w^{1}}\big{\|}_{L^{1}} |v~1L121|v~1v1+v1L12v~1L1\displaystyle\leq\frac{|\big{\|}{\tilde{v}^{1}}\big{\|}_{L^{1}}^{2}-1|}{\big{\|}{\tilde{v}^{1}-v^{1}+v^{1}}\big{\|}_{L^{1}}^{2}}\big{\|}{\tilde{v}^{1}}\big{\|}_{L^{1}}
v~1v1L12+2v1Lv~1v1L112uLv~1v1L1(v1L1+v~1v1L1)\displaystyle\leq\frac{\big{\|}{\tilde{v}^{1}-v^{1}}\big{\|}_{L^{1}}^{2}+2\big{\|}{v^{1}}\big{\|}_{L^{\infty}}\big{\|}{\tilde{v}^{1}-v^{1}}\big{\|}_{L^{1}}}{1-2\big{\|}{u}\big{\|}_{L^{\infty}}\big{\|}{\tilde{v}^{1}-v^{1}}\big{\|}_{L^{1}}}(\big{\|}{v^{1}}\big{\|}_{L^{1}}+\big{\|}{\tilde{v}^{1}-v^{1}}\big{\|}_{L^{1}})
ε13δ12δε1(n(Ω)1/2+ε1)12δn(Ω)1/2ε1,\displaystyle\leq\varepsilon_{1}\frac{3\delta}{1-2\delta\varepsilon_{1}}(\mathcal{L}_{n}(\Omega)^{1/2}+\varepsilon_{1})\leq 12\delta\mathcal{L}_{n}(\Omega)^{1/2}\varepsilon_{1},

by choosing ε1min{n(Ω)1/2,(2δ)1}\varepsilon_{1}\leq\min\{\mathcal{L}_{n}(\Omega)^{1/2},(2\delta)^{-1}\}. Thus, it follows that v1w1L1(1+12δn(Ω)1/2)ε1\big{\|}{v^{1}-w^{1}}\big{\|}_{L^{1}}\leq(1+12\delta\mathcal{L}_{n}(\Omega)^{1/2})\varepsilon_{1}. Similarly, we get

(2.7) w1Lv~1L12v1Lv~1v1L1δ12δe12δ,\big{\|}{w^{1}}\big{\|}_{L^{\infty}}\leq\frac{\big{\|}{\tilde{v}^{1}}\big{\|}_{L^{\infty}}}{1-2\big{\|}{v^{1}}\big{\|}_{L^{\infty}}\big{\|}{\tilde{v}^{1}-v^{1}}\big{\|}_{L^{1}}}\leq\frac{\delta}{1-2\delta e_{1}}\leq 2\delta,

which proves our claim for m=1m=1. By induction, suppose now that the claim holds for 1,,m11,\dots,m-1 and let us first estimate the distance from vmv^{m} to the orthogonalized function

w~m:={v~1if m=1,v~mi=1m1(nv~mwidx)wiif k>1\tilde{w}^{m}:=\begin{cases}\tilde{v}^{1}&\mbox{if }m=1,\\ \tilde{v}^{m}-\sum_{i=1}^{m-1}\left(\int_{\mathbb{R}^{n}}\tilde{v}^{m}w^{i}\mathrm{d}x\right)w^{i}&\mbox{if }k>1\end{cases}

with respect to the L1L^{1}-topology. By the inductive step,

i=1m1|nv~mwidx|\displaystyle\sum_{i=1}^{m-1}\big{\lvert}{\int_{\mathbb{R}^{n}}\tilde{v}^{m}w^{i}\mathrm{d}x}\big{\rvert} i=1m1n|v~mvm||wivi|+|vi||v~mvm|+|vm||wivi|dx\displaystyle\leq\sum_{i=1}^{m-1}\int_{\mathbb{R}^{n}}|\tilde{v}^{m}-v^{m}||w^{i}-v^{i}|+|v^{i}||\tilde{v}^{m}-v^{m}|+|v^{m}||w^{i}-v^{i}|\mathrm{d}x
Cm1εm1δ+(m1)δεm+δCm1εm1\displaystyle\leq C_{m-1}\varepsilon_{m-1}\delta+(m-1)\delta\varepsilon_{m}+\delta C_{m-1}\varepsilon_{m-1}
(2Cm1δ+(m1)δ)εm,\displaystyle\leq(2C_{m-1}\delta+(m-1)\delta)\varepsilon_{m},

which implies

(2.8) vmw~mL1vmv~mL1+i=1m1|nv~mw1dx|(viL1+wiviL1)[1+(n(Ω)1/2+ε¯m1)(2Cm1δ+(m1)δ)]εm\displaystyle\begin{aligned} \big{\|}{v^{m}-\tilde{w}^{m}}\big{\|}_{L^{1}}&\leq\big{\|}{v^{m}-\tilde{v}^{m}}\big{\|}_{L^{1}}+\sum_{i=1}^{m-1}\big{\lvert}{\int_{\mathbb{R}^{n}}\tilde{v}^{m}w^{1}\mathrm{d}x}\big{\rvert}\left(\big{\|}{v^{i}}\big{\|}_{L^{1}}+\big{\|}{w^{i}-v^{i}}\big{\|}_{L^{1}}\right)\\ &\leq\left[1+(\mathcal{L}_{n}(\Omega)^{1/2}+\bar{\varepsilon}_{m-1})(2C_{m-1}\delta+(m-1)\delta)\right]\varepsilon_{m}\end{aligned}

and

(2.9) w~mLδ(1+Cm1(2Cm1+m1)).\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{\infty}}\leq\delta\left(1+C_{m-1}(2C_{m-1}+m-1)\right).

By setting for simplicity C~m\tilde{C}_{m} as the largest of the constants appearing on the right hand side of (2.8) and (2.9), we get

vmw~mL1C~mεmandw~mLC~m.\big{\|}{v^{m}-\tilde{w}^{m}}\big{\|}_{L^{1}}\leq\tilde{C}_{m}\varepsilon_{m}\quad\text{and}\quad\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{\infty}}\leq\tilde{C}_{m}.

Recalling that wm=w~mL21w~mw^{m}=\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{2}}^{-1}\tilde{w}^{m}, we have

(2.10) |w~mL21||2nvm(vmw~m)dx+n(vmw~m)2dx|2δC~mεm+(δ+C~m)C~mεm.\displaystyle\begin{aligned} \big{\lvert}{\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{2}}-1}\big{\rvert}&\leq\big{\lvert}{2\int_{\mathbb{R}^{n}}v^{m}(v^{m}-\tilde{w}^{m})\mathrm{d}x+\int_{\mathbb{R}^{n}}(v^{m}-\tilde{w}^{m})^{2}\mathrm{d}x}\big{\rvert}\\ &\leq 2\delta\tilde{C}_{m}\varepsilon_{m}+(\delta+\tilde{C}_{m})\tilde{C}_{m}\varepsilon_{m}.\end{aligned}

Hence, assume that εmε¯m:=12(2δC~m+(δ+C~m)C~m)1\varepsilon_{m}\leq\overline{\varepsilon}_{m}:=\frac{1}{2}(2\delta\tilde{C}_{m}+(\delta+\tilde{C}_{m})\tilde{C}_{m})^{-1}. Thus, 1/2w~mL23/21/2\leq\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{2}}\leq 3/2 and we have

wmL=w~mL21w~mL2C~m.\big{\|}{w^{m}}\big{\|}_{L^{\infty}}=\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{2}}^{-1}\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{\infty}}\leq 2\tilde{C}_{m}.

On the other hand, as in (2.8), we get

vmwmL1vmw~mL1+w~mwmL1(1+12C~mn(Ω)1/2)C~mεm,\big{\|}{v^{m}-w^{m}}\big{\|}_{L^{1}}\leq\big{\|}{v^{m}-\tilde{w}^{m}}\big{\|}_{L^{1}}+\big{\|}{\tilde{w}^{m}-w^{m}}\big{\|}_{L^{1}}\leq(1+12\tilde{C}_{m}\mathcal{L}_{n}(\Omega)^{1/2})\tilde{C}_{m}\varepsilon_{m},

for εmε¯m\varepsilon_{m}\leq\overline{\varepsilon}_{m}, for some ε¯m>0\overline{\varepsilon}_{m}>0 small enough which depends on C~m,δ\tilde{C}_{m},\delta and n(Ω)\mathcal{L}_{n}(\Omega). The inductive claim follows by defining

Cm:=2(1+12C~mn(Ω)1/2)C~m.C_{m}:=2(1+12\tilde{C}_{m}\mathcal{L}_{n}(\Omega)^{1/2})\tilde{C}_{m}.

Thus, we can finally prove the main inequality by induction. First, by (2.7) for m=1m=1

[w1]Hs(n)=[v~1]Hs(n)v~1L2[v~1]Hs(n)12v1Lv1v~1L1(1+4δε1)[v~1]Hs(n),[w^{1}]_{H^{s}(\mathbb{R}^{n})}=\frac{[\tilde{v}^{1}]_{H^{s}(\mathbb{R}^{n})}}{\big{\|}{\tilde{v}^{1}}\big{\|}_{L^{2}}}\leq\frac{[\tilde{v}^{1}]_{H^{s}(\mathbb{R}^{n})}}{1-2\big{\|}{v^{1}}\big{\|}_{L^{\infty}}\big{\|}{v^{1}-\tilde{v}^{1}}\big{\|}_{L^{1}}}\leq(1+4\delta\varepsilon_{1})[\tilde{v}^{1}]_{H^{s}(\mathbb{R}^{n})},

while for m>1m>1 we obtain

[wm]Hs(n)\displaystyle[w^{m}]_{H^{s}(\mathbb{R}^{n})} =[w~m]Hs(n)w~mL2\displaystyle=\frac{[\tilde{w}^{m}]_{H^{s}(\mathbb{R}^{n})}}{\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{2}}}
11|w~mL21|[v~mi=1m1(nv~mwidx)wi]Hs(n)\displaystyle\leq\frac{1}{1-|\big{\|}{\tilde{w}^{m}}\big{\|}_{L^{2}}-1|}\left[\tilde{v}^{m}-\sum_{i=1}^{m-1}\left(\int_{\mathbb{R}^{n}}\tilde{v}^{m}w^{i}\mathrm{d}x\right)w^{i}\right]_{H^{s}(\mathbb{R}^{n})}
(1+2(2δC~m+(δ+C~m)C~m)εm)([v~m]Hs(n)+i=1m1(nv~mwidx)[wi]Hs(n)).\displaystyle\leq(1+2(2\delta\tilde{C}_{m}+(\delta+\tilde{C}_{m})\tilde{C}_{m})\varepsilon_{m})\left([\tilde{v}^{m}]_{H^{s}(\mathbb{R}^{n})}+\sum_{i=1}^{m-1}\left(\int_{\mathbb{R}^{n}}\tilde{v}^{m}w^{i}\mathrm{d}x\right)[w^{i}]_{H^{s}(\mathbb{R}^{n})}\right).

where we used (2.10). Using again

i=1m1|nv~mwidx|((m1)δ+2Cm1δ)εm,\sum_{i=1}^{m-1}\left\lvert\int_{\mathbb{R}^{n}}\tilde{v}^{m}w^{i}\mathrm{d}x\right\lvert\leq((m-1)\delta+2C_{m-1}\delta)\varepsilon_{m},

and the induction hypothesis, we get the claimed result. ∎

We are now ready to prove the validity of an almost-minimality condition for shape optimizer of (1.1).

Proposition 2.1.

Suppose that ΩD\Omega\subset D is a solution to (1.1) for some Λ>0\Lambda>0. Then, the vector V=(v1,,vm)H0s(Ω;m)V=(v^{1},\dots,v^{m})\in H^{s}_{0}(\Omega;\mathbb{R}^{m}) of normalized eigenfunctions on Ω\Omega satisfies the following almost-minimality condition: for every δVL(Ω)\delta\geq\big{\|}{V}\big{\|}_{L^{\infty}(\Omega)} there exist K,ε>0K,\varepsilon>0 such that

(2.11) [V]Hs(n)2+Λn({|V|>0})(1+KV~VL1(n))[V~]Hs(n)2+Λn({|V~|>0})[V]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{\big{\lvert}{V}\big{\rvert}>0\})\leq\left(1+K\|{\tilde{V}-V}\|_{L^{1}(\mathbb{R}^{n})}\right)[\tilde{V}]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{|\tilde{V}|>0\})

for every V~H0s(D;m)\tilde{V}\in H^{s}_{0}(D;\mathbb{R}^{m}) such that V~L(n)δ\|{\tilde{V}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta and VV~L1(n)ε\|{V-\tilde{V}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon.

Proof.

Let V~H0s(D;m)\tilde{V}\in H^{s}_{0}(D;\mathbb{R}^{m}) be the vector-valued function satisfying the assumptions of Proposition 2.1 and let WH0s(D;m)W\in H^{s}_{0}(D;\mathbb{R}^{m}) be the function obtained by the orthonormalizing procedure of Lemma 2.2 applied on V~\tilde{V}. Since VV is a solution of the minimization problem (2.5) and WW satisfies

nwiwjdx=δij,\int_{\mathbb{R}^{n}}w^{i}w^{j}\mathrm{d}x=\delta_{ij},

we can use it as a test function in (2.5) obtaining

[V]Hs(n)2+Λn({|V|>0})\displaystyle[V]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{\big{\lvert}{V}\big{\rvert}>0\}) [W]Hs(n)2+Λn({|W|>0})\displaystyle\leq[W]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{\big{\lvert}{W}\big{\rvert}>0\})
(1+C¯kVV~L1(n))[V~]Hs(n)2+Λn({|V~|>0}),\displaystyle\leq(1+\overline{C}_{k}\|{V-\tilde{V}}\|_{L^{1}(\mathbb{R}^{n})})[\tilde{V}]^{2}_{H^{s}(\mathbb{R}^{n})}+\Lambda\mathcal{L}_{n}(\{|\tilde{V}|>0\}),

where the last inequality follows by Lemma 2.2 and the fact that by construction n({|W|>0})n({|V~|>0})\mathcal{L}_{n}(\{\big{\lvert}{W}\big{\rvert}>0\})\subset\mathcal{L}_{n}(\{|\tilde{V}|>0\}). ∎

2.1. The extended formulation in n+1\mathbb{R}^{n+1}

The main problem of the previous formulation is the lack of monotonicity result for ss-harmonic function, which makes unworkable the strategy of the harmonic replacement. Hence, we overcome this difficulty by considering the local realization of the fractional Laplacian in terms of the Caffarelli-Silvestre extension [10], that gives an equivalent formulation of the previous non-local variational problem as a local problem defined in one extra dimension. This is nowadays a common strategy and it has been deeply used in the context of fractional problem with free boundary (see [11, 12, 13, 14, 15, 18, 21] among others).
Thus, we already know that for Ωn\Omega\subset\mathbb{R}^{n} ss-quasi-open and vH0s(Ω)v\in H^{s}_{0}(\Omega) it holds

(2.12) [v]Hs(n)2=ds+n+1|y|a|g|2dX,with ds=22s1Γ(s)Γ(1s),[v]^{2}_{H^{s}(\mathbb{R}^{n})}=d_{s}\int_{\mathbb{R}^{n+1}_{+}}|y|^{a}|\nabla g|^{2}\mathrm{d}X,\quad\mbox{with }d_{s}=2^{2s-1}\frac{\Gamma(s)}{\Gamma(1-s)},

where X=(x,y)n+1,a=12s(1,1)X=(x,y)\in\mathbb{R}^{n+1},a=1-2s\in(-1,1) and gH1,a(n+1)g\in H^{1,a}(\mathbb{R}^{n+1}) is uniquely defined as the LaL_{a}-extension of vv in +n+1\mathbb{R}^{n+1}_{+} (see [23, 22] for more result of weighted degenerate Sobolev spaces H1,aH^{1,a} and weighted degenerate Lebesgue spaces L2,aL^{2,a}).
Notice that, in this new setting, we can write the eigenvalue problem (2.1) as follows

(2.13) {Lag=0in +n+1yag=λdsgon Ω×{0}g=0on (nΩ)×{0},with yag(x,0)=limy0+yayg(x,y)\begin{cases}L_{a}g=0&\mbox{in }\mathbb{R}^{n+1}_{+}\\ -\partial^{a}_{y}g=\frac{\lambda}{d_{s}}g&\mbox{on }\Omega\times\{0\}\\ g=0&\mbox{on }(\mathbb{R}^{n}\setminus\Omega)\times\{0\},\end{cases}\quad\mbox{with }\partial^{a}_{y}g(x,0)=\lim_{y\to 0^{+}}y^{a}\partial_{y}g(x,y)

and g(x,0)=v(x)g(x,0)=v(x) in n\mathbb{R}^{n}. Similarly, we can reformulate the shape minimization problem (1.1) as

(2.14) i=1mλis(Ω)=dsmin{+n+1|y|a|G|2dX:G=(g1,,gm)HΩ1,a(+n+1;m),Ωgigjdx=δij}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)=d_{s}\text{min}\left\{\int_{\mathbb{R}^{n+1}_{+}}|y|^{a}|\nabla G|^{2}\mathrm{d}X\colon G=(g^{1},\dots,g^{m})\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1}_{+};\mathbb{R}^{m}),\int_{\Omega}g^{i}g^{j}\mathrm{d}x=\delta_{ij}\right\}

where gig^{i} is the LaL_{a}-extension of the normalized eigenfunction viv^{i} (that is, it solves (2.13) with λ=λis(Ω),v=vi\lambda=\lambda_{i}^{s}(\Omega),v=v^{i}), and

HΩ1,a(+n+1;m)={GH1,a(+n+1;m):|G|(x,0)0 in nΩ}.H^{1,a}_{\Omega}(\mathbb{R}^{n+1}_{+};\mathbb{R}^{m})=\left\{G\in H^{1,a}(\mathbb{R}^{n+1}_{+};\mathbb{R}^{m})\colon|G|(x,0)\equiv 0\mbox{ in }\mathbb{R}^{n}\setminus\Omega\right\}.

For the sake of simplicity, we will call G=(g1,,gm)G=(g^{1},\dots,g^{m}) in (2.14) as the vector of normalized eigenfunctions on Ω\Omega. Moreover, we can apply an even reflection through the thin-space {y=0}\{y=0\} and assume that each component of GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) is symmetric with respect to the yy-variable.
Thus, we can rephrase Proposition 2.1 as follows:

Let ΩD\Omega\subset D be a solution to (1.1) for some Λ>0\Lambda>0. Then, given Λ~=2Λ/ds\tilde{\Lambda}=2\Lambda/d_{s} and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m})

the vector of normalized eigenfunctions on Ω\Omega, for every δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} there exist K,ε>0K,\varepsilon>0 such that

(2.15) n+1|y|a|G|2dX+Λ~n({|G|>0}n)\displaystyle\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathbb{R}^{n})\leq (1+KG~GL1(n))n+1|y|a|G~|2dX+\displaystyle\,\left(1+K\|{\tilde{G}-G}\|_{L^{1}(\mathbb{R}^{n})}\right)\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X+
+Λ~n({|G~|>0}n)\displaystyle\,+\tilde{\Lambda}\mathcal{L}_{n}(\{|\tilde{G}|>0\}\cap\mathbb{R}^{n})

for G~HD1,a(n+1;m)\tilde{G}\in H^{1,a}_{D}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(n)δ\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta and GG~L1(n)ε\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon.

In the following Proposition we stress out how this new formulation allows to localize the problem. More precisely, the normalized eigenvector GG satisfies the following localized version of the almost-minimality condition.

Proposition 2.3.

Let ΩD\Omega\subset D be a solution to (1.1), for some Λ>0\Lambda>0, and G=(g1,,gm)HΩ1,a(n+1;m)G=(g^{1},\dots,g^{m})\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions. Then, for every δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} there exist σ,r0>0\sigma,r_{0}>0 such that, for every r(0,r0]r\in(0,r_{0}] and X0n×{0}X_{0}\in\mathbb{R}^{n}\times\{0\}

(2.16) Br(X0)|y|a|G|2dX+Λ~n({|G|>0}r(X0))\displaystyle\int_{B_{r}(X_{0})}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathcal{B}_{r}(X_{0}))\leq (1+σrn)Br(X0)|y|a|G~|2dX+\displaystyle\,\left(1+\sigma r^{n}\right)\int_{B_{r}(X_{0})}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X+
+Λ~n({|G~|>0}r(X0))\displaystyle\,+\tilde{\Lambda}\mathcal{L}_{n}(\{|\tilde{G}|>0\}\cap\mathcal{B}_{r}(X_{0}))
+σ2dsi=1mλis(Ω)rn,\displaystyle\,+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)r^{n},

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(r(X0))δ\|{\tilde{G}}\|_{L^{\infty}(\mathcal{B}_{r}(X_{0}))}\leq\delta and GG~H01,a(Br(X0);m)G-\tilde{G}\in H^{1,a}_{0}(B_{r}(X_{0});\mathbb{R}^{m}).

Proof.

Let δ>0\delta>0 and K,ε>0K,\varepsilon>0 be the constants defined in condition (2.15). Consider now r(0,r0]r\in(0,r_{0}] with

r0=(ε2ωnδ)1/n.r_{0}=\left(\frac{\varepsilon}{2\omega_{n}\delta}\right)^{1/n}.

Let G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be such that G~L(r(X0))δ\|{\tilde{G}}\|_{L^{\infty}(\mathcal{B}_{r}(X_{0}))}\leq\delta and GG~H01,a(Br(X0);m)G-\tilde{G}\in H^{1,a}_{0}(B_{r}(X_{0});\mathbb{R}^{m}), where GG~G-\tilde{G} is extended by zero in +n+1¯\overline{\mathbb{R}^{n+1}_{+}}. By the definition of r0>0r_{0}>0, we have

(2.17) GG~L1(n)2ωnδrn2ωnδr0nε.\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq 2\omega_{n}\delta r^{n}\leq 2\omega_{n}\delta r_{0}^{n}\leq\varepsilon.

Hence, by (2.15), we get

Br(X0)|y|a(|G|2|G~|2)dX+Λ~n({|G|>0}r(X0))\displaystyle\int_{B_{r}(X_{0})}|y|^{a}(|\nabla G|^{2}-|\nabla\tilde{G}|^{2})\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathcal{B}_{r}(X_{0}))\leq σrnn+1|y|a|G~|2dX+\displaystyle\,\sigma r^{n}\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X+
+Λ~n({G~>0}r(X0)),\displaystyle\,+\tilde{\Lambda}\mathcal{L}_{n}(\{\tilde{G}>0\}\cap\mathcal{B}_{r}(X_{0})),

with σ=2Kωnδ\sigma=2K\omega_{n}\delta. On the other hand, since by (2.14)

(2.18) n+1|y|a|G|2dX=2dsi=1mλis(Ω),\int_{\mathbb{R}^{n+1}}|y|^{a}\big{\lvert}{\nabla G}\big{\rvert}^{2}\mathrm{d}X=\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega),

it follows

n+1|y|a|G~|2dX\displaystyle\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X =n+1|y|a|G|2dX+Br(X0)|y|a|G~|2dXBr(X0)|y|a|G|2dX\displaystyle=\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\int_{B_{r}(X_{0})}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X-\int_{B_{r}(X_{0})}|y|^{a}|\nabla G|^{2}\mathrm{d}X
=Br(X0)|y|a(|G~|2|G|2)dX+2dsi=1mλis(Ω)\displaystyle=\int_{B_{r}(X_{0})}|y|^{a}\left(|\nabla\tilde{G}|^{2}-|\nabla G|^{2}\right)\mathrm{d}X+\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)

from which we deduce that

Br(X0)|y|a|G|2dX+Λ~n({|G|>0}r(X0))\displaystyle\int_{B_{r}(X_{0})}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathcal{B}_{r}(X_{0}))\leq (1+σrn)Br(X0)|y|a|G~|2dX+\displaystyle\,\left(1+\sigma r^{n}\right)\int_{B_{r}(X_{0})}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X+
+Λ~n({|G~|>0}r(X0))\displaystyle\,+\tilde{\Lambda}\mathcal{L}_{n}(\{|\tilde{G}|>0\}\cap\mathcal{B}_{r}(X_{0}))
+σ2dsi=1mλis(Ω)rn.\displaystyle\,+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)r^{n}.

Hence, given GH1,a(n+1;m),Λ~>0G\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}),\tilde{\Lambda}>0 let us consider the functional

𝒥(G)=n+1|y|a|G|2dX+Λ~n({|G|>0}n)\mathcal{J}(G)=\int_{\mathbb{R}^{n+1}}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathbb{R}^{n})

and its localized version

𝒥(G,Br(X0))=Br(X0)|y|a|G|2dX+Λ~n({|G|>0}r(X0)).\mathcal{J}(G,B_{r}(X_{0}))=\int_{B_{r}(X_{0})}|y|^{a}|\nabla G|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G}\big{\rvert}>0\}\cap\mathcal{B}_{r}(X_{0})).

Finally, we end the Section with the following statement, where we collect all the previous formulation. In the rest of the paper, we will always refer to the following two almost-minimality conditions.

Theorem 2.2.

Set Λ>0\Lambda>0 and let ΩD\Omega\subset D be a shape optimizer to (1.1) and G=(g1,,gm)HΩ1,a(n+1;m)G=(g^{1},\dots,g^{m})\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega.
Then, given Λ~=2Λ/ds\tilde{\Lambda}=2\Lambda/d_{s}, it holds the following almost-minimality condition in n+1\mathbb{R}^{n+1}: for every δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} there exist K,ε>0K,\varepsilon>0 such that

(2.19) 𝒥(G)(1+KG~GL1(n))𝒥(G~)\mathcal{J}(G)\leq\left(1+K\|{\tilde{G}-G}\|_{L^{1}(\mathbb{R}^{n})}\right)\mathcal{J}(\tilde{G})

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(n)δ\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta and GG~L1(n)ε\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon.
Similarly, under the same assumptions, the vector GG satisfies the following localized almost-minimality condition: for δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} there exist σ,r0>0\sigma,r_{0}>0 such that, for every r(0,r0]r\in(0,r_{0}] and X0n×{0}X_{0}\in\mathbb{R}^{n}\times\{0\}

(2.20) 𝒥(G,Br(X0))(1+σrn)𝒥(G~,Br(X0))+σ2dsi=1mλis(Ω)rn,\mathcal{J}(G,B_{r}(X_{0}))\leq\left(1+\sigma r^{n}\right)\mathcal{J}(\tilde{G},B_{r}(X_{0}))+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)r^{n},

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(r(X0))δ\|{\tilde{G}}\|_{L^{\infty}(\mathcal{B}_{r}(X_{0}))}\leq\delta and GG~H01,a(Br(X0);m)G-\tilde{G}\in H^{1,a}_{0}(B_{r}(X_{0});\mathbb{R}^{m}).

Remark 2.4.

We remark that the localized almost-minimality condition (2.20) scales according to the C0,sC^{0,s} rescaling. Indeed, if GG satisfies (2.20), then for every X0n×{0},ρ>0X_{0}\in\mathbb{R}^{n}\times\{0\},\rho>0

GX0,ρ(X)=1ρsG(X0+ρX)HΩX0,ρ1,a(n+1;m),with ΩX0,ρ=ΩX0ρ,G_{X_{0},\rho}(X)=\frac{1}{\rho^{s}}G(X_{0}+\rho X)\in H^{1,a}_{\Omega_{X_{0},\rho}}(\mathbb{R}^{n+1};\mathbb{R}^{m}),\quad\mbox{with }\,\Omega_{X_{0},\rho}=\frac{\Omega-X_{0}}{\rho},

satisfies the same condition with constants σρn\sigma\rho^{n} and r0/ρr_{0}/\rho. More precisely, for every r(0,r0/ρ]r\in(0,r_{0}/\rho]

𝒥(GX0,ρ,Br)(1+(σρn)rn)𝒥(G~,Br)+σ2dsi=1mλis(Ω)rn,\mathcal{J}(G_{X_{0},\rho},B_{r})\leq\left(1+(\sigma\rho^{n})r^{n}\right)\mathcal{J}(\tilde{G},B_{r})+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)r^{n},

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(r)δρs\|{\tilde{G}}\|_{L^{\infty}(\mathcal{B}_{r})}\leq\delta\rho^{-s} and GX0,ρG~H01,a(Br;m)G_{X_{0},\rho}-\tilde{G}\in H^{1,a}_{0}(B_{r};\mathbb{R}^{m}).

3. C0,sC^{0,s}- regularity and non-degeneracy of eigenfunctions

In this Section we start by focusing on some local properties of shape optimizers. In particular we obtain optimal regularity and non-degeneracy of the vector of normalized eigenfunctions. As a consequence, we prove that shape optimizers are open sets and they satisfy some density estimates.
In the end, we discuss the validity of the unique continuation principle for eigenfunctions on disconnected domain.

3.1. Optimal Regularity

The purpose of this subsection is to obtain C0,sC^{0,s}-regularity of the eigenfunction associated to a solution Ω\Omega of (1.1).
The following is a technical lemma about a decay type estimate for the gradient of LaL_{a}-harmonic functions.

Lemma 3.1.

Let uH1,a(B1)u\in H^{1,a}(B_{1}) be LaL_{a}-harmonic in B1B_{1} and symmetric with respect to {y=0}\{y=0\}. Then, there exists C=C(n,a)>0C=C(n,a)>0 such that

Br|y|a|uu(0)|2dXCr2B1|y|a|u|2dX,\fint_{B_{r}}|y|^{a}|\nabla u-\nabla u(0)|^{2}\mathrm{d}X\leq Cr^{2}\fint_{B_{1}}|y|^{a}|\nabla u|^{2}\mathrm{d}X,

for every r(0,1/2]r\in(0,1/2].

Proof.

By the Poincaré-type inequality due to Fabes, Kenig and Serapioni [22, Theorem 1.5], there exists C1=C1(n,a)>0C_{1}=C_{1}(n,a)>0 such that the following inequality holds

Br|y|a|uu¯Br|2dXC1r2Br|y|a|u|2dX\int_{B_{r}}|y|^{a}|u-\bar{u}_{B_{r}}|^{2}\mathrm{d}X\leq C_{1}r^{2}\int_{B_{r}}|y|^{a}|\nabla u|^{2}\mathrm{d}X

where u¯Br\bar{u}_{B_{r}} is the average of uu in the ball BrB_{r}, that is

u¯Br=1rn+1+a|B1|aBr|y|audX,with |B1|a=B1|y|adX.\bar{u}_{B_{r}}=\frac{1}{r^{n+1+a}|B_{1}|_{a}}\int_{B_{r}}|y|^{a}u\mathrm{d}X,\quad\mbox{with }|B_{1}|_{a}=\int_{B_{1}}|y|^{a}\mathrm{d}X.

By the mean value formula for LaL_{a}-harmonic function, we first get u¯Br=u(0)\bar{u}_{B_{r}}=u(0).
Then, since by [11, Theorem 2.6] we know that

Br|y|a|u|2dX(rR)n+1+aBR|y|a|u|2dX\int_{B_{r}}|y|^{a}|\nabla u|^{2}\mathrm{d}X\leq\left(\frac{r}{R}\right)^{n+1+a}\int_{B_{R}}|y|^{a}|\nabla u|^{2}\mathrm{d}X

for 0rR10\leq r\leq R\leq 1, by combining the previous inequalities with a Caccioppoli type estimate, for r1/2r\leq 1/2 we get

Br|y|a|uu(0)|2dX\displaystyle\int_{B_{r}}|y|^{a}|u-u(0)|^{2}\mathrm{d}X C1r2Br|y|a|u|2dX\displaystyle\leq C_{1}r^{2}\int_{B_{r}}|y|^{a}|\nabla u|^{2}\mathrm{d}X
C1r2(r1/2)n+1+aB1/2|y|a|u|2dX\displaystyle\leq C_{1}r^{2}\left(\frac{r}{1/2}\right)^{n+1+a}\int_{B_{1/2}}|y|^{a}|\nabla u|^{2}\mathrm{d}X
C2rn+1+a+2B1|y|au2dX,\displaystyle\leq C_{2}r^{n+1+a+2}\int_{B_{1}}|y|^{a}u^{2}\mathrm{d}X,

where C2=C2(n,a)C_{2}=C_{2}(n,a). Moreover, since for every i=1,,ni=1,\dots,n the derivative xiu\partial_{x_{i}}u is LaL_{a}-harmonic, we immediately get

(3.1) Br|y|a|xu(xu)(0)|2dXC1(n,a)r2B1|y|a|xu|2dX,\fint_{B_{r}}|y|^{a}|\nabla_{x}u-(\nabla_{x}u)(0)|^{2}\mathrm{d}X\leq C_{1}(n,a)r^{2}\fint_{B_{1}}|y|^{a}|\nabla_{x}u|^{2}\mathrm{d}X,

for every r(0,1/2]r\in(0,1/2]. On the other hand, the covariant derivative associated to the yy-direction

yau={|y|ayuif y00if y=0,\partial^{a}_{y}u=\begin{cases}|y|^{a}\partial_{y}u&\mbox{if }y\neq 0\\ 0&\mbox{if }y=0\end{cases},

is LaL_{-a}-harmonic, and consequently

(3.2) Br|y|a(yau)2dXC1(n,a)r2B1|y|a(yau)2dX.\fint_{B_{r}}|y|^{-a}(\partial^{a}_{y}u)^{2}\mathrm{d}X\leq C_{1}(n,-a)r^{2}\fint_{B_{1}}|y|^{-a}(\partial^{a}_{y}u)^{2}\mathrm{d}X.

Finally, summing (3.1) and (3.2) we get the claimed inequality with C=max{C1(n,a),C1(n,a)}C=\max\{C_{1}(n,a),C_{1}(n,-a)\}. ∎

The following result is a dichotomy which has been first proposed in the scalar case in [17, Proposition 2.1.] for the case of almost minimizer of the thin one-phase problem (see [16, Section 2] for a similar result in the local case).

Lemma 3.2.

Let σ>0,ε>0\sigma^{\prime}>0,\varepsilon>0 and GH1,a(B1;m)G\in H^{1,a}(B_{1};\mathbb{R}^{m}) be symmetric with respect to {y=0}\{y=0\} and such that

𝒥(G;B1)(1+σ)𝒥(G~;B1)+ε,\mathcal{J}(G;B_{1})\leq(1+\sigma^{\prime})\mathcal{J}(\tilde{G};B_{1})+\varepsilon,

where G~H1,a(B1;m)\tilde{G}\in H^{1,a}(B_{1};\mathbb{R}^{m}) satisfies G~L(1)GL(n)\big{\|}{\tilde{G}}\big{\|}_{L^{\infty}(\mathcal{B}_{1})}\leq\big{\|}{G}\big{\|}_{L^{\infty}(\mathbb{R}^{n})} and G~GH01,a(B1;m)\tilde{G}-G\in H^{1,a}_{0}(B_{1};\mathbb{R}^{m}). Thus, if we set

ai:=(B1|y|a|gi|2dX)1/2,a_{i}:=\left(\fint_{B_{1}}|y|^{a}\big{\lvert}{\nabla g^{i}}\big{\rvert}^{2}\mathrm{d}X\right)^{1/2},

there exists universal constants η(0,1),M,σ0>0\eta\in(0,1),M,\sigma_{0}>0 such that if σσ0\sigma^{\prime}\leq\sigma_{0}, then for every i=1,,mi=1,\dots,m

eitheraiMor(ηBη|y|a|gi|2dX)1/2ai2.\mbox{either}\quad a_{i}\leq M\quad\mbox{or}\quad\left(\eta\fint_{B_{\eta}}|y|^{a}\big{\lvert}{\nabla g^{i}}\big{\rvert}^{2}\mathrm{d}X\right)^{1/2}\leq\frac{a_{i}}{2}.
Proof.

Let g~iH1,a(n+1)L(n)\tilde{g}^{i}\in H^{1,a}(\mathbb{R}^{n+1})\cap L^{\infty}(\mathbb{R}^{n}) be the LaL_{a}-harmonic replacement of gig^{i} in B1B_{1}, that is

{Lag~i=0in B1g~i=gion B1.\begin{cases}L_{a}\tilde{g}^{i}=0&\mbox{in }B_{1}\\ \tilde{g}^{i}=g^{i}&\mbox{on }\partial B_{1}.\end{cases}

By the symmetry of the datum on B1\partial B_{1}, we know that g~i\tilde{g}^{i} is symmetric with respect to {y=0}\{y=0\}. By integration by parts, we easily deduce

(3.3) B1|y|ag~i,(gig~i)dX=0.\int_{B_{1}}|y|^{a}\langle\nabla\tilde{g}^{i},\nabla(g^{i}-\tilde{g}^{i})\mathrm{d}X=0.

Since g~iL(n)δ\big{\|}{\tilde{g}^{i}}\big{\|}_{L^{\infty}(\mathbb{R}^{n})}\leq\delta, consider now the admissible competitor G~=(g1,,g~i,,gm)\tilde{G}=(g^{1},\dots,\tilde{g}^{i},\dots,g^{m}). By the almost minimality condition, we first get

B1|y|a(|gi|2|g~i|2)dXσB1|y|a|G~|2dX+ε+Λ~ωn.\int_{B_{1}}|y|^{a}(|\nabla g^{i}|^{2}-|\nabla\tilde{g}^{i}|^{2})\mathrm{d}X\leq\sigma^{\prime}\int_{B_{1}}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X+\varepsilon+\tilde{\Lambda}\omega_{n}.

On the other hand, since

B1|y|a|G~|2dX+2dsi=1mλis(Ω)+B1|y|a(|g~i|2|gi|2)dX\int_{B_{1}}|y|^{a}|\nabla\tilde{G}|^{2}\mathrm{d}X\leq+\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)+\int_{B_{1}}|y|^{a}\left(|\nabla\tilde{g}^{i}|^{2}-|\nabla g^{i}|^{2}\right)\mathrm{d}X

we deduce, together with (3.3), that

B1|y|a|(gig~i)|2dXσB1|y|a|g~i|2dX+C~\int_{B_{1}}|y|^{a}|\nabla(g^{i}-\tilde{g}^{i})|^{2}\mathrm{d}X\leq\sigma^{\prime}\int_{B_{1}}|y|^{a}|\nabla\tilde{g}^{i}|^{2}\mathrm{d}X+\tilde{C}

with C~>0\tilde{C}>0. Since g~i\tilde{g}^{i} minimizes the Dirichlet’s type energy associated to the operator LaL_{a}, we get

B1|y|a|g~i|2dXai2,\fint_{B_{1}}|y|^{a}|\nabla\tilde{g}^{i}|^{2}\mathrm{d}X\leq a_{i}^{2},

and since g~i\tilde{g}^{i} is symmetric with respect to {y=0}\{y=0\}, by Lemma 3.1

Bη|y|a|g~ig~i(0)|2dXCη2ai2for η1/2.\fint_{B_{\eta}}|y|^{a}|\nabla\tilde{g}^{i}-\nabla\tilde{g}^{i}(0)|^{2}\mathrm{d}X\leq C\eta^{2}a^{2}_{i}\quad\mbox{for }\eta\leq 1/2.

On the other hand, since |xg~i||\nabla_{x}\tilde{g}^{i}| is LaL_{a}-subharmonic, by [41, Lemma A.2.] we get

xg~iL(B1/2)C0ai\big{\|}{\nabla_{x}\tilde{g}^{i}}\big{\|}_{L^{\infty}(B_{1/2})}\leq C_{0}a_{i}

for some C0C_{0} universal. Thus, if we denote q=g~i(0)=(xg~i(0),0)q=\nabla\tilde{g}^{i}(0)=(\nabla_{x}\tilde{g}^{i}(0),0) we conclude that |q|C0ai|q|\leq C_{0}a_{i}. Collecting the previous inequalities, for r(0,1/2]r\in(0,1/2] we first get

Br|y|a|giq|2dX2σai2rn1+2Cr2ai2+C¯rn1\fint_{B_{r}}|y|^{a}|\nabla g^{i}-q|^{2}\mathrm{d}X\leq 2\sigma^{\prime}a_{i}^{2}r^{-n-1}+2Cr^{2}a_{i}^{2}+\bar{C}r^{-n-1}

for some C¯>0\bar{C}>0 universal. Then, since |q|C0ai|q|\leq C_{0}a_{i}

(3.4) rBr|y|a|gi|2dX4σai2rn+4Cr3ai2+2C¯rn+2C02rai2.r\fint_{B_{r}}|y|^{a}|\nabla g^{i}|^{2}\mathrm{d}X\leq 4\sigma^{\prime}a_{i}^{2}r^{-n}+4Cr^{3}a_{i}^{2}+2\bar{C}r^{-n}+2C_{0}^{2}ra_{i}^{2}.

Now, if we choose rr small enough, such that

2C02r18,4Cr31242C_{0}^{2}r\leq\frac{1}{8},\quad 4Cr^{3}\leq\frac{1}{24}

and σ\sigma^{\prime} small so that

4σrn1244\sigma^{\prime}r^{-n}\leq\frac{1}{24}

we get the following dichotomy:

either2C¯rnai224or2C¯rn<ai224.\mbox{either}\quad 2\bar{C}r^{-n}\geq\frac{a_{i}^{2}}{24}\quad\mbox{or}\quad 2\bar{C}r^{-n}<\frac{a_{i}^{2}}{24}.

Hence, in the first case it follows the first alternative of the claimed dichotomy and in the second one, if we combine (3.4) with the choices of rr and σ\sigma^{\prime}, we provide the bounds of the second alternative. ∎

We can now state our main regularity result for the normalized eigenfunctions associated to the shape minimization problem (1.1).

Proposition 3.1.

Let DnD\subset\mathbb{R}^{n} be open and bounded and ΩD\Omega\subset D be a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega. Let r0>0r_{0}>0 be the constant associated to δ=GL(n)\delta=\|G\|_{L^{\infty}(\mathbb{R}^{n})} in the localized almost-minimality condition (2.20). Then, GCloc0,s(n+1;m)G\in C^{0,s}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m}) and

(3.5) [gi]C0,s(Br0/2(X0))C(1+λis(Ω)),[g^{i}]_{C^{0,s}(B_{r_{0}/2}(X_{0}))}\leq C\left(1+\lambda_{i}^{s}(\Omega)\right),

for every X0n×{0},i=1,,mX_{0}\in\mathbb{R}^{n}\times\{0\},i=1,\dots,m, with C>0C>0 universal constant.

Proof.

Let δ=GL(Ω)\delta=\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} and η,M,σ0\eta,M,\sigma_{0} be the constants from Lemma 3.2. Fix σ,r0>0\sigma,r_{0}>0 as the constants of the almost-minimality condition (2.20) associated to δ\delta. By the scaling invariance, it is not restrictive to assume that r0=1r_{0}=1 and σσ0\sigma\leq\sigma_{0}: if not, consider

GX0,ρ(X)=1ρsG(X0+ρX)G_{X_{0},\rho}(X)=\frac{1}{\rho^{s}}G(X_{0}+\rho X)

with X0n×{0}X_{0}\in\mathbb{R}^{n}\times\{0\} and ρ(0,r0]\rho\in(0,r_{0}]. By Remark 2.4, for every r(0,r0/ρ]r\in(0,r_{0}/\rho]

𝒥(GX0,ρ,Br)(1+(σρn)rn)𝒥(G~,Br)+σ2dsi=1mλis(Ω)rn,\mathcal{J}(G_{X_{0},\rho},B_{r})\leq\left(1+(\sigma\rho^{n})r^{n}\right)\mathcal{J}(\tilde{G},B_{r})+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)r^{n},

for every admissible G~H1,a(Br;m)\tilde{G}\in H^{1,a}(B_{r};\mathbb{R}^{m}). By choosing σ=σ0\sigma=\sigma_{0} and ρ=r0\rho=r_{0} we get

𝒥(GX0,r0,B1)(1+σ)𝒥(G~,B1)+ε,\mathcal{J}(G_{X_{0},r_{0}},B_{1})\leq\left(1+\sigma^{\prime}\right)\mathcal{J}(\tilde{G},B_{1})+\varepsilon,

with σ=σ0r0nσ0\sigma^{\prime}=\sigma_{0}r_{0}^{n}\leq\sigma_{0} and ε=σ2ds1i=1mλis(Ω)\varepsilon=\sigma 2d_{s}^{-1}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega). Thus, for i=1,,mi=1,\dots,m, set

ai(η)=(ηBη|y|a|gi|2dX)1/2.a_{i}(\eta)=\left(\eta\fint_{B_{\eta}}|y|^{a}\big{\lvert}{\nabla g^{i}}\big{\rvert}^{2}\mathrm{d}X\right)^{1/2}.

We first show by induction that for all N0N\geq 0 the following inequality holds

(3.6) ai(ηN)C(η)M+2Nai(1),a_{i}(\eta^{N})\leq C(\eta)M+2^{-N}a_{i}(1),

with C(η)C(\eta) a large constant. By the previous observations, we can apply Lemma 3.2 which gives the desired inequality for N=0N=0. Thus, let us suppose that it holds for NN and let us show that it holds also for the value N+1N+1. By the scaling invariance of Remark 2.4, if we rescaled Lemma 3.2 we get

eitherai(ηN)Morai(ηN+1)12ai(ηN).\mbox{either}\quad a_{i}(\eta^{N})\leq M\quad\mbox{or}\quad a_{i}(\eta^{N+1})\leq\frac{1}{2}a_{i}(\eta^{N}).

From the last alternative it follows immediately the validity of (3.6) for the value N+1N+1. Instead, in the first case, we can use that

ai(ηN+1)ηn+a2ai(ηN)C(η)M,a_{i}(\eta^{N+1})\leq\eta^{-\frac{n+a}{2}}a_{i}(\eta^{N})\leq C(\eta)M,

with C(η)=ηn+a2C(\eta)=\eta^{-\frac{n+a}{2}}. Finally ai(r)C(1+ai(1))a_{i}(r)\leq C(1+a_{i}(1)) for every r(0,1)r\in(0,1), namely

Br|y|a|gi|2dXC(1+λis(Ω))2rn.\int_{B_{r}}|y|^{a}|\nabla g^{i}|^{2}\mathrm{d}X\leq C(1+\lambda_{i}^{s}(\Omega))^{2}r^{n}.

Finally, by applying Morrey’s type embedding we get the claimed estimate. ∎

Since by (2.14) and (2.18), the vector GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) is bounded in H1,a(n+1;m)H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}), we can slightly simplify the almost-minimality condition (2.19) as follows:

for every δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} there exist K,ε>0K,\varepsilon>0 such that

(3.7) 𝒥(G)𝒥(G~)+K2dsi=1mλis(Ω)G~GL1(n)\mathcal{J}(G)\leq\mathcal{J}(\tilde{G})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G}\|_{L^{1}(\mathbb{R}^{n})}

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(n)δ\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta and GG~L1(n)ε\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon.

Similarly, we can change condition (2.20) as well:

for every δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} there exist σ,r0>0\sigma,r_{0}>0 such that, for every r(0,r0]r\in(0,r_{0}] and X0n×{0}X_{0}\in\mathbb{R}^{n}\times\{0\}

(3.8) 𝒥(G,Br(X0))𝒥(G~,Br(X0))+σ2dsi=1mλis(Ω)G~GL1(r(X0)).\mathcal{J}(G,B_{r}(X_{0}))\leq\mathcal{J}(\tilde{G},B_{r}(X_{0}))+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G}\|_{L^{1}(\mathcal{B}_{r}(X_{0}))}.

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(r(X0))δ\|{\tilde{G}}\|_{L^{\infty}(\mathcal{B}_{r}(X_{0}))}\leq\delta and GG~H01,a(Br(X0);m)G-\tilde{G}\in H^{1,a}_{0}(B_{r}(X_{0});\mathbb{R}^{m}).

It is more convenient to work with these alternative formulations since in these conditions the energies simplify in the regions where the competitor and the solution are equal.

Remark 3.2.

This new almost-minimality condition (3.8) holds true also if we remove the assumption on the upper bound in H1,a(+n+1;m)H^{1,a}(\mathbb{R}^{n+1}_{+};\mathbb{R}^{m}), due to (2.18). Precisely, as in [17], by combining the regularity result Proposition 3.1 and the validity of a Caccioppoli type inequality for every components of GG, we can conclude that it is still uniformly bounded in H1,a(Br;m)H^{1,a}(B_{r};\mathbb{R}^{m}) for every r(0,r0]r\in(0,r_{0}].

Remark 3.3.

By the previous result, we already know that {|G|>0}n\{|G|>0\}\cap\mathbb{R}^{n} and more generally {gi0}n\{g^{i}\neq 0\}\cap\mathbb{R}^{n} are open subsets of n\mathbb{R}^{n}. Moreover, as we see in Proposition 3.10, by the unique continuation principle we get

n(({|G|>0}n)({gi0}n))=0,\mathcal{L}_{n}\left(\left(\{|G|>0\}\cap\mathbb{R}^{n}\right)\setminus\left(\{g^{i}\neq 0\}\cap\mathbb{R}^{n}\right)\right)=0,

for every i=1,,mi=1,\dots,m.

3.2. Non-degeneracy

The non-degeneracy of solutions near the free boundary points will allow to understand the measure-theoretic structure of the free boundary via a blow-up analysis.
Inspired by [17, 16], our analysis is mainly based on the following Lemma, in which we compare every components of GG with its LaL_{a}-harmonic replacement.

Lemma 3.3.

Given ΩD\Omega\subset D a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) the vector of normalized eigenfunctions on Ω\Omega, let us assume that B1{|G|>0}B_{1}\subset\{|G|>0\}.
Then, for every i=1,,mi=1,\dots,m, given viv^{i} the LaL_{a}-harmonic replacement of gig^{i} in B7/8B_{7/8}, that is,

{Lavi=0in B7/8vi=gion B1B7/8.\begin{cases}L_{a}v^{i}=0&\mbox{in }B_{7/8}\\ v^{i}=g^{i}&\mbox{on }B_{1}\setminus B_{7/8}.\end{cases}

we get

(3.9) giviL(B1/2)ω(σ),\big{\|}{g^{i}-v^{i}}\big{\|}_{L^{\infty}(B_{1/2})}\leq\omega(\sigma),

where ω(σ)0\omega(\sigma)\to 0 as σ0\sigma\to 0.

Proof.

Fixed i=1,,mi=1,\dots,m, consider the competitor G~=(g1,,vi,,gm)\tilde{G}=(g^{1},\dots,v^{i},\dots,g^{m}) with viv^{i} is the LaL_{a}-harmonic replacement of gig^{i} in B7/8B_{7/8}. Then, by the localized almost-minimality condition (3.8)

𝒥(G,B1)𝒥(G~,B1)+σ2dsi=1mλis(Ω)G~GL1(1)\mathcal{J}(G,B_{1})\leq\mathcal{J}(\tilde{G},B_{1})+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G}\|_{L^{1}(\mathcal{B}_{1})}

and by Remark 3.3 we first deduce

B7/8|y|a|gi|2dXB7/8|y|a|vi|2dXσ2dsi=1mλis(Ω)G~GL1(1).\int_{B_{7/8}}|y|^{a}\big{\lvert}{\nabla g^{i}}\big{\rvert}^{2}\mathrm{d}X-\int_{B_{7/8}}|y|^{a}\big{\lvert}{\nabla v^{i}}\big{\rvert}^{2}\mathrm{d}X\leq\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G}\|_{L^{1}(\mathcal{B}_{1})}.

and then, by exploiting the harmonicity of viv^{i}, we get

B7/8|y|a|(givi)|2dXσ2dsi=1mλis(Ω)G~GL1(1).\int_{B_{7/8}}|y|^{a}\big{\lvert}{\nabla(g^{i}-v^{i})}\big{\rvert}^{2}\mathrm{d}X\leq\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G}\|_{L^{1}(\mathcal{B}_{1})}.

By Poincaré inequality, there exists C>0C>0 depending on nn and s(0,1)s\in(0,1) such that

B3/4|y|a(givi)2dXCσ,\int_{B_{3/4}}|y|^{a}(g^{i}-v^{i})^{2}\mathrm{d}X\leq C\sigma,

with givig^{i}-v^{i} uniformly C0,sC^{0,s} in B3/4B_{3/4}. Finally, if at some point XB1/2X\in B_{1/2} we have (givi)(X)M(g^{i}-v^{i})(X)\geq M, then

Mn+2Cσ,M^{n+2}\leq C\sigma,

for some C>0C>0. Thus (3.9) holds with ω(σ)=Cσ1/(n+2)\omega(\sigma)=C\sigma^{1/(n+2)}. ∎

Remark 3.4.

The previous result allows to overcome the absence of a minimality condition by comparing the almost-minimizer with its harmonic replacement. Nevertheless, since each eigenfunction gig^{i} solves (2.13) for λ=λis(Ω)\lambda=\lambda_{i}^{s}(\Omega), we can easily compute that |G|=|(g1,,gm)|HΩ1,a(+n+1)|G|=|(g^{1},\dots,g^{m})|\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1}_{+}) is a weak solution to

{La|G|0in +n+1ya|G|λms(Ω)|G|on Ω.\begin{cases}-L_{a}|G|\leq 0&\mbox{in }\mathbb{R}^{n+1}_{+}\\ -\partial^{a}_{y}|G|\leq\lambda_{m}^{s}(\Omega)|G|&\mbox{on }\Omega.\end{cases}

Indeed, it holds

La|G|\displaystyle L_{a}|G| =i=1m(1|G|giLagi+|y|a|gi|2|G||y|agigi,|G||G|2)\displaystyle=\sum_{i=1}^{m}\left(\frac{1}{|G|}g^{i}L_{a}g^{i}+|y|^{a}\frac{\big{\lvert}{\nabla g^{i}}\big{\rvert}^{2}}{|G|}-|y|^{a}\frac{g^{i}\langle\nabla g^{i},\nabla|G|\rangle}{|G|^{2}}\right)
=|y|a|G|3i,j((gj)2|gi|2gigjgi,gj)0,\displaystyle=\frac{|y|^{a}}{|G|^{3}}\sum_{i,j}\left((g^{j})^{2}\big{\lvert}{\nabla g^{i}}\big{\rvert}^{2}-g^{i}g^{j}\langle\nabla g^{i},\nabla g^{j}\rangle\right)\geq 0,

weakly in H1,a(+n+1)H^{1,a}(\mathbb{R}^{n+1}_{+}). Moreover, by (2.6) we get |G|L(n)m[C~n,sλms(Ω)]n4s\big{\|}{|G|}\big{\|}_{L^{\infty}(\mathbb{R}^{n})}\leq\sqrt{m}\left[\tilde{C}_{n,s}\lambda^{s}_{m}(\Omega)\right]^{\frac{n}{4s}} and consequently, by taking the symmetric extension through {y=0}\{y=0\}, we get

La(|G|+mC~n,sn/4s1aλms(Ω)n+4s4s|y|1a)0in +n+1.-L_{a}\left(|G|+\frac{\sqrt{m}\tilde{C}_{n,s}^{n/4s}}{1-a}\lambda^{s}_{m}(\Omega)^{\frac{n+4s}{4s}}|y|^{1-a}\right)\leq 0\quad\mbox{in $\mathbb{R}^{n+1}_{+}$}.

Similarly, we have

La(|gi|+C~n,sn/4s1aλis(Ω)n+4s4s|y|1a)0in +n+1-L_{a}\left(|g^{i}|+\frac{\tilde{C}_{n,s}^{n/4s}}{1-a}\lambda^{s}_{i}(\Omega)^{\frac{n+4s}{4s}}|y|^{1-a}\right)\leq 0\quad\mbox{in $\mathbb{R}^{n+1}_{+}$}

for every i=1,,mi=1,\dots,m.

We start by proving the following weak non-degeneracy condition.

Proposition 3.4.

Let ΩD\Omega\subset D be a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega. Set r0>0r_{0}>0 to be the constant associated to δ=GL(n)\delta=\|G\|_{L^{\infty}(\mathbb{R}^{n})} in the localized almost-minimality condition (3.8).
Then, there exists a universal constant c2>0c_{2}>0 such that

(3.10) |G|(X)c2dist(X,{|G|>0})sin r0/2+(G).\big{\lvert}{G}\big{\rvert}(X)\geq c_{2}\mathrm{dist}(X,\partial\{\big{\lvert}{G}\big{\rvert}>0\})^{s}\quad\text{in $\mathcal{B}_{r_{0}/2}^{+}(G)$}.
Proof.

By Remark 2.4, up to translation and rescaling it is enough to show that if GG satisfies (3.8) for σ>0\sigma>0 sufficiently small in a large ball and

(3.11) dist(0,{|G|>0})=1,\mathrm{dist}(0,\partial\{\big{\lvert}{G}\big{\rvert}>0\})=1,

then |G|(0)c2>0|G|(0)\geq c_{2}>0 for some c2c_{2} small to be made precise later.
Indeed, assume 1{|G|>0}\mathcal{B}_{1}\subset\{|G|>0\} and for every i=1,,mi=1,\dots,m set viv^{i} as the LaL_{a}-harmonic replacement of gig^{i} in B7/8B_{7/8}: according to Lemma 3.3 the result follows once we prove the claimed statement for the vector V=(v1,,vm)V=(v^{1},\dots,v^{m}). Now, the viv^{i}’s are uniformly bounded say in B3/4B_{3/4} and hence, since they are LaL_{a}-harmonic and symmetric with respect to {y=0}\{y=0\}, we get

|vi(X)vi(0)|K|X|,in B1/2,|v^{i}(X)-v^{i}(0)|\leq K|X|,\quad\text{in $B_{1/2}$},

for K>0K>0 universal. Thus,

vi(X)vi(0)+K|X|,in B1/2.v^{i}(X)\leq v^{i}(0)+K|X|,\quad\text{in $B_{1/2}$.}

Let

Vδ(X)=1δsV(δX),Gδ(X)=1δsG(δX),for XB1V_{\delta}(X)=\frac{1}{\delta^{s}}V(\delta X),\quad G_{\delta}(X)=\frac{1}{\delta^{s}}G(\delta X),\quad\mbox{for }X\in B_{1}

with δ>0\delta>0 universal to be chosen universal later. Then, we get

vδivi(0)δs+Kδ1sCvδi(0)in B1,v^{i}_{\delta}\leq v^{i}(0)\delta^{-s}+K\delta^{1-s}\leq Cv_{\delta}^{i}(0)\quad\text{in $B_{1}$},

for every i=1,,mi=1,\dots,m and δ\delta small enough. Moreover, since the vδiv^{i}_{\delta}’s are LaL_{a}-harmonic in B1B_{1}, the bound above and the validity of mean value formulas for the LaL_{a}-operator imply

vδiL(B1),xvδiL(B1/2),yavδiL(B1/2)Cvδi(0),\|v^{i}_{\delta}\|_{L^{\infty}(B_{1})},\|\nabla_{x}v^{i}_{\delta}\|_{L^{\infty}(B_{1/2})},\|\partial^{a}_{y}v^{i}_{\delta}\|_{L^{\infty}(B_{1/2})}\leq Cv^{i}_{\delta}(0),

with C>0C>0 universal (possibly changing from line to line). Let φC0(B1/2),0φ1\varphi\in C_{0}^{\infty}(B_{1/2}),0\leq\varphi\leq 1 such that φ1\varphi\equiv 1 in B1/4B_{1/4}, then

B1|y|a|vδi|2dXB1|y|a|(vδi(1φ))|2dXC(vδi(0))2\int_{B_{1}}|y|^{a}|\nabla v^{i}_{\delta}|^{2}\mathrm{d}X\geq\int_{B_{1}}|y|^{a}|\nabla(v^{i}_{\delta}(1-\varphi))|^{2}\mathrm{d}X-C(v_{\delta}^{i}(0))^{2}

and on the other hand

n({|Vδ|>0}B1)n({|Vδ|(1φ)>0}B1)+C0.\mathcal{L}_{n}(\{|V_{\delta}|>0\}\cap B_{1})\geq\mathcal{L}_{n}(\{|V_{\delta}|(1-\varphi)>0\}\cap B_{1})+C_{0}.

In conclusion, by the minimality of VδV_{\delta} with respect to its Dirichlet-type energy, we first deduce

𝒥(Vδ,B1)𝒥(Gδ,B1)𝒥(Vδ(1φ),B1)+σCn\mathcal{J}(V_{\delta},B_{1})\leq\mathcal{J}(G_{\delta},B_{1})\leq\mathcal{J}(V_{\delta}(1-\varphi),B_{1})+\sigma C_{n}

and then

C|Vδ|2(0)Λ~C0σCnC|V_{\delta}|^{2}(0)\geq\tilde{\Lambda}C_{0}-\sigma C_{n}

Finally, we reach the claim for σ>0\sigma>0 sufficiently small. ∎

Finally, we are able to prove the strong non-degeneracy near the free boundary.

Proposition 3.5.

Consider ΩD\Omega\subset D a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) the vector of normalized eigenfunctions on Ω\Omega.
Let r0>0r_{0}>0 as in Proposition 3.1, then if X0F(G)X_{0}\in F(G), for every r(0,r0/2)r\in(0,r_{0}/2) we have

(3.12) supr(X0)|G|crs,\sup_{\mathcal{B}_{r}(X_{0})}\big{\lvert}{G}\big{\rvert}\geq cr^{s},

for some universal constant c>0,r0>0c>0,r_{0}>0.

In view of Proposition 3.1 and Proposition 3.4, Proposition 3.5 follows by applying the next lemma to the function

f(X)=|G|(X)+mC~n,sn/4s1aλms(Ω)n+4s4s|y|1a,f(X)=|G|(X)+\frac{\sqrt{m}\tilde{C}_{n,s}^{n/4s}}{1-a}\lambda^{s}_{m}(\Omega)^{\frac{n+4s}{4s}}|y|^{1-a},

introduced in Remark 3.4.

Lemma 3.6.

Let f0f\geq 0 be defined in B1B_{1} and subharmonic in B1+(f).B^{+}_{1}(f). Assume that there is a small constant η>0\eta>0 such that

(3.13) fC0,s(B1)η1,\|f\|_{C^{0,s}(B_{1})}\leq\eta^{-1},

and ff satisfies the non-degeneracy condition on 1\mathcal{B}_{1},

(3.14) f(X)ηdist(X,{f=0})sfor every X1.f(X)\geq\eta\;\mathrm{dist}(X,\{f=0\})^{s}\quad\text{for every }X\in\mathcal{B}_{1}.

Then if 0F(f)0\in F(f), we get

suprfc(η)rs,for r1.\sup_{\mathcal{B}_{r}}f\geq c(\eta)\;r^{s},\quad\text{for }r\leq 1.
Proof.

The proof follows the lines of [11, Proposition 3.3](see also [19, Lemma 2.9]).
Given a point X01+(f)X_{0}\in\mathcal{B}_{1}^{+}(f), to be chosen close enough to 0F(f)0\in F(f), we construct a sequence of points (Xk)k1(X_{k})_{k}\subset\mathcal{B}_{1} such that

f(Xk+1)(1+δ)f(Xk),|Xk+1Xk|C(η)dist(Xk,{f=0}),f(X_{k+1})\geq(1+\delta)f(X_{k}),\quad|X_{k+1}-X_{k}|\leq C(\eta)\mathrm{dist}(X_{k},\{f=0\}),

with δ\delta small depending on η\eta.

Then, using (3.14) and that (f(Xk))k(f(X_{k}))_{k} grows geometrically, we find

|Xk+1X0|\displaystyle|X_{k+1}-X_{0}| i=0k|Xi+1Xi|C(η)i=0kdist(Xi,{f=0})\displaystyle\leq\sum_{i=0}^{k}|X_{i+1}-X_{i}|\leq C(\eta)\sum_{i=0}^{k}\mathrm{dist}(X_{i},\{f=0\})
C(η)η2i=0kf(Xi)1/sc(η)f(Xk+1)1/s.\displaystyle\leq\frac{C(\eta)}{\eta^{2}}\sum_{i=0}^{k}f(X_{i})^{1/s}\leq c(\eta)f(X_{k+1})^{1/s}.

Hence for a sequence of radii rk=dist(Xk,{f=0}),r_{k}=\mathrm{dist}(X_{k},\{f=0\}),we have that

suprk(X0)fcrks\sup_{\mathcal{B}_{r_{k}}(X_{0})}f\geq cr_{k}^{s}

from which we obtain that

supr(X0)fcrs,for all r|X0|.\sup_{\mathcal{B}_{r}(X_{0})}f\geq cr^{s},\quad\text{for all $r\geq|X_{0}|.$}

The conclusion follows by letting X0X_{0} go to 0F(f)0\in F(f).

We now show that the sequence of XkX_{k}’s exists. After scaling, assume we constructed XkX_{k} such that

f(Xk)=1.f(X_{k})=1.

Let us call with YkF(f)Y_{k}\in F(f) the point where the distance from XkX_{k} to {f=0}\{f=0\} is achieved. By (3.13) and (3.14), we get

c(η)rk=|XkYk|C(η).c(\eta)\leq r_{k}=|X_{k}-Y_{k}|\leq C(\eta).

Assume by contradiction that we cannot find Xk+1X_{k+1} in M(Xk)\mathcal{B}_{M}(X_{k}) with MM large to be specified later, such that

f(Xk+1)1+δ.f(X_{k+1})\geq 1+\delta.

Then f1+δ+wf\leq 1+\delta+w with ww a LaL_{a}-harmonic function in BM+(Xk)B_{M}^{+}(X_{k}) such that

w=0on {y=0},w=fon BM(Xk){y>0}.w=0\quad\text{on $\{y=0\}$},\quad w=f\quad\text{on $\partial B_{M}(X_{k})\cap\{y>0\}$}.

Thus, we have

wC(n)y2sMsupBM+(Xk)fCη1y2sMs1δin Brk(Xk),w\leq C(n)\frac{y^{2s}}{M}\sup_{B_{M}^{+}(X_{k})}f\leq C\eta^{-1}y^{2s}M^{s-1}\leq\delta\quad\text{in $B_{r_{k}}(X_{k}),$}

for MM sufficiently large depending on δ\delta. Thus,

(3.15) f1+2δin Brk(Xk).f\leq 1+2\delta\quad\text{in $B_{r_{k}}(X_{k}).$}

On the other hand, f(Yk)=0,YkBrk(Xk)f(Y_{k})=0,Y_{k}\in\partial B_{r_{k}}(X_{k}). Thus from the Hölder continuity of ff we find

(3.16) f12,in Bc(η)(Yk).f\leq\frac{1}{2},\quad\text{in $B_{c(\eta)}(Y_{k})$}.

If δ\delta is sufficiently small (3.15)-(3.16) contradict that

1=f(Xk)Brk(Xk)|y|afdX.1=f(X_{k})\leq\fint_{B_{r_{k}}(X_{k})}|y|^{a}f\mathrm{d}X.

Before stating the main result on density estimates, we need the following lemma in which we prove a compactness result for sequence of almost-minimizers uniformly bounded in H1,a(n+1;m)H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}).

Lemma 3.7.

Let (Gk)k(G_{k})_{k} be a sequence of almost-minimizer in B1B_{1} in the sense of condition (3.8), uniformly bounded in H1,a(B1;m)H^{1,a}(B_{1};\mathbb{R}^{m}). Then, up to a subsequence, there exists a limit function GG_{\infty} such that

  • GHloc1,a(B1;m)Cloc0,s(B1¯;m)G_{\infty}\in H^{1,a}_{{{\tiny{\mbox{loc}}}}}(B_{1};\mathbb{R}^{m})\cap C^{0,s}_{{\tiny{\mbox{loc}}}}(\overline{B_{1}};\mathbb{R}^{m});

  • GkGG_{k}\to G_{\infty} in Cloc0,α(B1¯;m)C^{0,\alpha}_{{\tiny{\mbox{loc}}}}(\overline{B_{1}};\mathbb{R}^{m}), for every α(0,s)\alpha\in(0,s);

  • GkGG_{k}\rightharpoonup G_{\infty} weakly in Hloc1,a(B1;m)H^{1,a}_{{\tiny{\mbox{loc}}}}(B_{1};\mathbb{R}^{m});

  • GG_{\infty} is an almost-minimizer in B1B_{1} in the sense of condition (3.8).

Proof.

By Proposition 3.1 we already know that GkGG_{k}\to G_{\infty} uniformly on every compact set of B1B_{1} and so in Cloc0,α(B¯)C^{0,\alpha}_{{\tiny{\mbox{loc}}}}(\overline{B}), for every α(0,s)\alpha\in(0,s). Moreover, by Ascoli-Arzelá theorem it follows that GC0,s(B¯)G_{\infty}\in C^{0,s}(\overline{B}).
By assumptions, the sequence is uniformly bounded in H1,a(B1;m)H^{1,a}(B_{1};\mathbb{R}^{m}) and sot it weakly converges to some GH1,a(B1;m)G_{\infty}\in H^{1,a}(B_{1};\mathbb{R}^{m}).

In conclusion, let us show that for every δGL(Ω)\delta\geq\big{\|}{G_{\infty}}\big{\|}_{L^{\infty}(\Omega)} there exist σ,r0>0\sigma,r_{0}>0 such that, for r(0,r0]r\in(0,r_{0}] we have

𝒥(G,Br)𝒥(G+Ψ,Br)+σ2dsi=1mλis(Ω)ΨL1(r),\mathcal{J}(G_{\infty},B_{r})\leq\mathcal{J}(G_{\infty}+\Psi,B_{r})+\sigma\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\big{\|}{\Psi}\big{\|}_{L^{1}(\mathcal{B}_{r})},

for every ΨH01,a(Br;m)\Psi\in H^{1,a}_{0}(B_{r};\mathbb{R}^{m}) such that ΨL(r)δ\|{\Psi}\|_{L^{\infty}(\mathcal{B}_{r})}\leq\delta. For the sake of simplicity, we denote

C~=2dsi=1mλis(Ω).\tilde{C}=\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega).

Since we already know by Proposition 3.1 that there exists a local minimizer Hölder continuous of class C0,sC^{0,s}, we can assume that Ψ\Psi is continuous. Therefore, for every k>0k>0 let us consider the competitor

Gk,ε=i=1m(gki+ψiεη)+ei(gki+ψi+εη)ei,G_{k,\varepsilon}=\sum_{i=1}^{m}(g^{i}_{k}+\psi^{i}-\varepsilon\eta)_{+}e^{i}-(g^{i}_{k}+\psi^{i}+\varepsilon\eta)_{-}e^{i},

with ηCc(B(1+r)/2)\eta\in C^{\infty}_{c}(B_{(1+r)/2}) such that 0η10\leq\eta\leq 1 and η1\eta\equiv 1 on a neighborhood of Br¯\overline{B_{r}}. Hence, by the almost minimality of GkG_{k} in B(r0+r)/2B_{(r_{0}+r)/2}, namely we get for r(0,r0]r\in(0,r_{0}]

𝒥(Gk,B(r0+r)/2)𝒥(Gk,ε,B(r0+r)/2)+σC~Gk,εGkL1((r0+r)/2),\mathcal{J}(G_{k},B_{(r_{0}+r)/2})\leq\mathcal{J}(G_{k,\varepsilon},B_{(r_{0}+r)/2})+\sigma\tilde{C}\big{\|}{G_{k,\varepsilon}-G_{k}}\big{\|}_{L^{1}(\mathcal{B}_{(r_{0}+r)/2})},

we have

n((r0+r)/2{|Gk|>0})\displaystyle\mathcal{L}_{n}(\mathcal{B}_{(r_{0}+r)/2}\cap\{|G_{k}|>0\})\leq i=1mB(r0+r)/2|y|a|ψi|2+2|y|aψi,gkidX+\displaystyle\,\sum_{i=1}^{m}\int_{B_{(r_{0}+r)/2}}{|y|^{a}\big{\lvert}{\nabla\psi^{i}}\big{\rvert}^{2}+2|y|^{a}\langle\nabla\psi^{i},\nabla g^{i}_{k}\rangle\mathrm{d}X}+
+εi=1msuppηBrε|y|a|η|2+2|y|aη,(gki+ψi)dX+\displaystyle\,+\varepsilon\sum_{i=1}^{m}\int_{\mathrm{supp}\eta\setminus B_{r}}{\varepsilon|y|^{a}\big{\lvert}{\nabla\eta}\big{\rvert}^{2}+2|y|^{a}\langle\nabla\eta,\nabla(g^{i}_{k}+\psi^{i})\rangle\mathrm{d}X}+
+n((r0+r)/2{|Gk,ε|>0})+σC~Gk,εGkL1((r0+r)/2).\displaystyle\,+\mathcal{L}_{n}(\mathcal{B}_{(r_{0}+r)/2}\cap\{|G_{k,\varepsilon}|>0\})+\sigma\tilde{C}\big{\|}{G_{k,\varepsilon}-G_{k}}\big{\|}_{L^{1}(\mathcal{B}_{(r_{0}+r)/2})}.

In particular, localizing the measure of the positivity set in r\mathcal{B}_{r}, we get

n(r{|Gk|>0})\displaystyle\mathcal{L}_{n}(\mathcal{B}_{r}\cap\{|G_{k}|>0\})\leq i=1mB(r0+r)/2|y|a|ψi|2+2|y|aψi,gkidX+Cε+σC~ΨL1(r)\displaystyle\,\sum_{i=1}^{m}\int_{B_{(r_{0}+r)/2}}{|y|^{a}\big{\lvert}{\nabla\psi^{i}}\big{\rvert}^{2}+2|y|^{a}\langle\nabla\psi^{i},\nabla g^{i}_{k}\rangle\mathrm{d}X}+C\varepsilon+\sigma\tilde{C}\big{\|}{\Psi}\big{\|}_{L^{1}(\mathcal{B}_{r})}
+(r0+r)/2χ{|Gk,ε|>0}dx(r0+r)/2r¯χ{|Gk|>0}dx,\displaystyle\,+\int_{\mathcal{B}_{(r_{0}+r)/2}}{\chi_{\{|G_{k,\varepsilon}|>0\}}\mathrm{d}x}-\int_{\mathcal{B}_{(r_{0}+r)/2}\setminus\overline{\mathcal{B}_{r}}}{\chi_{\{|G_{k}|>0\}}\mathrm{d}x},

where we used that (Gk)k(G_{k})_{k} is uniformly bounded in H1,a(B(r0+r)/2)H^{1,a}(B_{(r_{0}+r)/2}). Since

{gkiεη>0}Br¯\displaystyle\{g^{i}_{k}-\varepsilon\eta>0\}\setminus\overline{B_{r}} {gki>0}Br¯\displaystyle\subseteq\{g^{i}_{k}>0\}\setminus\overline{B_{r}}
{gki+εη<0}Br¯\displaystyle\{g^{i}_{k}+\varepsilon\eta<0\}\setminus\overline{B_{r}} {gki<0}Br¯\displaystyle\subseteq\{g^{i}_{k}<0\}\setminus\overline{B_{r}}

and by the uniform convergence

{gki+ψiε>0}Br¯\displaystyle\{g^{i}_{k}+\psi^{i}-\varepsilon>0\}\cap\overline{B_{r}} {gi+ψi>0}Br¯\displaystyle\subseteq\{g^{i}_{\infty}+\psi^{i}>0\}\cap\overline{B_{r}}
{gki+ψi+ε<0}Br¯\displaystyle\{g^{i}_{k}+\psi^{i}+\varepsilon<0\}\cap\overline{B_{r}} {gi+ψi<0}Br¯,\displaystyle\subseteq\{g^{i}_{\infty}+\psi^{i}<0\}\cap\overline{B_{r}},

we deduce

n(r{|Gk|>0})\displaystyle\mathcal{L}_{n}(\mathcal{B}_{r}\cap\{|G_{k}|>0\})\leq B(r0+r)/2|y|a(|Ψ|2+2Ψ,Gk)dX\displaystyle\,\int_{B_{(r_{0}+r)/2}}{|y|^{a}\left(\big{\lvert}{\nabla\Psi}\big{\rvert}^{2}+2\langle\nabla\Psi,\nabla G_{k}\rangle\right)\mathrm{d}X}
+n(r{|G+Ψ|>0})+σC~ΨL1(r)+Cε.\displaystyle\,+\mathcal{L}_{n}(\mathcal{B}_{r}\cap\{|G_{{\infty}}+\Psi|>0\})+\sigma\tilde{C}\big{\|}{\Psi}\big{\|}_{L^{1}(\mathcal{B}_{r})}+C\varepsilon.

Now, using that GkGG_{k}\rightharpoonup G_{\infty} weakly in Hloc1(B1)H^{1}_{{\tiny{\mbox{loc}}}}(B_{1}) and uniformly on Br¯\overline{B_{r}}, we obtain

𝒥(G,Br)Br|y|a|(G+Ψ)|2dX+n(r{|G+Ψ|>0})+σC~ΨL1(r)+Cε\mathcal{J}(G_{\infty},B_{r})\leq\int_{B_{r}}{|y|^{a}\big{\lvert}{\nabla(G_{\infty}+\Psi)}\big{\rvert}^{2}\mathrm{d}X}+\mathcal{L}_{n}(\mathcal{B}_{r}\cap\{|G_{\infty}+\Psi|>0\})+\sigma\tilde{C}\big{\|}{\Psi}\big{\|}_{L^{1}(\mathcal{B}_{r})}+C\varepsilon

for every ε>0\varepsilon>0, which implies the desired inequality. ∎

The following density estimates for the positivity set of |G|\big{\lvert}{G}\big{\rvert} are obtained by a straightforward combination of the non-degeneracy condition (3.12) and the optimal regularity of local minimizer.

Corollary 3.8.

Let ΩD\Omega\subset D be a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega. Let r0>0r_{0}>0 be as in Proposition 3.1 and X0F(G)X_{0}\in F(G).
Then, for every r(0,r0/2)r\in(0,r_{0}/2)

(3.17) ε0ωnrnn(r(X0){|G|>0})(1ε0)ωnrn,\varepsilon_{0}\omega_{n}r^{n}\leq\mathcal{L}_{n}(\mathcal{B}_{r}(X_{0})\cap\{|G|>0\})\leq(1-\varepsilon_{0})\omega_{n}r^{n},

for some ε0>0\varepsilon_{0}>0.

Proof.

For the sake of simplicity, assume that X0=0F(G)X_{0}=0\in F(G). The proof of the left hand side is a combination of Proposition 3.1 and Proposition 3.5. More precisely, on one hand for rr small enough there exists Xrr+(G)X_{r}\in\mathcal{B}^{+}_{r}(G) such that |G|(Xr)Crs\big{\lvert}{G}\big{\rvert}(X_{r})\geq Cr^{s}. On the other one, since |G|\big{\lvert}{G}\big{\rvert} is of class C0,sC^{0,s}, by setting

C0=min{1,C[|G|]C0,s},C_{0}=\min\left\{1,\frac{C}{[|G|]_{C^{0,s}}}\right\},

we have that |G|>0\big{\lvert}{G}\big{\rvert}>0 in C0r+(Xr)\mathcal{B}_{C_{0}r}^{+}(X_{r}), which proves the claimed lower bound.
On the other hand, since |G|\big{\lvert}{G}\big{\rvert} is non-negative, up to rescaling, the upper bound in (3.17) is equivalent to

n(1{|G|=0})ε0.\mathcal{L}_{n}(\mathcal{B}_{1}\cap\{\big{\lvert}{G}\big{\rvert}=0\})\geq\varepsilon_{0}.

Thus, suppose by contradiction that there exists a sequence (Gk)k(G_{k})_{k} of eigenfunctions associated to the problem (1.1) such that 0F(Gk)0\in F(G_{k}) and

limkn(1{|Gk|=0})=0.\lim_{k\to\infty}\mathcal{L}_{n}(\mathcal{B}_{1}\cap\{\big{\lvert}{G_{k}}\big{\rvert}=0\})=0.

By Proposition 3.1 and Lemma 3.7, we already know that GkGG_{k}\to G_{\infty} weakly in H1,a(B1/2)H^{1,a}(B_{1/2}) and uniformly on every compact set of B1/2B_{1/2}. Moreover, GHloc1,a(B1/2+)Cloc0,s(B1/2¯)G_{\infty}\in H^{1,a}_{{{\tiny{\mbox{loc}}}}}(B_{1/2}^{+})\cap C^{0,s}_{{\tiny{\mbox{loc}}}}(\overline{B_{1/2}}) is a local minimizer in B1/2B_{1/2}. Now, fixed r(0,1/2)r\in(0,1/2) and X0B1/2X_{0}\in B_{1/2}, consider g~ki:Br(X0)\tilde{g}^{i}_{k}\colon B_{r}(X_{0})\to\mathbb{R} be the LaL_{a}-harmonic replacement of gkig^{i}_{k} in Br(X0)B_{r}(X_{0}), that is be such that

{Lag~ki=0in Br(X0)g~ki=gkion Br(X0).\begin{cases}L_{a}\tilde{g}^{i}_{k}=0&\mbox{in }B_{r}(X_{0})\\ \tilde{g}^{i}_{k}=g^{i}_{k}&\mbox{on }\partial B_{r}(X_{0}).\end{cases}

By the almost-minimality condition (3.8) for GkG_{k}, given the competitor G~k=(gk1,,g~ki,,gkm)\widetilde{G}_{k}=(g^{1}_{k},\dots,\tilde{g}^{i}_{k},\dots,g^{m}_{k}), we deduce

(3.18) Br(X0)|y|a|(gkig~ki)|2dXn(r(X0){|Gk|=0})+σC[|Gk|]C0,s(Br0/2)rn+s.\int_{B_{r}(X_{0})}{|y|^{a}|\nabla(g^{i}_{k}-\tilde{g}^{i}_{k})|^{2}\mathrm{d}X}\leq\mathcal{L}_{n}(\mathcal{B}_{r}(X_{0})\cap\{\big{\lvert}{G_{k}}\big{\rvert}=0\})+\sigma C[|G_{k}|]_{C^{0,s}(B_{r_{0}/2})}r^{n+s}.

Since the Hölder seminorm is uniformly bounded, up to a subsequence, the sequence (G~k)k(\widetilde{G}_{k})_{k} do converge uniformly on every compact set of Br(X0)B_{r}(X_{0}) to some function G~Hloc1(Br+(X0))\widetilde{G}_{\infty}\in H^{1}_{{{\tiny{\mbox{loc}}}}}(B_{r}^{+}(X_{0})). Thus, by applying Fatou’s Lemma to (3.18), we get

Br(X0)|y|a|(gig~i)|2dXσCrn+s,for r(0,1/2).\int_{B_{r}(X_{0})}{|y|^{a}|\nabla(g^{i}_{\infty}-\tilde{g}^{i}_{\infty})|^{2}\mathrm{d}X}\leq\sigma Cr^{n+s},\quad\mbox{for }r\in(0,1/2).

for some C>0C>0. Finally, by [11, Theorem 2.6], for r<ρ<1/2r<\rho<1/2 we get

Br(X0)|y|a|gi|2dXCρn+s+C(rρ)n+22sBρ(X0)|y|a|gi|2dX.\int_{B_{r}(X_{0})}{|y|^{a}|\nabla g^{i}_{\infty}|^{2}\mathrm{d}X}\leq C\rho^{n+s}+C\left(\frac{r}{\rho}\right)^{n+2-2s}\int_{B_{\rho}(X_{0})}{|y|^{a}|\nabla g^{i}_{\infty}|^{2}\mathrm{d}X}.

Hence, fixed δ<1/2\delta<1/2 such that q=Cδ<1q=C\delta<1, if ρ=δk1,r=δk\rho=\delta^{k-1},r=\delta^{k} and μ=δn+s\mu=\delta^{n+s} we get

Bδk(X0)|y|a|gi|2dXCμk1+Cμδ2sBδk1(X0)|y|a|gi|2dX\int_{B_{\delta^{k}}(X_{0})}|y|^{a}|\nabla g^{i}_{\infty}|^{2}\mathrm{d}X\leq C\mu^{k-1}+C\mu\delta^{2-s}\int_{B_{\delta^{k-1}}(X_{0})}|y|^{a}|\nabla g^{i}_{\infty}|^{2}\mathrm{d}X

and iterating the previous estimate, with δ>0\delta>0 such that q=Cδ2s<1q=C\delta^{2-s}<1, we get

Bδk(X0)|y|a|gi|2dXCμk1i=0k1qiCμk111q.\int_{B_{\delta^{k}}(X_{0})}|y|^{a}|\nabla g^{i}_{\infty}|^{2}\mathrm{d}X\leq C\mu^{k-1}\sum_{i=0}^{k-1}q^{i}\leq C\mu^{k-1}\frac{1}{1-q}.

Hence, there exists a universal constant C~>0\tilde{C}>0 such that

Br(X0)|y|a|gi|2dXC~rn+s,\int_{B_{r}(X_{0})}{|y|^{a}|\nabla g^{i}_{\infty}|^{2}\mathrm{d}X}\leq\tilde{C}r^{n+s},

for every r(0,1/2)r\in(0,1/2). By a covering argument and Morrey’s embeddings, the function gig^{i}_{\infty} is Hölder continuous of order C0,s+εC^{0,s+\varepsilon} with ε=min(s/2,(1s)/2)\varepsilon=\min(s/2,(1-s)/2), in contradiction with Proposition 3.5. ∎

Remark 3.9.

In the local case [31], the authors highlight how the sign of the first eigenfunction plays a major role in the proof of the upper bound on the density. Indeed, in the local setting is fundamental to know that the first eigenfunction of the shape optimizer Ω\Omega is non-degenerate near Ω\partial\Omega (see [31, Lemma 2.10]) and that Ω\Omega is connected (see [31, Corollary 4.3]).
Instead, in our case the validity of the upper bound only relies on the different local regularity of fractional eigenfunctions near their zero set depending on whether or not they change sign.

3.3. Unique continuation of eigenfunctions

In this last part of the Section we collect few observation related to some specific feature due to the nonlocal attitude of the problem (2.1). More precisely, since we are not able to prove that shape optimizer of (1.1) are connected set, we show some nonlocal effect for eigenfunctions in disconnected sets.

Remark 3.5.

In the local case [31, Corollary 4.3], the authors proved that every shape optimizer is a connected set. Their proof strictly relies on the invariance of the spectrum of the union of disjoint sets under the translations of each component and on the decomposition of the whole spectrum as the union of the spectrum of each connected components.
In the nonlocal case, we can not use a similar strategy since the mutual position of the connected components impacts the value of each eigenvalues (see [5] for some consequences of this feature in the case of the second eigenvalue). Obviously, the decomposition of the spectrum in each component is false in the nonlocal case.

It is known by [5, Theorem 2.8.] that the first normalized eigenfunction on an open bounded set Ω\Omega, even disconnected, is strictly positive (or negative) in the whole domain.
On the other hand, if we restrict our attention to shape optimizers of (1.1) we can prove that the support of the first mm normalized eigenfunction coincides with Ω\Omega, up to a (n1)(n-1)-dimensional set.

Proposition 3.10.

Let ΩD\Omega\subset D be a shape optimizer to (1.1), even disconnected, and viH0s(Ω)v^{i}\in H^{s}_{0}(\Omega) be the ii-th normalized eigenfunction on Ω\Omega. Then

if vi0 on some ball Ω then vi0 on Ω.\text{if $v^{i}\equiv 0$ on some ball $\mathcal{B}\subset\Omega$ then $v^{i}\equiv 0$ on $\Omega$}.

Consequently, n(Ω)=n({vi0})\mathcal{L}_{n}(\Omega)=\mathcal{L}_{n}(\{v^{i}\neq 0\}).

Proof.

By the Caffarelli-Silvestre extension, the previous statement is equivalent to

if gi0 on some ball Ω then gi0 on Ω,\text{if $g^{i}\equiv 0$ on some ball $\mathcal{B}\subset\Omega$ then $g^{i}\equiv 0$ on $\Omega$},

where gig^{i} solves (2.13) with λ=λis(Ω)\lambda=\lambda_{i}^{s}(\Omega). Thus, suppose that gi0g^{i}\equiv 0 on r(X0)Ω\mathcal{B}_{r}(X_{0})\subset\Omega for some X0Ω,r>0X_{0}\in\Omega,r>0, then by localizing the system (2.13) in Br(X0)B_{r}(X_{0}) we get

{Lagi=0in Br(X0){y>0}yag=g=0on r(X0).\begin{cases}L_{a}g^{i}=0&\mbox{in }B_{r}(X_{0})\cap\{y>0\}\\ -\partial^{a}_{y}g=g=0&\mbox{on }\mathcal{B}_{r}(X_{0}).\end{cases}

Thus, by the unique continuation principle for LaL_{a}-harmonic functions (see [35] and reference therein) it follows that gi0g^{i}\equiv 0 on Br(X0)¯{y0}\overline{B_{r}(X_{0})}\cap\{y\geq 0\}.
Finally, by exploiting the validity of the unique continuation principle for elliptic divergence form operator, we get that gi0g^{i}\equiv 0 in n+1{y>ε}\mathbb{R}^{n+1}\cap\{y>\varepsilon\}, for every ε>0\varepsilon>0. Since every eigenfunctions is Hölder continuous, by letting ε\varepsilon goes to zero we get that the trace g(,0)g(\cdot,0) is identically zero, in contradiction with the normalization of the eigenfunction.
Finally, since {gi=0}Ω\{g^{i}=0\}\cap\Omega has empty interior, we get the second part of the statement. ∎

Lastly, the proof of Theorem 1.1 follows by combining Proposition 2.1, Proposition 3.1, Remark 3.3 and Proposition 3.10.

4. Weiss monotonicity formula

In this Section we establish a Weiss type monotonicity formula in the spirit of [19, 31]. As it is well known in the literature, this result implies the characterization of the possible blow-up limits at free boundary points.

For a vector-valued function GH1,a(B1;m)G\in H^{1,a}(B_{1};\mathbb{R}^{m}), let us consider

(4.1) W(X0,G,r)=1rn𝒥(G,Br(X0))srn+1Br(X0)|y|a|G|2dσW(X_{0},G,r)=\frac{1}{r^{n}}\mathcal{J}(G,B_{r}(X_{0}))-\frac{s}{r^{n+1}}\int_{\partial B_{r}(X_{0})}{|y|^{a}\big{\lvert}{G}\big{\rvert}^{2}\mathrm{d}\sigma}

The monotonicity result is an essential tool for proving the regularity of the free boundary and to estimate the dimension of its singular part.

Theorem 4.1.

Let σ,r0>0\sigma,r_{0}>0 and suppose that GH1,a(Br0(X0);m)C0,s(Br0(X0);m)G\in H^{1,a}(B_{r_{0}}(X_{0});\mathbb{R}^{m})\cap C^{0,s}(B_{r_{0}}(X_{0});\mathbb{R}^{m}) satisfies the localized almost-minimality condition (3.8) for every r(0,r0]r\in(0,r_{0}]. Then, if X0F(G)X_{0}\in F(G), we get

(4.2) ddrW(X0,G,r)1rn+2i=1m+Br+(X0)|y|a(gi,XX0sgi)2dσ2σC~[|G|]C0,s(Br0)rs1,\frac{d}{dr}W(X_{0},G,r)\geq\frac{1}{r^{n+2}}\sum_{i=1}^{m}\int_{\partial^{+}B^{+}_{r}(X_{0})}{|y|^{a}\left(\langle\nabla g^{i},X-X_{0}\rangle-sg^{i}\right)^{2}\mathrm{d}\sigma}-2\sigma\tilde{C}[|G|]_{C^{0,s}(B_{r_{0}})}r^{s-1},

with C~=2ωndsi=1mλis(Ω)\tilde{C}=\frac{2\omega_{n}}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega). Thus, there exists finite the limit W(X0,G,0+)=limr0+W(X0,G,r)W(X_{0},G,0^{+})=\lim_{r\to 0^{+}}W(X_{0},G,r) and for σ=0\sigma=0 the function W(X0,G,)W(X_{0},G,\cdot) is constant in (0,+)(0,+\infty) if and only if GG is ss-homogeneous with respect to X0X_{0}.

In order to simplify the notations, since the problem is invariant under translation on {y=0}\{y=0\}, in the following computations we will assume X0=0X_{0}=0 and denote W(r)=W(0,G,r)W(r)=W(0,G,r).

Lemma 4.2.

Let σ,r0>0\sigma,r_{0}>0 and suppose that GH1,a(Br0;m)C0,s(Br0;m)G\in H^{1,a}(B_{r_{0}};\mathbb{R}^{m})\cap C^{0,s}(B_{r_{0}};\mathbb{R}^{m}) satisfies the localized almost-minimality condition (3.8) for every r(0,r0]r\in(0,r_{0}].
Then, if 0F(G)0\in F(G), we get

Br+|y|a|G|2dX+n(Br+(G)n)\displaystyle\int_{B^{+}_{r}}{|y|^{a}\big{\lvert}{\nabla G}\big{\rvert}^{2}\mathrm{d}X}+\mathcal{L}_{n}(B^{+}_{r}(G)\cap\mathbb{R}^{n})\leq ran+Br+|y|a(r2(1s)|SnG|2+s2|G|2r2s)dσ+\displaystyle\,\,\frac{r^{-a}}{n}\int_{\partial^{+}B^{+}_{r}}{|y|^{a}\left(r^{2(1-s)}\big{\lvert}{\nabla_{S^{n}}G}\big{\rvert}^{2}+s^{2}\frac{\big{\lvert}{G}\big{\rvert}^{2}}{r^{2s}}\right)\mathrm{d}\sigma}+
+rnn1(Br+(G)n)+2σC~[|G|]C0,s(Br0)rn+s,\displaystyle\,+\frac{r}{n}\mathcal{H}^{n-1}(\partial B^{+}_{r}(G)\cap\mathbb{R}^{n})+2\sigma\tilde{C}[|G|]_{C^{0,s}(B_{r_{0}})}r^{n+s},

where Br+(G)=Br+{|G|>0}\partial B^{+}_{r}(G)=\partial B^{+}_{r}\cap\{\big{\lvert}{G}\big{\rvert}>0\} and C~=2ωndsi=1mλis(Ω)\tilde{C}=\frac{2\omega_{n}}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega).

Proof.

Let us consider now the ss-homogeneous extension G~=(g~1,,g~m)\widetilde{G}=(\tilde{g}^{1},\dots,\tilde{g}^{m}) of the trace of GG on Br\partial B_{r}, defined by

G~(X)=|X|srsG(Xr|X|).\widetilde{G}(X)=\frac{\big{\lvert}{X}\big{\rvert}^{s}}{r^{s}}G\left(X\frac{r}{\big{\lvert}{X}\big{\rvert}}\right).

Then, for every i=1,,mi=1,\dots,m we get

|g~i|2(X)=s21r2s|X|2(1s)gi(Xr|X|)2+r2(1s)|X|2(1s)|Sngi|2(Xr|X|).\big{\lvert}{\nabla\tilde{g}^{i}}\big{\rvert}^{2}(X)=s^{2}\frac{1}{r^{2s}\big{\lvert}{X}\big{\rvert}^{2(1-s)}}g^{i}\left(X\frac{r}{\big{\lvert}{X}\big{\rvert}}\right)^{2}+\frac{r^{2(1-s)}}{\big{\lvert}{X}\big{\rvert}^{2(1-s)}}\big{\lvert}{\nabla_{S^{n}}g^{i}}\big{\rvert}^{2}\left(X\frac{r}{\big{\lvert}{X}\big{\rvert}}\right).

Integrating over Br+B_{r}^{+} and summing for i=1,,mi=1,\dots,m, we obtain

Br+|y|a|G~|2dX\displaystyle\int_{B^{+}_{r}}{|y|^{a}|\nabla\widetilde{G}|^{2}\mathrm{d}X} =0r1ρ2(1s)+Bρ+|y|a(s21r2s|G|2(Xrρ)+r2(1s)|SnG|2(Xrρ))dσdρ\displaystyle=\int_{0}^{r}\frac{1}{\rho^{2(1-s)}}\int_{\partial^{+}B^{+}_{\rho}}{|y|^{a}\left(s^{2}\frac{1}{r^{2s}}\big{\lvert}{G}\big{\rvert}^{2}\left(X\frac{r}{\rho}\right)+r^{2(1-s)}\big{\lvert}{\nabla_{S^{n}}G}\big{\rvert}^{2}\left(X\frac{r}{\rho}\right)\right)\mathrm{d}\sigma}\mathrm{d}\rho
=ran+Br+|y|a(s2|G|2r2s+r2(1s)|SnG|2)dσ,\displaystyle=\frac{r^{-a}}{n}\int_{\partial^{+}B^{+}_{r}}{|y|^{a}\left(s^{2}\frac{\big{\lvert}{G}\big{\rvert}^{2}}{r^{2s}}+r^{2(1-s)}\big{\lvert}{\nabla_{S^{n}}G}\big{\rvert}^{2}\right)\mathrm{d}\sigma},

while for the measure term we have that

n(Br+(G)n)=rnn1(Br+(G)n)\mathcal{L}_{n}(B^{+}_{r}(G)\cap\mathbb{R}^{n})=\frac{r}{n}\mathcal{H}^{n-1}(\partial B^{+}_{r}(G)\cap\mathbb{R}^{n})

Finally, since G~=G\widetilde{G}=G on Br\partial B_{r}, G~L(r)=GL(r)\|{\tilde{G}}\|_{L^{\infty}(\mathcal{B}_{r})}=\|{G}\|_{L^{\infty}(\mathcal{B}_{r})} and

G~GL1(r)2ωn[|G|]C0,s(Br0)rn+s.\|{\tilde{G}-G}\|_{L^{1}(\mathcal{B}_{r})}\leq 2\omega_{n}[|G|]_{C^{0,s}(B_{r_{0}})}r^{n+s}.

the almost-minimality condition (3.8) gives the claimed inequality. ∎

Proof of Theorem 4.1.

By the estimate of Lemma 4.2, we immediately get

W(r)=\displaystyle W^{\prime}(r)= 1rn(+Br+|y|a|G|2dX+n1(Br+(G)n))+\displaystyle\,\frac{1}{r^{n}}\left(\int_{\partial^{+}B^{+}_{r}}{|y|^{a}\big{\lvert}{\nabla G}\big{\rvert}^{2}\mathrm{d}X}+\mathcal{H}^{n-1}(\partial B^{+}_{r}(G)\cap\mathbb{R}^{n})\right)+
nrn+1(Br+|y|a|G|2dX+n(Br+(G)n))+\displaystyle\,-\frac{n}{r^{n+1}}\left(\int_{B^{+}_{r}}{|y|^{a}\big{\lvert}{\nabla G}\big{\rvert}^{2}\mathrm{d}X}+\mathcal{L}_{n}(B^{+}_{r}(G)\cap\mathbb{R}^{n})\right)+
2srn+1i=1m+Br+|y|agirgidσ+2s2rn+2+Br+|y|a|G|2dσ\displaystyle\,-\frac{2s}{r^{n+1}}\sum_{i=1}^{m}\int_{\partial^{+}B^{+}_{r}}{|y|^{a}g^{i}\partial_{r}g^{i}\mathrm{d}\sigma}+2\frac{s^{2}}{r^{n+2}}\int_{\partial^{+}B^{+}_{r}}{|y|^{a}\big{\lvert}{G}\big{\rvert}^{2}\mathrm{d}\sigma}
\displaystyle\geq 1rni=1m+Br+|y|a(|rgi|22sgirui+s2r2|gi|2)dσ2σC~[|G|]C0,s(Br0)rs1\displaystyle\,\frac{1}{r^{n}}\sum_{i=1}^{m}\int_{\partial^{+}B^{+}_{r}}{|y|^{a}\left(\big{\lvert}{\partial_{r}g^{i}}\big{\rvert}^{2}-2sg^{i}\partial_{r}u^{i}+\frac{s^{2}}{r^{2}}\big{\lvert}{g^{i}}\big{\rvert}^{2}\right)\mathrm{d}\sigma}-2\sigma\tilde{C}[|G|]_{C^{0,s}(B_{r_{0}})}r^{s-1}
=\displaystyle= 1rni=1m+Br+|y|a(rgisrgi)2dσ2σC~[|G|]C0,s(Br0)rs1.\displaystyle\,\frac{1}{r^{n}}\sum_{i=1}^{m}\int_{\partial^{+}B^{+}_{r}}{|y|^{a}\left(\partial_{r}g^{i}-\frac{s}{r}g^{i}\right)^{2}\mathrm{d}\sigma}-2\sigma\tilde{C}[|G|]_{C^{0,s}(B_{r_{0}})}r^{s-1}.

with C~=2ωndsi=1mλis(Ω)\tilde{C}=\frac{2\omega_{n}}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega). Hence, we deduce that

(4.3) rW(r)+2σC~s[|G|]C0,s(Br0)rsr\mapsto W(r)+2\sigma\frac{\tilde{C}}{s}[|G|]_{C^{0,s}(B_{r_{0}})}r^{s}

is monotone non-decreasing for r(0,1)r\in(0,1) and consequently there exists the limit W(0+)=limr0+W(0,G,r)W(0^{+})=\lim_{r\to 0^{+}}W(0,G,r). Finally, by exploiting the C0,sC^{0,s} regularity of GG we can easily exclude the possibility that W(0+)=W(0^{+})=-\infty: indeed, since 0F(G)0\in F(G) we get

W(r)srn+1Br|y|a|G|2dσs[|G|]C0,s2B1|y|adσ,W(r)\geq-\frac{s}{r^{n+1}}\int_{\partial B_{r}}{|y|^{a}\big{\lvert}{G}\big{\rvert}^{2}\mathrm{d}\sigma}\geq-s[|G|]^{2}_{C^{0,s}}\int_{\partial B_{1}}|y|^{a}\mathrm{d}\sigma,

for every r>0r>0. Finally, if σ=0\sigma=0 the right hand side of (4.2) is non-negative, and so we deduce that W(r)0W^{\prime}(r)\equiv 0 for r(0,+)r\in(0,+\infty) if and only if

gi(X),X|X|=s|X|gi(X) in n+1,\left\langle\nabla g^{i}(X),\frac{X}{\big{\lvert}{X}\big{\rvert}}\right\rangle=\frac{s}{\big{\lvert}{X}\big{\rvert}}g^{i}(X)\quad\text{ in $\mathbb{R}^{n+1}$,}

namely the components gig^{i} are ss-homogeneous in n+1\mathbb{R}^{n+1}. ∎

5. Compactness and convergence of blow-up sequences

This Section is dedicated to the convergence of the blow-up sequences and the analysis of the blow-up limits, both being essential for determining the local behavior of the free boundary and for the characterization of the Regular and Singular strata.

Let us recall the notion of blow-up sequence associated to a local minimizer GG in B1B_{1}. Given (Xk)kF(G)(X_{k})_{k}\subset F(G) and rk0+r_{k}\searrow 0^{+} such that Brk(Xk)B1B_{r_{k}}(X_{k})\subset B_{1}, we define a blow-up sequence by

(5.1) GXk,rk(X)=1rksG(Xk+rkX).G_{X_{k},r_{k}}(X)=\frac{1}{r^{s}_{k}}G(X_{k}+r_{k}X).

Thus, the sequence (GXk,rk)k(G_{X_{k},r_{k}})_{k} is uniformly Hölder continuous in the class C0,sC^{0,s} and locally uniformly bounded in n+1\mathbb{R}^{n+1}. Thus, by Lemma 3.7 we already know that, up to a subsequence, (GXk,rk)k(G_{X_{k},r_{k}})_{k} converges locally uniformly on every compact set to a function G0Hloc1,a(n+1;m)Cloc0,s(n+1;m)G_{0}\in H^{1,a}_{{{\tiny{\mbox{loc}}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m})\cap C^{0,s}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that, for every R>0R>0 the following properties hold

  • GXk,rkG0G_{X_{k},r_{k}}\to G_{0} in Cloc0,α(BR¯;m)C^{0,\alpha}_{{\tiny{\mbox{loc}}}}(\overline{B_{R}};\mathbb{R}^{m}), for every α(0,s)\alpha\in(0,s);

  • GXk,rkG0G_{X_{k},r_{k}}\rightharpoonup G_{0} weakly in H1,a(BR;m)H^{1,a}(B_{R};\mathbb{R}^{m});

  • G0G_{0} is an almost-minimizer in BRB_{R} in the sense of definition (3.7) (and also (3.8)).

In order to improve the previous compactness result, we need to show that the blow-up sequence satisfies a scaled almost-minimality condition. As we show in the following Lemma, this is a direct consequence from the scaling properties of the functional.

Lemma 5.1.

Suppose that GH1,a(n+1;m)L(n+1;m)G\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m})\cap L^{\infty}(\mathbb{R}^{n+1};\mathbb{R}^{m}) satisfies the almost-minimality condition (3.7) for some constants K,ε>0K,\varepsilon>0. Then, for every X0{y=0},r>0X_{0}\in\{y=0\},r>0 the function GX0,rG_{X_{0},r} defined in (5.1) satisfies:

(5.2) 𝒥(GX0,r)𝒥(G~)+K2dsi=1mλis(Ω)G~GX0,rL1(n)rs\mathcal{J}(G_{X_{0},r})\leq\mathcal{J}(\tilde{G})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G_{X_{0},r}}\|_{L^{1}(\mathbb{R}^{n})}r^{s}

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~L(n)δrs\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta r^{-s} and GG~L1(n)εrns\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon r^{-n-s}.

Proof.

Since the problem is translation invariance in {y=0}\{y=0\}, suppose X0=0X_{0}=0 and denote Gr=G0,rG_{r}=G_{0,r}. Let G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be such that

G~L(n)δrsandGG~L1(n)εrn+s.\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\frac{\delta}{r^{s}}\quad\mbox{and}\quad\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\frac{\varepsilon}{r^{n+s}}.

Consider now the function Φ=GrG~\Phi=G_{r}-\tilde{G} and Φ1/r(X)=rsΦ(X/r),G~1/r(X)=rsG~(X/r)\Phi_{1/r}(X)=r^{s}\Phi(X/r),\tilde{G}_{1/r}(X)=r^{s}\tilde{G}(X/r). We notice that

Φ1/rL1(n)=rn+sΦL1(n)εandG~1/rL(n)=rsG~L(n)δ,\big{\|}{\Phi_{1/r}}\big{\|}_{L^{1}(\mathbb{R}^{n})}=r^{n+s}\big{\|}{\Phi}\big{\|}_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon\quad\mbox{and}\quad\|{\tilde{G}_{1/r}}\|_{L^{\infty}(\mathbb{R}^{n})}=r^{s}\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta,

which allows to consider exploit the almost-minimality of GG with respect to the competitor G~1/r=G+Φ1/r\tilde{G}_{1/r}=G+\Phi_{1/r}. Thus, we get

𝒥(Gr)\displaystyle\mathcal{J}(G_{r}) =1rn𝒥(G)\displaystyle=\frac{1}{r^{n}}\mathcal{J}(G)
1rn𝒥(G~1/r)+K2dsi=1mλis(Ω)G~1/rGL1(n)rn\displaystyle\leq\frac{1}{r^{n}}\mathcal{J}(\tilde{G}_{1/r})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\frac{\|{\tilde{G}_{1/r}-G}\|_{L^{1}(\mathbb{R}^{n})}}{r^{n}}
𝒥(G~)+K2dsi=1mλis(Ω)G~GrL1(n)rs,\displaystyle\leq\mathcal{J}(\tilde{G})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G_{r}}\|_{L^{1}(\mathbb{R}^{n})}r^{s},

as we claimed. ∎

Proposition 5.2.

Let ΩD\Omega\subset D be a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega. Given (Xk)kF(G)(X_{k})_{k}\subset F(G) and rk0+r_{k}\searrow 0^{+}, for every R>0R>0 the following properties hold (up to extracting a subsequence):

  • GXk,rkG0G_{X_{k},r_{k}}\to G_{0} strongly in H1,a(BR;m)H^{1,a}(B_{R};\mathbb{R}^{m});

  • the sequence of the characteristic functions

    χ({|GXk,rk|>0})χ({|G0|>0})\chi(\{|G_{X_{k},r_{k}}|>0\})\to\chi(\{|G_{0}|>0\})

    strongly in L1(R)L^{1}(\mathcal{B}_{R});

  • the sequence of the closed sets R+(GXk,rk)¯\overline{\mathcal{B}^{+}_{R}(G_{X_{k},r_{k}})} and its complement in n\mathbb{R}^{n}, converge in the Hausdorff sense respectively to R+(G0)¯\overline{\mathcal{B}^{+}_{R}(G_{0})} and nR+(G0)¯\mathbb{R}^{n}\setminus\overline{\mathcal{B}^{+}_{R}(G_{0})}

  • the blow-up limit G0G_{0} is non-degenerate at zero, namely there exists a dimensional constant c0>0c_{0}>0 such that

    supr|G0|c0rsfor every r>0.\sup_{\mathcal{B}_{r}}\big{\lvert}{G_{0}}\big{\rvert}\geq c_{0}r^{s}\quad\text{for every $r>0$}.
Proof.

For notational simplicity, we set Gk=GXk,rkG_{k}=G_{X_{k},r_{k}}. Since |Gk|\big{\lvert}{G_{k}}\big{\rvert} converges locally uniformly to |G0|\big{\lvert}{G_{0}}\big{\rvert}, we get

χ({|G0|>0})lim infkχ({|Gk|>0})\chi(\{|G_{0}|>0\})\leq\liminf_{k\to\infty}\chi(\{|G_{k}|>0\})

Now, let us prove that GkG_{k} converges strongly in Hloc1,a(n+1;m)H^{1,a}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m}) to G0G_{0} and that the characteristic functions χ({|Gk|>0})\chi(\{|G_{k}|>0\}) converge to χ({|Gk|>0})\chi(\{|G_{k}|>0\}) in L1L^{1}. Namely, fixed a radius R>0R>0, it is sufficient to prove that

limkBR|y|a|Gk|2dX+n(r+(Gk))=BR|y|a|G0|2dX+n(r+(G0)).\lim_{k\to\infty}\int_{B_{R}}{|y|^{a}\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}\mathrm{d}X}+\mathcal{L}_{n}(\mathcal{B}_{r}^{+}(G_{k}))=\int_{B_{R}}{|y|^{a}\big{\lvert}{\nabla G_{0}}\big{\rvert}^{2}\mathrm{d}X}+\mathcal{L}_{n}(\mathcal{B}_{r}^{+}(G_{0})).

Consider now ηCc(n+1),0η1\eta\in C^{\infty}_{c}(\mathbb{R}^{n+1}),0\leq\eta\leq 1 such that η1\eta\equiv 1 on BRB_{R}, and the competitor G~kH1,a(n+1;m)\widetilde{G}_{k}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) defined by

G~k=ηG0+(1η)Gk.\widetilde{G}_{k}=\eta G_{0}+(1-\eta)G_{k}.

For the sake of notational simplicity, let us set:

Ωk={|Gk|>0}n,Ω~k={|G~k|>0}nandΩ0={|G0|>0}n.\Omega_{k}=\{\big{\lvert}{G_{k}}\big{\rvert}>0\}\cap\mathbb{R}^{n},\quad\widetilde{\Omega}_{k}=\{|\widetilde{G}_{k}|>0\}\cap\mathbb{R}^{n}\quad\mbox{and}\quad\Omega_{0}=\{\big{\lvert}{G_{0}}\big{\rvert}>0\}\cap\mathbb{R}^{n}.

Since G~k=Gk\widetilde{G}_{k}=G_{k} on {η=0}\{\eta=0\}, by the almost-minimality condition (5.2) applied on GkG_{k}, given δGL(Ω)\delta\geq\big{\|}{G}\big{\|}_{L^{\infty}(\Omega)} and K,ε>0K,\varepsilon>0 the associated constants, we get

𝒥(Gk)𝒥(G~k)+K2dsi=1mλis(Ω)G~kGkL1(n)rks,\mathcal{J}(G_{k})\leq\mathcal{J}(\tilde{G}_{k})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}_{k}-G_{k}}\|_{L^{1}(\mathbb{R}^{n})}r^{s}_{k},

and consequently

(5.3) {η>0}|y|a|Gk|2dX+Λ~n(Ωk{η>0}){η>0}|y|a|G~k|2dX+Λ~n(Ω~k{η>0})+K2dsi=1mλis(Ω)G~kGkL1(n)rks{η>0}|G~k|2dX+Λ~n(Ω0{η=1})+Λ~n({0<η<1})++K2dsi=1mλis(Ω)η(G0Gk)L1(n)rks\displaystyle\begin{aligned} \int_{\{\eta>0\}}{|y|^{a}\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}\mathrm{d}X}&+\tilde{\Lambda}\mathcal{L}_{n}(\Omega_{k}\cap\{\eta>0\})\\ \leq&\int_{\{\eta>0\}}{|y|^{a}|\nabla\widetilde{G}_{k}|^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\widetilde{\Omega}_{k}\cap\{\eta>0\})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\big{\|}{\tilde{G}_{k}-G_{k}}\big{\|}_{L^{1}(\mathbb{R}^{n})}r^{s}_{k}\\ \leq&\int_{\{\eta>0\}}{|\nabla\widetilde{G}_{k}|^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\Omega_{0}\cap\{\eta=1\})+\tilde{\Lambda}\mathcal{L}_{n}(\{0<\eta<1\})+\\ &\,+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\big{\|}{\eta(G_{0}-G_{k})}\big{\|}_{L^{1}(\mathbb{R}^{n})}r^{s}_{k}\end{aligned}

On {η>0}\{\eta>0\} we calculate

|Gk|2|G~k|2=\displaystyle\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\nabla\widetilde{G}_{k}|^{2}= |Gk|2|ηG0+(1η)Gk+(G0Gk)η|2\displaystyle\,\,\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\eta\nabla G_{0}+(1-\eta)\nabla G_{k}+(G_{0}-G_{k})\nabla\eta|^{2}
=\displaystyle= (1(1η)2)|Gk|2η2|G0|2|G0Gk|2|η|2+\displaystyle\,\,(1-(1-\eta)^{2})\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-\eta^{2}|\nabla G_{0}|^{2}-|G_{0}-G_{k}|^{2}|\nabla\eta|^{2}+
2(G0Gk)η,ηG0+(1η)Gk2η(1η)G0,Gk.\displaystyle-2(G_{0}-G_{k})\langle\nabla\eta,\eta\nabla G_{0}+(1-\eta)\nabla G_{k}\rangle-2\eta(1-\eta)\langle\nabla G_{0},\nabla G_{k}\rangle.

Since GkG_{k} converges strongly in L2,a(BR;m)L^{2,a}(B_{R};\mathbb{R}^{m}) and weakly Hloc1,a(n+1;m)H^{1,a}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m}) to G0G_{0}, we can estimate

lim supk\displaystyle\limsup_{k\to\infty} {η>0}|y|a(|Gk|2|G~k|2)dX=\displaystyle\int_{\{\eta>0\}}{|y|^{a}\left(\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\nabla\widetilde{G}_{k}|^{2}\right)\mathrm{d}X}=
=lim supk{η>0}|y|a((1(1η)2)|Gk|2η2|G0|22η(1η)G0,Gk)dX\displaystyle=\limsup_{k\to\infty}\int_{\{\eta>0\}}{|y|^{a}\left((1-(1-\eta)^{2})\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-\eta^{2}|\nabla G_{0}|^{2}-2\eta(1-\eta)\langle\nabla G_{0},\nabla G_{k}\rangle\right)\mathrm{d}X}
=lim supk{η>0}|y|a(1(1η)2)(|Gk|2|G0|2)dX\displaystyle=\limsup_{k\to\infty}\int_{\{\eta>0\}}{|y|^{a}(1-(1-\eta)^{2})\left(\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\nabla G_{0}|^{2}\right)\mathrm{d}X}
lim supk{η=1}|y|a(|Gk|2|G0|2)dX,\displaystyle\geq\limsup_{k\to\infty}\int_{\{\eta=1\}}{|y|^{a}\left(\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\nabla G_{0}|^{2}\right)\mathrm{d}X},

where in the last inequality we used that |Gk|\big{\lvert}{\nabla G_{k}}\big{\rvert} weakly converges in L2,a({0<η<1})L^{2,a}(\{0<\eta<1\}) to |G0|\big{\lvert}{\nabla G_{0}}\big{\rvert}.

Combining this fact with inequality (5.3), we obtain

lim supk\displaystyle\limsup_{k\to\infty} ({η=1}|y|a(|Gk|2|G0|2)dX+n(Ωk{η=1})n(Ω0{η=1}))\displaystyle\left(\int_{\{\eta=1\}}{|y|^{a}\left(\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\nabla G_{0}|^{2}\right)\mathrm{d}X}+\mathcal{L}_{n}(\Omega_{k}\cap\{\eta=1\})-\mathcal{L}_{n}(\Omega_{0}\cap\{\eta=1\})\right)
lim supk({η>0}|y|a(|Gk|2|G~k|2)dX+n(Ωk{η=1})n(Ω0{η=1}))\displaystyle\leq\limsup_{k\to\infty}\left(\int_{\{\eta>0\}}{|y|^{a}\left(\big{\lvert}{\nabla G_{k}}\big{\rvert}^{2}-|\nabla\widetilde{G}_{k}|^{2}\right)\mathrm{d}X}+\mathcal{L}_{n}(\Omega_{k}\cap\{\eta=1\})-\mathcal{L}_{n}(\Omega_{0}\cap\{\eta=1\})\right)
lim supkn(Ωk{η=1})n(Ωk{η>0})+n({0<η<1})\displaystyle\leq\limsup_{k\to\infty}\mathcal{L}_{n}(\Omega_{k}\cap\{\eta=1\})-\mathcal{L}_{n}(\Omega_{k}\cap\{\eta>0\})+\mathcal{L}_{n}(\{0<\eta<1\})
n({0<η<1}).\displaystyle\leq\mathcal{L}_{n}(\{0<\eta<1\}).

Finally, since η\eta is arbitrary outside BRB_{R}, the right hand side can be made arbitrarily small, and this implies the desired equality.

By Corollary 3.8, we already know that

(5.4) ε0ωnrnn(r{|Gk|>0})(1ε0)ωnrn,for r<r0/rk,\varepsilon_{0}\omega_{n}r^{n}\leq\mathcal{L}_{n}(\mathcal{B}_{r}\cap\{|G_{k}|>0\})\leq(1-\varepsilon_{0})\omega_{n}r^{n},\quad\text{for $r<r_{0}/r_{k}$},

and for every k>0k>0. Now, it is well-known that the convergence of the sequence of characteristic functions in the strong topology of L1L^{1}, together with (5.4), implies the Hausdorff convergence of ΩkBR¯\overline{\Omega_{k}\cap B_{R}} to Ω0BR¯\overline{\Omega_{0}\cap B_{R}} locally in n\mathbb{R}^{n}. Obviously, the same result holds for the complements Ωkc\Omega_{k}^{c} in n\mathbb{R}^{n}.

Finally, the non-degeneracy of the blow-up limit is a straightforward combination of the uniform convergence and the non-degeneracy condition (3.12). Namely, by Proposition 3.5, for every k>0k>0 the rescaled function GkG_{k} is non-degenerate in the sense

for every yΩk¯,r12rksupr(y)|Gk|c0rs.\text{for every }y\in\overline{\Omega_{k}},r\leq\frac{1}{2r_{k}}\quad\sup_{\mathcal{B}_{r}(y)}\big{\lvert}{G_{k}}\big{\rvert}\geq c_{0}r^{s}.

The previous inequality is obtained by applying (3.12) in Brkr(y)B_{r_{k}r}(y) for the local minimizer GG. Finally, by the uniform convergence of GkG_{k} and the Hausdorff convergence of ΩkBr\Omega_{k}\cap B_{r} in n\mathbb{R}^{n}, for every yΩ0¯y\in\overline{\Omega_{0}} we get

supr(y)|Gk|c0rs,for every r>0.\sup_{\mathcal{B}_{r}(y)}\big{\lvert}{G_{k}}\big{\rvert}\geq c_{0}r^{s},\,\,\text{for every }r>0.

Proposition 5.3.

Let ΩD\Omega\subset D be a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega. Assume that X0F(G)X_{0}\in F(G), then every blow-up limit G0G_{0} is a global minimizer of the functional 𝒥\mathcal{J}. More precisely, for every R>0R>0 and for every G~0Hloc1,a(n+1;m)\tilde{G}_{0}\in H^{1,a}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~0G0H01,a(BR;m)\tilde{G}_{0}-G_{0}\in H^{1,a}_{0}(B_{R};\mathbb{R}^{m}) we have

BR|y|a|G0|2dX+Λ~n(R{|G0|>0})BR|y|a|G~0|2dX+Λ~n(R{|G~0|>0}).\int_{B_{R}}{|y|^{a}\big{\lvert}{\nabla G_{0}}\big{\rvert}^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}_{R}\cap\{|G_{0}|>0\})\leq\int_{B_{R}}{|y|^{a}|\nabla\tilde{G}_{0}|^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}_{R}\cap\{|\tilde{G}_{0}|>0\}).
Proof.

By Remark 2.4, it is not restrictive to assume X0=0X_{0}=0 and to consider Gr=G0,rG_{r}=G_{0,r} a blow-up sequence centered at 0F(G)0\in F(G). By Lemma 5.1, for every R>0R>0 it holds

(5.5) 𝒥(Gr)𝒥(G~)+K2dsi=1mλis(Ω)G~GrL1(n)rs\mathcal{J}(G_{r})\leq\mathcal{J}(\tilde{G})+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}-G_{r}}\|_{L^{1}(\mathbb{R}^{n})}r^{s}

for every G~H1,a(n+1;m)\tilde{G}\in H^{1,a}(\mathbb{R}^{n+1};\mathbb{R}^{m}) such that G~GrH1,a(BR;m)\tilde{G}-G_{r}\in H^{1,a}(B_{R};\mathbb{R}^{m}) and

G~L(n)δrsandGG~L1(n)εrns.\|{\tilde{G}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\delta r^{-s}\quad\mbox{and}\quad\|{G-\tilde{G}}\|_{L^{1}(\mathbb{R}^{n})}\leq\varepsilon r^{-n-s}.

Let now G~0Hloc1,a(n+1;m)Lloc(n+1;m)\tilde{G}_{0}\in H^{1,a}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m})\cap L^{\infty}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be such that G~0G0H01,a(BR;m)\tilde{G}_{0}-G_{0}\in H^{1,a}_{0}(B_{R};\mathbb{R}^{m}) and let ηCc(BR)\eta\in C^{\infty}_{c}(B_{R}) be such that 0η10\leq\eta\leq 1. Set GrkG_{r_{k}} the blow-up sequence associated to rk0+r_{k}\searrow 0^{+} converging to G0G_{0} in the sense of Lemma 3.7 and Proposition 5.2, and consider the new test function

G~rk=G~0+(1η)(GrkG0).\tilde{G}_{r_{k}}=\tilde{G}_{0}+(1-\eta)(G_{r_{k}}-G_{0}).

Since G~0=G0\tilde{G}_{0}=G_{0} outside of BRB_{R}, we get G~rk=Grk\tilde{G}_{r_{k}}=G_{r_{k}} outside of BRB_{R}. Moreover, since

G~rkGrk=G~0G0η(GrkG0).\tilde{G}_{r_{k}}-G_{r_{k}}=\tilde{G}_{0}-G_{0}-\eta(G_{r_{k}}-G_{0}).

and GrkG0G_{r_{k}}\to G_{0} in L1(R)L^{1}(\mathcal{B}_{R}), there exists k0=k0(n,s)>0k_{0}=k_{0}(n,s)>0 such that

G~rkGrkL1(n)2G~0G0L1(n)andG~rkGrkL(n)2G~0G0L(n)\|{\tilde{G}_{r_{k}}-G_{r_{k}}}\|_{L^{1}(\mathbb{R}^{n})}\leq 2\|{\tilde{G}_{0}-G_{0}}\|_{L^{1}(\mathbb{R}^{n})}\quad\mbox{and}\quad\|{\tilde{G}_{r_{k}}-G_{r_{k}}}\|_{L^{\infty}(\mathbb{R}^{n})}\leq 2\|{\tilde{G}_{0}-G_{0}}\|_{L^{\infty}(\mathbb{R}^{n})}

for kk0k\geq k_{0}. Hence, by testing (5.5) with G~rk\tilde{G}_{r_{k}} we deduce

BR|y|a|Grk|2dX+Λ~n(R+(Grk))\displaystyle\int_{B_{R}}|y|^{a}|\nabla G_{r_{k}}|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}^{+}_{R}(G_{r_{k}}))\leq BR|y|a|G~rk|2dX+Λ~n(R+(G~0){η=1})+\displaystyle\,\int_{B_{R}}|y|^{a}|\nabla\tilde{G}_{r_{k}}|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}^{+}_{R}(\tilde{G}_{0})\cap\{\eta=1\})+
+Λ~n({0<η<1}n)+\displaystyle\,+\tilde{\Lambda}\mathcal{L}_{n}(\{0<\eta<1\}\cap\mathbb{R}^{n})+
+K2dsi=1mλis(Ω)G~rkGrkL1(n)rks.\displaystyle\,+K\frac{2}{d_{s}}\sum_{i=1}^{m}\lambda_{i}^{s}(\Omega)\|{\tilde{G}_{r_{k}}-G_{r_{k}}}\|_{L^{1}(\mathbb{R}^{n})}r^{s}_{k}.

Since G~rkG~0,GrkG0\tilde{G}_{r_{k}}\to\tilde{G}_{0},G_{r_{k}}\to G_{0} strongly in H1,a(BR;m)H^{1,a}(B_{R};\mathbb{R}^{m}), we get

BR|y|a|G0|2dX+Λ~n({|G0|>0}R)\displaystyle\int_{B_{R}}|y|^{a}|\nabla G_{0}|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{\big{\lvert}{G_{0}}\big{\rvert}>0\}\cap\mathcal{B}_{R})\leq BR|y|a|G~0|2dX+Λ~n({|G~0|>0}R)+\displaystyle\,\int_{B_{R}}|y|^{a}|\nabla\tilde{G}_{0}|^{2}\mathrm{d}X+\tilde{\Lambda}\mathcal{L}_{n}(\{|\tilde{G}_{0}|>0\}\cap\mathcal{B}_{R})+
+Λ~n({0<η<1}R)\displaystyle\,+\tilde{\Lambda}\mathcal{L}_{n}(\{0<\eta<1\}\cap\mathcal{B}_{R})

Finally, by choosing η\eta such that n({η=1}R)\mathcal{L}_{n}(\{\eta=1\}\cap\mathcal{B}_{R}) arbitrarily close to n(R)\mathcal{L}_{n}(\mathcal{B}_{R}), we get the claimed inequality. ∎

The following is a straightforward application of the Weiss monotonicity formula to the previous characterization of the blow-up limits.

Corollary 5.4.

Let ΩD\Omega\subset D be a shape optimizer to (1.1) and GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on Ω\Omega. Assume that X0F(G)X_{0}\in F(G), then every blow-up limit G0=(g0i,,g0m)G_{0}=(g^{i}_{0},\dots,g^{m}_{0}) of GG at X0X_{0} is ss-homogeneous in n+1\mathbb{R}^{n+1}, that is,

g0i(X),X|X|=s|X|g0i(X) in n+1,\left\langle\nabla g^{i}_{0}(X),\frac{X}{\big{\lvert}{X}\big{\rvert}}\right\rangle=\frac{s}{\big{\lvert}{X}\big{\rvert}}g^{i}_{0}(X)\,\text{ in $\mathbb{R}^{n+1}$},

for every i=1,,mi=1,\dots,m. Moreover, the Lebesgue density of F(G)F(G) exists finite at every X0F(G)X_{0}\in F(G) and it satisfies

(5.6) {|G|>0}(γ)={X0F(G):limr0+n(r(X0){|G|>0})n(r)=γ}={X0F(G):W(X0,G,0+)=ωnγ}.\displaystyle\begin{aligned} \left\{\big{\lvert}{G}\big{\rvert}>0\right\}^{(\gamma)}&=\left\{X_{0}\in F(G)\colon\lim_{r\to 0^{+}}\frac{\mathcal{L}_{n}(\mathcal{B}_{r}(X_{0})\cap\{|G|>0\})}{\mathcal{L}_{n}(\mathcal{B}_{r})}=\gamma\right\}\\ &=\left\{X_{0}\in F(G)\colon W(X_{0},G,0^{+})=\omega_{n}\gamma\right\}.\end{aligned}
Proof.

Let X0F(G)X_{0}\in F(G) and G0G_{0} a blow-up limit of GG at X0X_{0} associated to a sequence rk0+r_{k}\searrow 0^{+}. By Lemma 3.7, Proposition 5.2 and Proposition 6.2, we already know that G0G_{0} is a global minimizer of 𝒥\mathcal{J}. On the other hand, by the definition of the Weiss formula, for every ρ,r>0\rho,r>0 we get

W(X0,G,rρ)=W(0,GX0,r,ρ).W(X_{0},G,r\rho)=W(0,G_{X_{0},r},\rho).

Fixed R>0R>0, since up to a subsequence GX0,rkG0G_{X_{0},r_{k}}\to G_{0} uniformly and strongly in H1,a(BR;m)H^{1,a}(B_{R};\mathbb{R}^{m}), we deduce

W(0,G0,R)=limkW(0,GX0,rk,R)=limkW(X0,G,Rrk)=limr0+W(X0,G,r),W(0,G_{0},R)=\lim_{k\to\infty}W(0,G_{X_{0},r_{k}},R)=\lim_{k\to\infty}W(X_{0},G,Rr_{k})=\lim_{r\to 0^{+}}W(X_{0},G,r),

where the last limit is unique and it does not depend on the sequence (rk)k(r_{k})_{k}, thanks to the monotonicity result Theorem 4.1. Finally, since G0G_{0} is a global minimizer of 𝒥\mathcal{J} and RW(0,G0,R)R\mapsto W(0,G_{0},R) is constant we get that the blow-up limit is ss-homogeneous (see case σ=0\sigma=0 of Theorem 4.1).
Moreover, the homogeneity of the blow-up limits and the strong convergence of the blow-up sequences imply

(5.7) n(1{|G0|>0})=W(0,G0,1)=limr0+W(X0,G,r)=limr0+n(r(X0){|G|>0})rn.\mathcal{L}_{n}(\mathcal{B}_{1}\cap\{|G_{0}|>0\})=W(0,G_{0},1)=\lim_{r\to 0^{+}}W(X_{0},G,r)=\lim_{r\to 0^{+}}\frac{\mathcal{L}_{n}(\mathcal{B}_{r}(X_{0})\cap\{|G|>0\})}{r^{n}}.

Hence, the density W(X0,G,0+)W(X_{0},G,0^{+}) coincides, up to a multiplicative constant, with the Lebesgue density of the free boundary. ∎

Remark 5.5.

Notice that, by (5.7), the measure of the positivity set of the blow-up limit at some X0F(G)X_{0}\in F(G) in 1\mathcal{B}_{1} does not depend on the blow-up limit itself.

Remark 5.6.

In the classification of the blow-up limits, we will use some results related to eigenvalues of the weighted Laplace-Beltrami operator related to the operator LaL_{a}. By direct computation, the operator LaL_{a} can be decomposed as

Lau=sina(θn)1rnr(rn+2aru)+1r2aLaSnuL_{a}u=\sin^{a}(\theta_{n})\frac{1}{r^{n}}\partial_{r}\left(r^{n+2-a}\partial_{r}u\right)+\frac{1}{r^{2-a}}L_{a}^{S^{n}}u

where y=rsin(θn)y=r\sin(\theta_{n}) and the Laplace-Beltrami type operator is defined as

LaSn=divSn(sina(θn)Sn),L_{a}^{S^{n}}=\mbox{div}_{S^{n}}(\sin^{a}(\theta_{n})\nabla_{S^{n}}),

with divSn\mbox{div}_{S^{n}} and Sn\nabla_{S^{n}} respectively the tangential divergence and gradient on SnS^{n}. In particular, the following results hold true.

Let ωSn1×{0}\omega\subset S^{n-1}\times\{0\} be an open subset of the (n1)(n-1)-sphere and let Σω={rθ:θω,r>0}×{0}\Sigma_{\omega}=\{r\theta\colon\theta\in\omega,r>0\}\times\{0\} be the cone generated by ω\omega in {y=0}\{y=0\}. Then, gg is a α\alpha-homogeneous solution of

{Lag=0in +n+1yag=0on Σωg=0on {y=0}Σω,\begin{cases}-L_{a}g=0&\emph{in }\mathbb{R}^{n+1}_{+}\\ \partial^{a}_{y}g=0&\emph{on }\Sigma_{\omega}\\ g=0&\emph{on }\{y=0\}\setminus\Sigma_{\omega},\end{cases}

if and only if its trace φ=g|Sn\varphi=g\lvert_{S^{n}} on the sphere satisfies

(5.8) {LaSnφ=λ(α)φin S+nθnaφ=0on ωφ=0on (Sn1×{0})ω,\begin{cases}-L_{a}^{S^{n}}\varphi=\lambda(\alpha)\varphi&\emph{in }S^{n}_{+}\\ \partial^{a}_{\theta_{n}}\varphi=0&\emph{on }\omega\\ \varphi=0&\emph{on }(S^{n-1}\times\{0\})\setminus\omega,\end{cases}

with λ(α)=α(α+n+a1)\lambda(\alpha)=\alpha(\alpha+n+a-1) the characteristic eigenvalue associated to the Section ω\omega. Moreover, both the map ωα(ω)\omega\mapsto\alpha(\omega) and ωλ(α(ω))\omega\mapsto\lambda(\alpha(\omega)) are monotone with respect to the inclusion of spherical sets.

In particular, if α<min(1,2s)\alpha<\min(1,2s), then φ\varphi cannot change sign and it is indeed a multiple of the principle eigenvalue. Finally, for every spherical set ωSn1\omega\subset S^{n-1} such that n1(S)nωn/2\mathcal{H}^{n-1}(S)\leq n\omega_{n}/2 we have the inequality

λ1s(ns),\lambda_{1}\geq s\left(n-s\right),

and the equality is achieved if and only if, up to a rotation, ω=Sn1{xn>0}\omega=S^{n-1}\cap\{x_{n}>0\}.

The proof of these claims uses the monotonicity of the eigenvalue with respect to the inclusion of spherical set and the Pólya-Szegö inequality for the Schwarz symmetrization applied to the eigenvalue problem (5.8) (see [36, 37] for further details).

The following Lemma characterizes the structure of the blow-up limits. In particular, we can prove that the norm of every blow-up limit is a global minimizer of a scalar thin one-phase problem.

Proposition 5.7.

Under the same hypotheses of Corollary 5.4, every blow-up limit G0G_{0} is of the form

G0(X)=ξ|G0|(X)whereξm,|ξ|=1G_{0}(X)=\xi\big{\lvert}{G_{0}}\big{\rvert}(X)\quad\mbox{where}\quad\xi\in\mathbb{R}^{m},\big{\lvert}{\xi}\big{\rvert}=1

and |G0|\big{\lvert}{G_{0}}\big{\rvert} is a global minimizer of the scalar thin functional

(5.9) 𝒥(g,BR)=BR|y|a|g|2dX+Λ~n(R{g>0}),for R>0.\mathcal{J}(g,B_{R})=\int_{B_{R}}{|y|^{a}\big{\lvert}{\nabla g}\big{\rvert}^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}_{R}\cap\{g>0\}),\quad\mbox{for }R>0.

Moreover, there exists a dimensional constant δ(0,1/2)\delta\in(0,1/2) such that one of the following possibilities holds:

  1. 1.

    The Lebesgue density of {|G|>0}\{|G|>0\} at X0X_{0} is 1/21/2 and every blow-up limit G0G_{0} is of the form

    (5.10) G0(X)=ΛΓ(1+s)U(x,ν,y)ξwhereξm,|ξ|=1,νSn1×{0}.G_{0}(X)=\frac{\sqrt{\Lambda}}{\Gamma(1+s)}U(\langle x,\nu\rangle,y)\xi\quad\mbox{where}\quad\xi\in\mathbb{R}^{m},\big{\lvert}{\xi}\big{\rvert}=1,\nu\in S^{n-1}\times\{0\}.
  2. 2.

    The Lebesgue density of {|G|>0}\{|G|>0\} at X0X_{0} satisfies

    (5.11) 12+δlimr0|r(X0){|G|>0})||r|1δ,\frac{1}{2}+\delta\leq\lim_{r\to 0}\frac{|\mathcal{B}_{r}(X_{0})\cap\{|G|>0\})|}{|\mathcal{B}_{r}|}\leq 1-\delta,

    and |G0||G_{0}| is a nonnegative global minimizer of (5.9) with singularity in zero.

Proof.

Let X0F(G)X_{0}\in F(G) and G0G_{0} a blow-up limit of GG at X0X_{0}. By Corollary 5.4, we already know that G0G_{0} is an ss-homogeneous global minimizer such that

|1+(G0)|=γ|1|,|\mathcal{B}^{+}_{1}(G_{0})|=\gamma|\mathcal{B}_{1}|,

for some γ(0,1)\gamma\in(0,1) (because of the density estimates). As in [19, Lemma 2.4], by computing the first variation of the functional 𝒥(,BR)\mathcal{J}(\cdot,B_{R}) with respect to a direction ξei\xi e^{i}, with ξC({|G|>0})\xi\in C^{\infty}(\{\big{\lvert}{G}\big{\rvert}>0\}), we deduce that the blow-up limit satisfies

{Lag0i=0in +n+1yag0i=0on {|G0|>0}ng0i=0on n{|G0|>0}.\begin{cases}L_{a}g^{i}_{0}=0&\mbox{in }\mathbb{R}^{n+1}_{+}\\ -\partial^{a}_{y}g^{i}_{0}=0&\mbox{on }\{|G_{0}|>0\}\cap\mathbb{R}^{n}\\ g^{i}_{0}=0&\mbox{on }\mathbb{R}^{n}\setminus\{|G_{0}|>0\}.\end{cases}

Hence, in view of Remark 5.6 all the components are equal up to a multiplicative constant. Moreover by nondegeneracy, G0G_{0} cannot be identically zero.

Thus, there exists ξm\xi\in\mathbb{R}^{m} such that |ξ|=1\big{\lvert}{\xi}\big{\rvert}=1 and G0=ξgG_{0}=\xi g, where |G0|=g|G_{0}|=g and gg is a global minimizer of (5.9). Indeed, for every R>0R>0 let g~Hloc1,a(n)\tilde{g}\in H^{1,a}_{{\tiny{\mbox{loc}}}}(\mathbb{R}^{n}) be such that supp(gg~)BR\mbox{supp}(g-\tilde{g})\subseteq B_{R}. Then, given the competitor G~=ξg~\widetilde{G}=\xi\tilde{g}, we easily get that 𝒥(G0,BR)𝒥(G~,BR)\mathcal{J}(G_{0},B_{R})\leq\mathcal{J}(\widetilde{G},B_{R}) is equivalent to

BR|y|a|g|2dX+Λ~n(R{g>0})BR|y|a|g~|2dX+Λ~n(R{g~>0}).\int_{B_{R}}{|y|^{a}\big{\lvert}{\nabla g}\big{\rvert}^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}_{R}\cap\{g>0\})\leq\int_{B_{R}}{|y|^{a}\big{\lvert}{\nabla\tilde{g}}\big{\rvert}^{2}\mathrm{d}X}+\tilde{\Lambda}\mathcal{L}_{n}(\mathcal{B}_{R}\cap\{\tilde{g}>0\}).

The desired claims now follow by the known results for the scalar case (for s=1/2s=1/2 see [14, Proposition 5.3], for s(0,1)s\in(0,1) see [21, Section 5.2]). ∎

Remark 5.8.

The constant in the expression (5.10) is different from the one appearing in [11]. Unfortunately, this is due to a computational mistake already noticed by [14, 24]: we refer to [24, Proposition 2.1.] for a detailed computation of the explicit constant in front of the one-plane solution UU.

In the same fashion of [19], the problem of the existence of singular points coincide with the existence of singular global minimizer of the scalar problem (5.9). Indeed, by [14, 21] we have

(5.12) n=inf{k:there exists an s-homogeneous global minimizer of (5.9) with singularity in zero}3.n^{*}=\inf\{k\in\mathbb{N}\colon\text{there exists an $s$-homogeneous global minimizer of \eqref{J} with singularity in zero}\}\geq 3.
Corollary 5.9.

Let n<nn<n^{*}, then every blow-up limit G0G_{0} is of the form (5.10).

Finally, we introduce the notion of regular and singular part of F(G)F(G).

Definition 5.10.

Let ΩD\Omega\subset D be a shape optimizer to (1.1) and GG be the vector of normalized eigenfunctions. Set X0F(G)X_{0}\in F(G), we say that

  • X0X_{0} is a regular point in Reg(F(G))\mathrm{Reg}(F(G)), if the Lebesgue density of {|G|>0}\{|G|>0\} at X0X_{0} is 1/21/2;

  • X0X_{0} is a singular point in Sing(F(G))\mathrm{Sing}(F(G)), if X0Reg(F(G))X_{0}\not\in\mathrm{Reg}(F(G)).

We conclude the Section with two results on the strata of F(G)F(G): first we initiate the analysis of the regularity of regular part of F(G)F(G) by proving that Reg(F(G))\mathrm{Reg}(F(G)) is relatively open in F(G)F(G). Then, by adapting a well-known approach in the context of Weiss-type monotonicity formulas, we estimate the Hausdorff dimension of the Singular stratum with respect to the threshold nn^{*}.

Corollary 5.11.

The regular part Reg(F(G))\mathrm{Reg}(F(G)) is an open subset of F(G)F(G).

Proof.

This result is deeply based on the upper semi-continuity of the function X0W(X0,G,0+)X_{0}\mapsto W(X_{0},G,0^{+}). Thus, let X0Reg(F(G))X_{0}\in\mathrm{Reg}(F(G)) and suppose by contradiction that there exists a sequence (Xk)kSing(F(G))(X_{k})_{k}\in\mathrm{Sing}(F(G)) converging to X0X_{0}. By Corollary 5.4 and Proposition 5.7, we know that

W(X0,G,0+)=ωn2andW(Xk,G,0+)ωn(12+δ),W(X_{0},G,0^{+})=\frac{\omega_{n}}{2}\quad\mbox{and}\quad W(X_{k},G,0^{+})\geq\omega_{n}\left(\frac{1}{2}+\delta\right),

for some universal δ(0,1/2)\delta\in(0,1/2). On the other hand, fixed r>0r>0, the function XW(X,G,r)X\mapsto W(X,G,r) is continuous and so

W(X0,G,r)=limkW(Xk,G,r),W(X_{0},G,r)=\lim_{k\to\infty}W(X_{k},G,r),

Moreover, exploiting Theorem 4.1 and the monotonicity of (4.3)

W(X0,G,r)+σC¯rs=limkW(Xk,G,r)+σC¯rsωn(12+δ),W(X_{0},G,r)+\sigma\overline{C}r^{s}=\lim_{k\to\infty}W(X_{k},G,r)+\sigma\overline{C}r^{s}\geq\omega_{n}\left(\frac{1}{2}+\delta\right),

for some C¯>0\overline{C}>0 defined in Theorem 4.1. Finally, passing to the limit as r0r\to 0 we obtain W(X0,G,0+)ωn(1/2+δ)W(X_{0},G,0^{+})\geq\omega_{n}(1/2+\delta), in contradiction with the assumption X0{|G|>0}(1/2)X_{0}\in\{|G|>0\}^{(1/2)}. ∎

The following result on the smallness of the singular set is nowadays standard and our approach follows essentially the strategy developed in [42] and then generalized in [31] for the local counterpart of (1.1).

Proposition 5.12.

The singular part Sing(F(G))\mathrm{Sing}(F(G)) satisfies the following estimate:

  • if n<nn<n^{*}, Sing(F(G))\mathrm{Sing}(F(G)) is empty;

  • if n=nn=n^{*}, Sing(F(G))\mathrm{Sing}(F(G)) contains at most a finite number of isolated points;

  • if n>nn>n^{*}, the (nn)(n-n^{*})-dimensional Hausdorff measure of Sing(F(G))\mathrm{Sing}(F(G)) is locally finite in 1\mathcal{B}_{1};

where n3n^{*}\geq 3 is defined in (5.12).

Proof.

We split the proof in three cases:

  1. (i)

    if n<nn<n^{*}, then by Corollary 5.9, for every point X0F(G)X_{0}\in F(G) the blow-up limits are of the form (5.10) and so the Lebesgue density of {|G|>0}\{|G|>0\} at X0X_{0} is 1/21/2 and F(G)=Reg(F(G))F(G)=\mathrm{Reg}(F(G));

  2. (ii)

    if n=nn=n^{*}, suppose there exists a sequence (Xk)kSing(F(G))(X_{k})_{k}\subset\mathrm{Sing}(F(G)) such that XkX0Sing(F(G))X_{k}\to X_{0}\in\mathrm{Sing}(F(G)). Set rk=|XkX0|r_{k}=|X_{k}-X_{0}| and consider the blow-up sequence

    Gk(X)=GXk,rk(X)=1rksG(Xk+rkX),G_{k}(X)=G_{X_{k},r_{k}}(X)=\frac{1}{r_{k}^{s}}G(X_{k}+r_{k}X),

    which converges to some blow-up limit G0G_{0}.

    Case 1: if Sing(F(G0)){0}\mathrm{Sing}(F(G_{0}))\setminus\{0\}\neq\emptyset, then there is a direction ξSn1F(G0)\xi\in S^{n-1}\cap F(G_{0}) such that tξt\xi is a singular point of F(G0)F(G_{0}), for every t>0t>0. Thus, by performing a blow-up analysis of G0G_{0} at ξ\xi we will get a new blow-up limit G00G_{00} such that

    {tξ:t}Sing(F(G00))\{t\xi\colon t\in\mathbb{R}\}\subset\mathrm{Sing}(F(G_{00}))

    and the whole function is invariant along the direction ξSn1\xi\in S^{n-1}. Finally, by a standard dimension reduction argument, the restriction |G00||n|G_{00}|\lvert_{\mathbb{R}^{n}} is a global minimizer of (5.9) with a non-trivial singular set, in contradiction with the definition of nn^{*}.

    Case 2: if Sing(F(G0)){0}=\mathrm{Sing}(F(G_{0}))\setminus\{0\}=\emptyset, then ξk=(X0Xk)rk1ξSn1Reg(F(G0))\xi_{k}=(X_{0}-X_{k})r_{k}^{-1}\to\xi\in S^{n-1}\cap\mathrm{Reg}(F(G_{0})). By Corollary 5.4, Proposition 5.7, there exists R>0R>0 such that

    W(ξ,G0,R)ωn(12+δ4),W(\xi,G_{0},R)\leq\omega_{n}\left(\frac{1}{2}+\frac{\delta}{4}\right),

    for δ(0,1/2)\delta\in(0,1/2) as in (5.11). By the convergence of GkG_{k}, we first get

    W(ξ,Gk,R)ωn(12+δ3),W(\xi,G_{k},R)\leq\omega_{n}\left(\frac{1}{2}+\frac{\delta}{3}\right),

    which implies that

    W(ξk,Gk,R)W(ξ0,Gk,R)+C(n,s)([|G|]C0,s2+Λ~nωn)(|ξkξ0|R)min{2s,1},W(\xi_{k},G_{k},R)\leq W(\xi_{0},G_{k},R)+C(n,s)\left([|G|]^{2}_{C^{0,s}}+\tilde{\Lambda}n\omega_{n}\right)\left(\frac{|\xi_{k}-\xi_{0}|}{R}\right)^{\min\{2s,1\}},

    for |ξ0ξk|<R|\xi_{0}-\xi_{k}|<R and small enough. Thus, by choosing kk large enough, we get that

    W(X0,G,rkr0)=W(ξk,Gk,r0)ωn(12+δ2),W(X_{0},G,r_{k}r_{0})=W(\xi_{k},G_{k},r_{0})\leq\omega_{n}\left(\frac{1}{2}+\frac{\delta}{2}\right),

    in contradiction with the assumption X0Sing(F(G))X_{0}\in\mathrm{Sing}(F(G));

  3. (iii)

    if n>nn>n^{*}, assume that for some t>0t>0 we have nn+t(Sing(F(G)))>0\mathcal{H}^{n-n^{*}+t}(\mathrm{Sing}(F(G)))>0. Then, it can be proved that there exists some point X0Sing(F(G))X_{0}\in\mathrm{Sing}(F(G)) and a blow-up limit G0G_{0} at X0X_{0} such that nn+t(Sing(F(G0)))>0\mathcal{H}^{n-n^{*}+t}(\mathrm{Sing}(F(G_{0})))>0. Since |G0||G_{0}| is a global minimizer of (5.9), this is in contradiction with the estimates of the dimension of the singular set in the scalar case (see [21]).

6. Viscosity formulation on Reg(F(G))\mathrm{Reg}(F(G))

In this short Section we recall some basic facts about the scalar thin one-phase free boundary problem, and we state the viscosity formulation satisfied by the normalized eigenfunctions on optimal domains near the free boundary.
We show that the eigenfunctions are indeed viscosity solutions of thin-problem. Thus, the study of the regular part of the free boundary can be performed with the viscosity approach of [12, 17, 18, 19] trying to reduce to the method carried out in the scalar case [17] with the methodology introduced in [19], both for s=1/2s=1/2.

6.1. The viscosity problem associated to (5.9)

In this subsection we collect basic definitions and results for the scalar thin one-phase problem arising from the critical condition of the functional (5.9). Consider

(6.1) {Lag=0,in B1+(g):=B1{(x,0):g(x,0)=0},gts=α,on F(g):=1n{(x,0):g(x,0)>0},\begin{cases}L_{a}g=0,\quad\textrm{in $B_{1}^{+}(g):=B_{1}\setminus\{(x,0):g(x,0)=0\},$}\\ \frac{\partial g}{\partial t^{s}}=\alpha,\quad\textrm{on $F(g):=\mathcal{B}_{1}\cap\partial_{\mathbb{R}^{n}}\{(x,0):g(x,0)>0\}$},\end{cases}

where α>0\alpha>0 and

(6.2) gts(x0):=limt0+g(x0+tν(x0),0)ts,X0=(x0,0)F(g),\dfrac{\partial g}{\partial t^{s}}(x_{0}):=\displaystyle\lim_{t\rightarrow 0^{+}}\frac{g(x_{0}+t\nu(x_{0}),0)}{t^{s}},\quad\textrm{$X_{0}=(x_{0},0)\in F(g)$},

with ν(x0)\nu(x_{0}) the unit normal to the free boundary F(g)F(g) at x0x_{0} pointing toward 1+(g)\mathcal{B}_{1}^{+}(g). For further details and proofs, we refer the reader to [11, 12, 14, 13, 15].

First, we state the notion of viscosity solutions to (6.1), as introduced in [12].

Definition 6.1.

Given g,φg,\varphi continuous, we say that φ\varphi touches gg by below (resp. above) at X0B1X_{0}\in B_{1} if g(X0)=φ(X0),g(X_{0})=\varphi(X_{0}), and

g(X)φ(X)(resp. g(X)φ(X))in a neighborhood O of X0.g(X)\geq\varphi(X)\quad(\text{resp. $g(X)\leq\varphi(X)$})\quad\text{in a neighborhood $O$ of $X_{0}$.}

If this inequality is strict in O{X0}O\setminus\{X_{0}\}, we say that φ\varphi touches gg strictly by below (resp. above).

Definition 6.2.

We say that φC(B1)\varphi\in C(B_{1}) is a strict comparison subsolution (resp. supersolution) to (6.1) if φ\varphi is a non-negative function in B1B_{1} which is even with respect to {y=0}\{y=0\} and it satisfies

  1. (i)

    φ\varphi is C2C^{2} and Laφ0L_{a}\varphi\geq 0 (resp. Laφ0L_{a}\varphi\leq 0) in B1+(φ)B_{1}^{+}(\varphi);

  2. (ii)

    F(φ)F(\varphi) is C2C^{2} and if x0F(φ)x_{0}\in F(\varphi) we have

    (6.3) φ(x0+tν(x0),z)=α(x0)U(t,z)+o(|(t,z)|s),as |(t,z)|0+,\varphi(x_{0}+t\nu(x_{0}),z)=\alpha(x_{0})U(t,z)+o(|(t,z)|^{s}),\quad\textrm{as $|(t,z)|\rightarrow 0^{+},$}

    with

    α(x0)α(resp. α(x0)α),\alpha(x_{0})\geq\alpha\quad\text{(resp. $\alpha(x_{0})\leq\alpha$)},

    where ν(x0)\nu(x_{0}) denotes the unit normal at x0x_{0} to F(φ)F(\varphi) pointing toward 1+(φ);\mathcal{B}_{1}^{+}(\varphi);

  3. (iii)

    Either φ\varphi is not harmonic in B1+(φ)B_{1}^{+}(\varphi) or α(x0)>α\alpha(x_{0})>\alpha (resp. α(x0)<α\alpha(x_{0})<\alpha) at all x0F(φ).x_{0}\in F(\varphi).

Notice that if F(φ)F(\varphi) is C2C^{2} then any function φ\varphi which is harmonic in B1+(φ)B^{+}_{1}(\varphi) has an asymptotic expansion at a point X0F(φ),X_{0}\in F(\varphi),

(6.4) φ(x,y)=α(x0)U((xx0)ν(x0),y)+o(|xx0|s+ys).\varphi(x,y)=\alpha(x_{0})U((x-x_{0})\cdot\nu(x_{0}),y)+o(|x-x_{0}|^{s}+y^{s}).
Definition 6.3.

We say that gg is a viscosity solution to (6.1) if gg is a continuous non-negative function in B1B_{1} which is even with respect to {y=0}\{y=0\} and it satisfies

  1. (i)

    Lag=0L_{a}g=0  in B1+(g)B_{1}^{+}(g);

  2. (ii)

    Any (strict) comparison subsolution (resp. supersolution) cannot touch gg by below (resp. by above) at a point X0=(x0,0)F(g).X_{0}=(x_{0},0)\in F(g).

Let us introduce the notion of ε\varepsilon-domain variation first introduced in [12] for the case s=1/2s=1/2 and then generalized in [18]. This methodology allows to “linearize" the problem (6.1), as long as an appropriate Harnack type inequality is established. Recently, this strategy has been adapted to the scalar thin almost-minimizers in [17] and to vectorial thin one-phase problem in [19].

We denote by PP the half-hyperplane

P:={Xn+1:xn0,y=0}P:=\{X\in\mathbb{R}^{n+1}:x_{n}\leq 0,y=0\}

and by

L:={Xn+1:xn=0,y=0}.L:=\{X\in\mathbb{R}^{n+1}:x_{n}=0,y=0\}.

Let gg be a continuous non-negative function in B¯ρ\overline{B}_{\rho}. We define the multivalued map g~\tilde{g} which associate to each Xn+1PX\in\mathbb{R}^{n+1}\setminus P the set g~(X)\tilde{g}(X)\subset\mathbb{R} via the formula

(6.5) U(X)=g(Xwen),wg~(X).U(X)=g(X-we_{n}),\quad\forall w\in\tilde{g}(X).

We write g~(X)\tilde{g}(X) to denote any of the values in this set.

Notice that if g satisfies

(6.6) U(Xεen)g(X)U(X+εen)in Bρ, for ε>0U(X-\varepsilon e_{n})\leq g(X)\leq U(X+\varepsilon e_{n})\quad\textrm{in $B_{\rho},$ for $\varepsilon>0$}

then g~(X)\tilde{g}(X)\neq\emptyset for XBρεPX\in B_{\rho-\varepsilon}\setminus P and |g~(X)|ε,|\tilde{g}(X)|\leq\varepsilon, thus we can associate to gg a possibly multi-valued map g~\tilde{g} defined at least on BρεPB_{\rho-\varepsilon}\setminus P and taking values in [ε,ε][-\varepsilon,\varepsilon] which satisfies

(6.7) U(X)=g(Xg~(X)en).U(X)=g(X-\tilde{g}(X)e_{n}).

Moreover if gg is strictly monotone along the ene_{n}-direction in Bρ+(g)B^{+}_{\rho}(g), then g~\tilde{g} is also single-valued.
We refer to [18, Section 2] and [12, Section 3] for other basic properties of the ε\varepsilon-domain variations.

6.2. The viscosity formulation of the eigenvalue problem.

Consider now the vector valued thin type problem

(6.8) {LaG=0in Br{y=0};yaG=σGin r+(G);ts|G|=αon F(G)r,\begin{cases}-L_{a}G=0&\text{in $B_{r}\setminus\{y=0\}$;}\\ -\partial^{a}_{y}G=\sigma G&\text{in $\mathcal{B}_{r}^{+}(G)$;}\\ \frac{\partial}{\partial t^{s}}|G|=\alpha&\text{on $F(G)\cap\mathcal{B}_{r},$}\end{cases}

where α>0\alpha>0 and σ=(σ1,,σm)\sigma=(\sigma^{1},\dots,\sigma^{m}), with σi>0\sigma^{i}>0 for i=1,,mi=1,\dots,m.

Definition 6.4.

We say that G=(g1,,gm)C(Br,m)G=(g^{1},\ldots,g^{m})\in C(B_{r},\mathbb{R}^{m}) is a viscosity solution to (6.8) in BrB_{r} if each gig^{i} is even with respect to {y=0}\{y=0\},

(6.9) {Lagi=0in Br{y=0}yagi=σigiin r+(G),i=1,,m,\begin{cases}-L_{a}g^{i}=0&\mbox{in }B_{r}\setminus\{y=0\}\\ -\partial^{a}_{y}g^{i}=\sigma^{i}g^{i}&\mbox{in }\mathcal{B}_{r}^{+}(G)\end{cases},\quad\forall i=1,\ldots,m,

and the free boundary condition is satisfied in the following sense. Given X0F(G)X_{0}\in F(G), and a continuous function φ\varphi in a neighborhood of X0X_{0}, then

  1. (i)

    If φ\varphi is a strict comparison subsolution to (6.1), then for all unit directions ff, G,f\langle G,f\rangle cannot be touched by below by φ\varphi at X0.X_{0}.

  2. (ii)

    If φ\varphi is a strict comparison supersolution to (6.1), then |G|\big{\lvert}{G}\big{\rvert} cannot be touched by above by φ\varphi at X0.X_{0}.

In the next proposition we prove that the vector of normalized eigenfunctions associated to shape optimizers of (1.1) are indeed viscosity solutions to (6.8) in B1B_{1} for some specific σm\sigma\in\mathbb{R}^{m} and α>0\alpha>0.

Proposition 6.5.

Let GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions associated to a shape optimizer Ω\Omega of (1.1). Then, GG is a viscosity solution of

(6.10) {LaG=0in n+1{y=0};yaG=λGin {|G|>0}n;ts|G|=ΛΓ(1+s)on F(G)with λ=(λ1s(Ω),,λms(Ω)),\begin{cases}-L_{a}G=0&\text{in $\mathbb{R}^{n+1}\setminus\{y=0\}$;}\\ -\partial^{a}_{y}G=\lambda G&\text{in $\{|G|>0\}\cap\mathbb{R}^{n}$;}\\ \frac{\partial}{\partial t^{s}}|G|=\frac{\sqrt{\Lambda}}{\Gamma(1+s)}&\text{on $F(G)$}\end{cases}\quad\text{with $\lambda=(\lambda_{1}^{s}(\Omega),\dots,\lambda_{m}^{s}(\Omega))$,}

in the sense of Definition 6.4.

Proof.

By (2.13) we already know that the first two conditions of (6.9) are satisfied. Hence, let φ\varphi be a strict comparison subsolution to (6.1), and suppose by contradiction that there exists a unit direction ff in m\mathbb{R}^{m} such that G,f\langle G,f\rangle is touched by below by φ\varphi at Y0F(G)Y_{0}\in F(G).
Consider now the blow-up sequences centered in the touching point

Gk(X)=1rksG(Y0+rkX)andφk(X)=1rksφ(Y0+rkX),G_{k}(X)=\frac{1}{r_{k}^{s}}G(Y_{0}+r_{k}X)\quad\mbox{and}\quad\varphi_{k}(X)=\frac{1}{r_{k}^{s}}\varphi(Y_{0}+r_{k}X),

for some sequence of radii rk0+r_{k}\to 0^{+}. Up to a subsequence, they converge respectively to some G0G_{0} and φ0\varphi_{0} uniformly on every compact set of n+1\mathbb{R}^{n+1}. By Definition 6.2, we get, up to rotation, that

(6.11) φ0(X)=αU(xn,y)with α>ΛΓ(1+s),\varphi_{0}(X)=\alpha U\left(x_{n},y\right)\quad\text{with $\alpha>\frac{\sqrt{\Lambda}}{\Gamma(1+s)}$},

On the other side, by Proposition 5.7 the norm |G0||G_{0}| is a ss-homogeneous global minimizer of the scalar thin one-phase functional (5.9) such that {|G0|>0}{y=0}{y=0,xn>0}\{|G_{0}|>0\}\cap\{y=0\}\supset\{y=0,x_{n}>0\}. By Remark 5.6, we deduce that {|G0|=0}{y=0}=P\{|G_{0}|=0\}\cap\{y=0\}=P and consequently

(6.12) G0(X)=ΛΓ(1+s)U(xn,y)ξwhereξm,|ξ|=1.G_{0}(X)=\frac{\sqrt{\Lambda}}{\Gamma(1+s)}U(x_{n},y)\xi\quad\mbox{where}\quad\xi\in\mathbb{R}^{m},\big{\lvert}{\xi}\big{\rvert}=1.

Hence, we immediately deduce that

φ0(X)=αU(xn,y)G0,f=ΛΓ(1+s)ξ,fU(xn,y),\varphi_{0}(X)=\alpha U(x_{n},y)\leq\langle G_{0},f\rangle=\frac{\sqrt{\Lambda}}{\Gamma(1+s)}\langle\xi,f\rangle U(x_{n},y),

in contradiction with the hypothesis (6.11).
On the other hand, let φ\varphi be a comparison strict supersolution and let us assume that |G|\big{\lvert}{G}\big{\rvert} is touched by above by φ\varphi at some Y0F(G)Y_{0}\in F(G). By the same blow-up procedure we get, up to rotation, that

φ0(X)=αU(xn,y)with α<ΛΓ(1+s),\varphi_{0}(X)=\alpha U\left(x_{n},y\right)\quad\text{with $\alpha<\frac{\sqrt{\Lambda}}{\Gamma(1+s)}$},

and that |G0||G_{0}| is a ss-homogeneous global minimizer of the scalar thin one-phase functional (5.9) such that {|G0|=0}{y=0}P\{|G_{0}|=0\}\cap\{y=0\}\subset P. As before, we get (6.12) and since |G0|φ0\big{\lvert}{G_{0}}\big{\rvert}\leq\varphi_{0}, the absurd follows from the fact that α<ΛΓ(1+s)\alpha<\frac{\sqrt{\Lambda}}{\Gamma(1+s)}. ∎

Remark 6.6.

We remark that if GG is a viscosity solution to (6.8) in Br(X0)B_{r}(X_{0}) for some α>0,λm,X0F(G)\alpha>0,\lambda\in\mathbb{R}^{m},X_{0}\in F(G), then

GX0,r(X)=1rsG(X0+rX),XB1G_{X_{0},r}(X)=\frac{1}{r^{s}}G(X_{0}+rX),\quad X\in B_{1}

is a viscosity solution to (6.8) in B1B_{1} with α>0\alpha>0 and σ~=rsσm\tilde{\sigma}=r^{s}\sigma\in\mathbb{R}^{m}.
Similarly, by Proposition 6.5, given a vector GG of normalized eigenfunctions associated to a shape optimizer Ω\Omega and X0F(G),r>0X_{0}\in F(G),r>0, we get that GX0,rG_{X_{0},r} is a viscosity solution to

(6.13) {LaGX0,r=0in n+1{y=0};yaGX0,r=(λrs)GX0,rin {|GX0,r|>0}n;ts|GX0,r|=ΛΓ(1+s)on F(GX0,r),\begin{cases}-L_{a}G_{X_{0},r}=0&\text{in $\mathbb{R}^{n+1}\setminus\{y=0\}$;}\\ -\partial^{a}_{y}G_{X_{0},r}=\left(\lambda r^{s}\right)G_{X_{0},r}&\text{in $\{|G_{X_{0},r}|>0\}\cap\mathbb{R}^{n}$;}\\ \frac{\partial}{\partial t^{s}}|G_{X_{0},r}|=\frac{\sqrt{\Lambda}}{\Gamma(1+s)}&\text{on $F(G_{X_{0},r})$},\end{cases}

with λ=(λ1s(Ω),,λms(Ω))\lambda=(\lambda_{1}^{s}(\Omega),\dots,\lambda_{m}^{s}(\Omega)). The behaviour of (6.8), under translations and rescaling, reflects that for r>0r>0 sufficiently small, the solutions near F(G)F(G) resemble the ones with σ=0\sigma=0.

Remark 6.7.

Notice that, if GG is a viscosity solution to (6.8) for some α>0,σm\alpha>0,\sigma\in\mathbb{R}^{m}, as in Remark 3.4, the function

|G|+|σ|1a|y|1a|G|+\frac{|\sigma|}{1-a}|y|^{1-a}

is a viscosity subsolution to the scalar thin one-phase problem (6.1) in the sense of Definition 6.3. Indeed, by the free boundary condition in Definition 6.4, we easily deduce the validity of its scalar counterpart in Definition 6.3.

7. Regularity of Reg(F(G))\mathrm{Reg}(F(G))

In this Section we start by introducing the basic tools for our study of the regular part Reg(F(G))\mathrm{Reg}(F(G)) and then we state and prove an Harnack type inequality which is crucial for the linearization. Lastly, we conclude the Section with an improvement of flatness type lemma, from which the main result of C1,αC^{1,\alpha} regularity for flat free boundary follows by standard arguments (see for example [20, 19]).

7.1. Flat solutions

In view of Definition 5.10, Proposition 6.5, Remark 6.6 and the non-degeneracy property, we can reduce our analysis to understanding flat viscosity solutions defined below.

Definition 7.1.

Let α=1\alpha=1 and GG be a viscosity solution to (6.8) in B1B_{1} for some σ>0\sigma>0. We say that GG is ε\varepsilon-flat in the (f,ν)(f,\nu)-directions in B1B_{1}, if there exist some unit directions fm,νnf\in\mathbb{R}^{m},\nu\in\mathbb{R}^{n},

(7.1) |G(X)U(x,ν,y)f|εin B1,|G(X)-U(\langle x,\nu\rangle,y)f|\leq\varepsilon\quad\text{in $B_{1}$,}

and

(7.2) |G|0in 1{x,ν<ε}.|G|\equiv 0\quad\text{in $\mathcal{B}_{1}\cap\{\langle x,\nu\rangle<-\varepsilon\}$}.

For the sake of simplicity we consider the notion of ε\varepsilon-flat solution for α=1\alpha=1, but, up to a multiplicative constant, this is not a restrictive assumption.

In [19], in collaboration with D. De Silva we introduce a two different approach in order to prove the validity of an Harnack type inequality near the free boundary in the case of vectorial problems. More precisely, in the thin case the reduction to a scalar problem just requires the construction of two appropriate barriers. In our case, we will adapt the same strategy to the case of almost-minimizer.

Remark 7.1.

Let GHΩ1,a(n+1;m)G\in H^{1,a}_{\Omega}(\mathbb{R}^{n+1};\mathbb{R}^{m}) be the vector of normalized eigenfunctions on a shape optimizer Ω\Omega. Then, by the minimum principle [5, Theorem 2.8.], we already know that

g1>0in {|G|>0}.g^{1}>0\quad\text{in $\{|G|>0\}$.}

This observations is fundamental and allows to treat g1g^{1} as a supersolution to the scalar thin one-phase problem (6.1) in the sense of Definition 6.2. Notice that, as in [19, Proposition 6.2.], the positivity of g1g^{1} can actually be proved just by assuming the flatness condition (7.1).

The following result translates the flatness hypothesis on the vector-valued function GG into the property that its first component is trapped between nearby translation of a one-plane solution of (6.1), while the remaining ones are sufficiently small.

Lemma 7.2.

Let GG be a viscosity solution to (6.8) in B1B_{1} for α=1,σ>0\alpha=1,\sigma>0. There exists ε0>0\varepsilon_{0}>0 universal such that, if G is ε0\varepsilon_{0}-flat in the (f1,en)(f^{1},e_{n})-directions in B1B_{1}, then

  1. (i)

    for i=2,,m,i=2,\ldots,m,

    (7.3) |gi|Cε0U(X+ε0en)in B1/2;|g^{i}|\leq C\varepsilon_{0}U(X+\varepsilon_{0}e_{n})\quad\text{in $B_{1/2};$}
  2. (ii)
    (7.4) U(XCε0en)g1|G|U(X+Cε0en)in B1/2,U(X-C\varepsilon_{0}e_{n})\leq g^{1}\leq|G|\leq U(X+C\varepsilon_{0}e_{n})\quad\text{in $B_{1/2}$},

with C>0C>0 universal.

Proof.

For the bound (i),(i), notice first that ya|gi|λi|gi|λiε0-\partial^{a}_{y}|g^{i}|\leq\lambda^{i}|g^{i}|\leq\lambda^{i}\varepsilon_{0} in B1B_{1} and so the function

hi=|gi|+λiε01a|y|1ah^{i}=|g^{i}|+\frac{\lambda^{i}\varepsilon_{0}}{1-a}|y|^{1-a}

is LaL_{a}-subharmonic in B1B_{1} and it satisfies

hiCε0,hi0on {X1:xnε0},h^{i}\leq C\varepsilon_{0},\quad h^{i}\equiv 0\quad\text{on $\{X\in\mathcal{B}_{1}\colon x_{n}\leq-\varepsilon_{0}\}$},

for some dimensional constant C>0C>0. Then, let vv be the LaL_{a}-harmonic function in B1{X1:xn<ε0}B_{1}\setminus\{X\in\mathcal{B}_{1}\colon x_{n}<-\varepsilon_{0}\} such that

v=Cε0on B1,v=0on {X1:xnε0}.v=C\varepsilon_{0}\quad\text{on $\partial B_{1}$},\quad v=0\quad\text{on $\{X\in\mathcal{B}_{1}\colon x_{n}\leq-\varepsilon_{0}\}$}.

Hence, by comparison principle hivh^{i}\leq v in B1B_{1} and so |gi|v|g^{i}|\leq v in B1B_{1}. Then by the boundary Harnack inequality, say for X¯=12en\bar{X}=\frac{1}{2}e_{n}, we deduce

v(X)C¯v(X¯)U(X¯+ε0en)U(X+ε0en)Cε0U(X+ε0en)in B1/2,v(X)\leq\bar{C}\frac{v(\bar{X})}{U(\bar{X}+\varepsilon_{0}e_{n})}U(X+\varepsilon_{0}e_{n})\leq C\varepsilon_{0}U(X+\varepsilon_{0}e_{n})\quad\text{in $B_{1/2}$},

with C>0C>0 universal.
For the bounds in (ii),(ii), we already know by Remark 7.1 that g1g^{1} is strictly positive and it satisfies

U(X)ε0g1(X)U(X)+ε0in B1,U(X)-\varepsilon_{0}\leq g^{1}(X)\leq U(X)+\varepsilon_{0}\quad\text{in $B_{1}$},

Moreover, by taking ε0\varepsilon_{0} possibly smaller, we have

{X1:xnε0}{X1:g1=0}{X1:xnε0}.\{X\in\mathcal{B}_{1}\colon x_{n}\leq-\varepsilon_{0}\}\subset\{X\in\mathcal{B}_{1}\colon g^{1}=0\}\subset\{X\in\mathcal{B}_{1}\colon x_{n}\leq\varepsilon_{0}\}.

Now, since yag10-\partial^{a}_{y}g^{1}\geq 0 in 1+(g1)\mathcal{B}_{1}^{+}(g^{1}) we get that gig^{i} is LaL_{a}-superharmonic in B1+(g1)B_{1}^{+}(g^{1}).
Thus, according to the proof of [12, Lemma 5.3.] for the case s=1/2s=1/2, we get the lower bound in (ii)(ii): here details are omitted as they apply verbatim by considering harmonic function with respect to the LaL_{a}-operator.
Now, since g1|G|g^{1}\leq|G|, the upper bound follows by exploiting the observations of Remark 3.4. Thus, consider

v(x,y)=|G|(x,y)+C¯|y|1a,with C¯=kC~n,sn/4s1aλks(Ω)n+4s4s.v(x,y)=|G|(x,y)+\overline{C}|y|^{1-a},\quad\mbox{with }\overline{C}=\frac{\sqrt{k}\tilde{C}_{n,s}^{n/4s}}{1-a}\lambda^{s}_{k}(\Omega)^{\frac{n+4s}{4s}}.

Let f1f_{1} be the LaL_{a}-harmonic function in the set

B1ε0=B1{X1:xnε0},B_{1}^{\varepsilon_{0}}=B_{1}\setminus\{X\in\mathcal{B}_{1}\colon x_{n}\leq-\varepsilon_{0}\},

such that f1=vf_{1}=v on B1\partial B_{1} and f1=0f_{1}=0 on {X1:xnε0}\{X\in\mathcal{B}_{1}\colon x_{n}\leq-\varepsilon_{0}\}. Since vv is LaL_{a}-subharmonic, by the maximum principle it follows that

(7.5) vf1in B1¯|G|f1in B1¯.v\leq f_{1}\quad\text{in $\overline{B_{1}}$}\longmapsto|G|\leq f_{1}\quad\text{in $\overline{B_{1}}$}.

Finally, we can prove that there exits C>0C>0 universal such that

f1(X)U(X+Cε0en)in B1/2,f_{1}(X)\leq U(X+C\varepsilon_{0}e_{n})\quad\text{in $B_{1/2}$},

which implies the claimed result. However, since the proof of this inequality follows essentially the ideas of the one of [12, Lemma 5.3.], by replacing the Laplacian with the LaL_{a}-operator, we omit the details. ∎

7.2. Harnack type inequality

In the case of local minimizers of vectorial problems with thin free boundary [19, Lemma 6.6.], with D. De Silva we prove an Harnack inequality by using the observation that |G||G| and g1g^{1} are respectively a subsolution and a supersolution for the scalar one phase problem in B1B_{1}, which means that the strategy of the scalar case applies straightforwardly also in that context (see [20, Lemma 2.4] for the similar result for the local case).
In the case of normalized eigenfunctions for the fractional Laplacian, we adapt the same strategy by using that locally the eigenvalue problem (6.8) resembles the vectorial one-phase problem in [19] (see Remark 6.6).
Therefore, most details are omitted as the results of [18] (or [12] for s=1/2s=1/2) can be applied directly, after observing that in their proofs it is enough for the function to be either a subsolution or a supersolution of (6.1) (depending on the desired bound).
As pointed out in [17], alternatively one could prove the validity of an Harnack inequality by specifying the size of the neighborhood around the contact point between the solution and some explicit barriers.

Theorem 7.3.

Let α=1\alpha=1 and GG be a solution of (6.8) in B1B_{1} satisfying

𝒥(G,B1)𝒥(G~,B1)+σ,\mathcal{J}(G,B_{1})\leq\mathcal{J}(\tilde{G},B_{1})+\sigma,

for every G~GH01,a(B1;m)\tilde{G}-G\in H^{1,a}_{0}(B_{1};\mathbb{R}^{m}), with σε2(n+2)\sigma\leq\varepsilon^{2(n+2)}. Then, there exists a universal constant ε¯>0\overline{\varepsilon}>0 such that if

(7.6) U(X+εa0en)g1|G|U(X+εb0en)in Br(X0)B1,U(X+\varepsilon a_{0}e_{n})\leq g^{1}\leq|G|\leq U(X+\varepsilon b_{0}e_{n})\quad\text{in $B_{r}(X_{0})\subset B_{1}$,}

with

ε(b0a0)ε¯r,\varepsilon(b_{0}-a_{0})\leq\bar{\varepsilon}r,

and

(7.7) |gi|rs(b0a0rε)5/8in B1/2(X0),  i=2,…, m,|g^{i}|\leq r^{s}\left(\frac{b_{0}-a_{0}}{r}\varepsilon\right)^{5/8}\quad\text{in $B_{1/2}(X_{0})$, \quad i=2,\ldots, m,}

then

(7.8) U(X+εa1en)g1|G|U(X+εb1en) in Bηr(X0),U(X+\varepsilon a_{1}e_{n})\leq g^{1}\leq|G|\leq U(X+\varepsilon b_{1}e_{n})\quad\text{ in $B_{\eta r}(X_{0})$,}

with

a0a1b1b0,b1a1=(1η)(b0a0),a_{0}\leq a_{1}\leq b_{1}\leq b_{0},\quad b_{1}-a_{1}=(1-\eta)(b_{0}-a_{0}),

for a small universal constant η>0\eta>0.

Here g~ε1\tilde{g}^{1}_{\varepsilon} and |Gε|~\widetilde{|G_{\varepsilon}|} are the ε\varepsilon-domain variations associated to g1g^{1} and |G||G| respectively and

aε:={(X,g~ε1(X)):XB1εP}andAε:={(X,|Gε|~(X)):XB1εP}.a_{\varepsilon}:=\left\{(X,\tilde{g}^{1}_{\varepsilon}(X))\colon X\in B_{1-\varepsilon}\setminus P\right\}\quad\text{and}\quad A_{\varepsilon}:=\left\{(X,\widetilde{|G_{\varepsilon}|}(X))\colon X\in B_{1-\varepsilon}\setminus P\right\}.

Since domain variations may be multivalued, we mean that given XX all pairs (X,g~ε1(X))(X,\tilde{g}^{1}_{\varepsilon}(X)) and (X,|Gε|~(X))(X,\widetilde{|G_{\varepsilon}|}(X)) belong respectively to aεa_{\varepsilon} and AεA_{\varepsilon}, for all possible values of the ε\varepsilon-domain variations. The following key corollary follows directly from the previous result.

Corollary 7.4.

Let α=1\alpha=1 and GG be a solution of (6.8) in B1B_{1} such that

𝒥(G,B1)𝒥(G~,B1)+σ,\mathcal{J}(G,B_{1})\leq\mathcal{J}(\tilde{G},B_{1})+\sigma,

for every G~GH01,a(B1;m)\tilde{G}-G\in H^{1,a}_{0}(B_{1};\mathbb{R}^{m}), with σε2(n+2)\sigma\leq\varepsilon^{2(n+2)}. Then, there exists a universal constant ε¯>0\overline{\varepsilon}>0 such that if

U(Xεen)g1|G|U(X+εen)in B1,U(X-\varepsilon e_{n})\leq g^{1}\leq|G|\leq U(X+\varepsilon e_{n})\quad\text{in $B_{1}$,}

and

|gi|ε3/4in B1/2,  i=2,…, m,|g^{i}|\leq\varepsilon^{3/4}\quad\text{in $B_{1/2}$, \quad i=2,\ldots, m,}

with εε¯/2\varepsilon\leq\bar{\varepsilon}/2 and k0>0k_{0}>0 such that

(7.9) 4ε(1η)k0ηk0ε¯,εC(1η)5k0,4\varepsilon(1-\eta)^{k_{0}}\eta^{-k_{0}}\leq\overline{\varepsilon},\quad\varepsilon\leq C(1-\eta)^{5k_{0}},

for some C>0C>0 universal, then the sets aε(B1/2×[1,1])a_{\varepsilon}\cap(B_{1/2}\times[-1,1]) and Aε(B1/2×[1,1])A_{\varepsilon}\cap(B_{1/2}\times[-1,1]) are trapped above the graph of a function y=aε(X)y=a_{\varepsilon}(X) and below the graph of a function y=bε(X)y=b_{\varepsilon}(X) with

bεaε2(1η)k01,b_{\varepsilon}-a_{\varepsilon}\leq 2(1-\eta)^{k_{0}-1},

where aε,bεa_{\varepsilon},b_{\varepsilon} have modulus of continuity bounded by the Hölder function γtβ\gamma t^{\beta}, with γ,β\gamma,\beta depending only on η\eta.

Indeed, by iterating the Harnack inequality for k=0,,k0k=0,\ldots,k_{0} (the second inequality in (7.9) guarantees that (7.7) is preserved), we obtain

(7.10) U(X+εaken)g1|G|U(X+εbken)in BηkU(X+\varepsilon a_{k}e_{n})\leq g^{1}\leq|G|\leq U(X+\varepsilon b_{k}e_{n})\quad\text{in $B_{\eta^{k}}$}

with bkak=2(1η)kb_{k}-a_{k}=2(1-\eta)^{k}. Thus, by the properties of the ε\varepsilon-domain variations (see [18, Lemma 3.1] and [12, Lemma 3.1] for s=1/2s=1/2) we get

akg~ε1|Gε|~bkin Bηkε,a_{k}\leq\tilde{g}^{1}_{\varepsilon}\leq\widetilde{|G_{\varepsilon}|}\leq b_{k}\quad\text{in $B_{\eta^{k}-\varepsilon},$}

and

aε(Bηkε×[1,1])\displaystyle a_{\varepsilon}\cap(B_{\eta^{k}-\varepsilon}\times[-1,1]) Bηkε×[ak,bk],\displaystyle\subset B_{\eta^{k}-\varepsilon}\times[a_{k},b_{k}],
Aε(Bηkε×[1,1])\displaystyle A_{\varepsilon}\cap(B_{\eta^{k}-\varepsilon}\times[-1,1]) Bηkε×[ak,bk],\displaystyle\subset B_{\eta^{k}-\varepsilon}\times[a_{k},b_{k}],

for k=0,,k0.k=0,\ldots,k_{0}.

Lemma 7.5.

Let α=1\alpha=1 and GG be a solution to (6.8) in B1B_{1} for some σm\sigma\in\mathbb{R}^{m}. There exists ε0>0\varepsilon_{0}>0 universal such that, if for 0<εε00<\varepsilon\leq\varepsilon_{0}

U(X)g1(X)in B1/2,U(X)\leq g^{1}(X)\quad\text{in $B_{1/2},$}

and at X¯=(x¯,y¯)B1/8(14en)\overline{X}=(\bar{x},\bar{y})\in B_{1/8}(\frac{1}{4}e_{n}) we have U(X¯+εen)g1(X¯),U(\overline{X}+\varepsilon e_{n})\leq g^{1}(\overline{X}), then

U(X+τεen)g1(X)in Bδ,U(X+\tau\varepsilon e_{n})\leq g^{1}(X)\quad\text{in $B_{\delta}$},

for universal constants τ,δ>0\tau,\delta>0. On the other hand, by taking σε2(n+2)\sigma\leq\varepsilon^{2(n+2)}, if

|G|(X)U(X)in B1/2|G|(X)\leq U(X)\quad\text{in $B_{1/2}$}

and

(7.11) |gi|ε5/8in B1/2,i=2,,m,|g^{i}|\leq\varepsilon^{5/8}\quad\text{in $B_{1/2}$},\quad i=2,\ldots,m,

then if g1(X¯)U(X¯εen)g^{1}(\overline{X})\leq U(\overline{X}-\varepsilon e_{n}) we get

|G|(X)U(Xτεen) in Bδ,|G|(X)\leq U(X-\tau\varepsilon e_{n})\quad\text{ in $B_{\delta},$}

for some universal δ>0\delta>0.

Proof of Lemma 7.5.

The first statement follows immediately from the fact that g1g^{1} is a supersolution to (6.1) hence we can apply [18, Lemma 4.3] (see [12, Lemma 6.3] for s=1/2s=1/2).

Now, let us consider the case

|G|(X)U(X),|gi|ε5/8 for i=2,,m,|G|(X)\leq U(X),\quad|g^{i}|\leq\varepsilon^{5/8}\,\,\mbox{ for }i=2,\ldots,m,

in B1/2B_{1/2}. Since |σ|ε2(n+2)|\sigma|\leq\varepsilon^{2(n+2)}, by Remark 6.7 the function |G|+Cε2(n+2)|y|1a|G|+C\varepsilon^{2(n+2)}|y|^{1-a} is a subsolution and in order to apply again [18, Lemma 4.3], we need to check that

|G|(X¯)+Cε2(n+2)|y¯|1aU(X¯cεen)|G|(\bar{X})+C\varepsilon^{2(n+2)}|\bar{y}|^{1-a}\leq U(\bar{X}-c\varepsilon e_{n})

for some c>0c>0 universal. Since g1(X¯)U(X¯εen)g^{1}(\overline{X})\leq U(\overline{X}-\varepsilon e_{n}), we get

(7.12) g1(X¯)U(X¯)U(X¯εen)U(X¯)=tU(X¯λen)εcε,λ(0,ε)g^{1}(\overline{X})-U(\overline{X})\leq U(\overline{X}-\varepsilon e_{n})-U(\overline{X})=-\partial_{t}U(\overline{X}-\lambda e_{n})\varepsilon\leq-c\varepsilon,\,\lambda\in(0,\varepsilon)

and

|G|(X¯)+Cε2(n2)|y¯|1aU(X¯)g1(X¯)+2Cε5/4U(X¯)c2ε.|G|(\overline{X})+C\varepsilon^{2(n-2)}|\bar{y}|^{1-a}-U(\overline{X})\leq g^{1}(\overline{X})+2C\varepsilon^{5/4}-U(\overline{X})\leq-\frac{c}{2}\varepsilon.

The desired bound follows arguing as in (7.12). ∎

We are now ready to sketch the proof of the Harnack inequality.

Proof of Theorem 7.3.

Without loss of generality, let us assume a0=1a_{0}=-1 and b0=1b_{0}=1. Also, up to rescaling, we can take r=1r=1 and so 2εε¯2\varepsilon\leq\overline{\varepsilon}. Moreover, we denote with ε0\varepsilon_{0} and δ\delta the universal constants in Lemma 7.5, and choose ε¯=ε0\bar{\varepsilon}=\varepsilon_{0}, with εε¯,σε2(n+2)\varepsilon\leq\bar{\varepsilon},\sigma\leq\varepsilon^{2(n+2)}.
Now, we distinguish two cases depending on the position of Br(X0)B_{r}(X_{0}).

Case 1: If dist(X0,{xn=ε,y=0})δ/2\mathrm{dist}(X_{0},\{x_{n}=-\varepsilon,y=0\})\leq\delta/2 we aim to apply Lemma 7.5. Assume that for X¯=1/4en\overline{X}=1/4e_{n}

g1(X¯)U(X¯).g^{1}(\overline{X})\leq U(\overline{X}).

Since,

g1|G|U(X+εen)in B1/2(εen)B1(X0),g^{1}\leq|G|\leq U(X+\varepsilon e_{n})\quad\text{in $B_{1/2}(-\varepsilon e_{n})\subset B_{1}(X_{0})$,}

and for ε\varepsilon small enough, it holds X¯B1/8((ε+1/4)en)\overline{X}\in B_{1/8}((-\varepsilon+1/4)e_{n}), by (7.7) we can apply Lemma 7.5 in B1/2(εen)B_{1/2}(-\varepsilon e_{n}) and deduce that

g1|G|U(X+(1η)εen)in Bδ(εen).g^{1}\leq|G|\leq U(X+(1-\eta)\varepsilon e_{n})\quad\text{in $B_{\delta}(-\varepsilon e_{n})$.}

Finally, the improvement follows by choosing η<δ/2\eta<\delta/2, which implies that Bη(X0)Bδ(εen)B_{\eta}(X_{0})\subset B_{\delta}(-\varepsilon e_{n}). If instead g1(X¯)U(X¯)g^{1}(\overline{X})\geq U(\overline{X}), we can proceed analogously as in the scalar result [12, Theorem 6.1].

Case 2. If dist(X0,{xn=ε,y=0})>δ/2\mathrm{dist}(X_{0},\{x_{n}=-\varepsilon,y=0\})>\delta/2, then g1>0g^{1}>0 and LaL_{a}-superharmonic in B1(X0){g1>0}B_{1}(X_{0})\cap\{g^{1}>0\}. Let v1v^{1} be the LaL_{a}-harmonic replacement of g1g^{1} in B7/8(X0)B_{7/8}(X_{0}). Exploiting Lemma 3.3 and the choice σε2(n+2)\sigma\leq\varepsilon^{2(n+2)} we get

g1v1L(B1/2(X0))Cε2.\|{g^{1}-v^{1}}\|_{L^{\infty}(B_{1/2}(X_{0}))}\leq C\varepsilon^{2}.

Thus,

U(Xεen)Cε2v1U(X+εen)+Cε2in B1/2(X0),U(X-\varepsilon e_{n})-C\varepsilon^{2}\leq v^{1}\leq U(X+\varepsilon e_{n})+C\varepsilon^{2}\quad\mbox{in }B_{1/2}(X_{0}),

which implies that U(X+εa0en)v1U(X+εb0en)U(X+\varepsilon a_{0}e_{n})\leq v^{1}\leq U(X+\varepsilon b_{0}e_{n}) in B1/2(X0)B_{1/2}(X_{0}) for some a0<b0a_{0}<b_{0}.
Then, by applying directly [18, Lemma 4.3] (or [12, Theorem 6.1] for s=1/2s=1/2), as in this case we only need that v1v^{1} is a positive LaL_{a}-harmonic function in B1+(v1)B_{1}^{+}(v^{1}). Thus, going back to g1g^{1}, the conclusion

(7.13) U(X+εa1en)g1U(X+εb1en) in Bη(X0),U(X+\varepsilon a_{1}e_{n})\leq g^{1}\leq U(X+\varepsilon b_{1}e_{n})\quad\text{ in $B_{\eta}(X_{0})$,}

does hold for η\eta small. On the other hand, reasoning as in Lemma 7.2-(i) we have in the same ball,

|G|U(X+εb1en)+Cε5/8U(X+εen)U(X+b¯1εen),|G|\leq U(X+\varepsilon b_{1}e_{n})+C\varepsilon^{5/8}U(X+\varepsilon e_{n})\leq U(X+\bar{b}_{1}\varepsilon e_{n}),

and our claim is proved. ∎

7.3. The improvement of flatness lemma

We can finally conclude the Section with an improvement of flatness lemma, from which the main result of C1,αC^{1,\alpha} regularity of a flat free boundary follows by standard arguments (see for example [20]). Lastly, we give a straightforward proof of Theorem 1.2 by applying the results for flat-solutions of (6.8).
In view of Lemma 7.2, the flatness can be expressed as in (7.14)-(7.15).

Lemma 7.6.

Let α=1\alpha=1 and GG be a viscosity solution of (6.8) in B1B_{1} such that

𝒥(G,B1)𝒥(G~,B1)+σ,\mathcal{J}(G,B_{1})\leq\mathcal{J}(\tilde{G},B_{1})+\sigma,

for every G~GH01,a(B1;m)\tilde{G}-G\in H^{1,a}_{0}(B_{1};\mathbb{R}^{m}), with σε2(n+2)\sigma\leq\varepsilon^{2(n+2)}. Suppose that 0F(G)0\in F(G) and

(7.14) U(Xεen)g1|G|U(X+εen)in B1,U(X-\varepsilon e_{n})\leq g^{1}\leq|G|\leq U(X+\varepsilon e_{n})\quad\text{in $B_{1}$},

with

(7.15) |Gg1f1|ε3/4in B1.|G-g^{1}f^{1}|\leq\varepsilon^{3/4}\quad\text{in $B_{1}$}.

Then, there exists a universal ρ0>0\rho_{0}>0 such that if ρ(0,ρ0]\rho\in(0,\rho_{0}] and ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}] for some ε0=ε0(ρ)\varepsilon_{0}=\varepsilon_{0}(\rho), then for unit vectors νn\nu\in\mathbb{R}^{n} and fmf\in\mathbb{R}^{m},

(7.16) U(x,νε2ρ,y)G,f|G|U(x,ν+ε2ρ,y)in Bρ,U\left(\langle x,\nu\rangle-\frac{\varepsilon}{2}\rho,y\right)\leq\langle G,f\rangle\leq|G|\leq U\left(\langle x,\nu\rangle+\frac{\varepsilon}{2}\rho,y\right)\quad\text{in $B_{\rho}$},

and

(7.17) |GG,ff|(ε2)3/4ρsin Bρ,|G-\langle G,f\rangle f|\leq\left(\frac{\varepsilon}{2}\right)^{3/4}\rho^{s}\quad\text{in $B_{\rho}$},

with |νen|,|ff1|Cε|\nu-e_{n}|,|f-f^{1}|\leq C\varepsilon, for some universal constant C>0.C>0.

Proof.

Following the strategies of [20, 19, 12] , we proceed with a contradiction argument.

Step 1 - Compactness and linearization. Set ρρ0\rho\leq\rho_{0} to be made precise later. Let us suppose there exist εk0,σkεk2(n+2)\varepsilon_{k}\to 0,\sigma_{k}\leq\varepsilon_{k}^{2(n+2)} and a sequence of solutions (Gk)k(G_{k})_{k} of (6.8) such that 0F(Gk)0\in F(G_{k}) and

(7.18) U(Xεken)gk1|Gk|U(X+εken)in B1,U(X-\varepsilon_{k}e_{n})\leq g^{1}_{k}\leq|G_{k}|\leq U(X+\varepsilon_{k}e_{n})\quad\text{in $B_{1}$,}

and

(7.19) |Gkgk1f1|εk3/4in B1,|G_{k}-g^{1}_{k}f^{1}|\leq\varepsilon_{k}^{3/4}\quad\text{in $B_{1},$}

but either of the conclusions (7.16) or (7.17) does not hold. Denote with g~k1\tilde{g}^{1}_{k} and |G|k~\widetilde{|G|_{k}} the εk\varepsilon_{k}-domain variations associated to gk1g^{1}_{k} and |Gk||G_{k}| respectively. In view of (7.18)-(7.19), we can apply Corollary 7.4 and Ascoli-Arzelà theorem to conclude that, up to a subsequence, the sets

ak:={(X,g~k1(X)):XB1εkP}andAk:={(X,|Gk|~(X)):XB1εkP},a_{k}:=\left\{(X,\tilde{g}^{1}_{k}(X))\colon X\in B_{1-\varepsilon_{k}}\setminus P\right\}\quad\text{and}\quad A_{k}:=\left\{(X,\widetilde{|G_{k}|}(X))\colon X\in B_{1-\varepsilon_{k}}\setminus P\right\},

converge uniformly (in Hausdorff distance) in B1/2PB_{1/2}\setminus P to the graphs

a:={(X,g~1(X)):XB1/2P}andA:={(X,|G|~(X)):XB1/2P},a_{\infty}:=\left\{(X,\tilde{g}^{1}_{\infty}(X))\colon X\in B_{1/2}\setminus P\right\}\quad\text{and}\quad A_{\infty}:=\left\{(X,\widetilde{|G_{\infty}|}(X))\colon X\in B_{1/2}\setminus P\right\},

with g~1\tilde{g}^{1}_{\infty} and |G|~\widetilde{|G_{\infty}|} Hölder continuous functions in B1/2B_{1/2}. Moreover,

(7.20) |G|~g~1in B1/2.\widetilde{|G_{\infty}|}\equiv\tilde{g}_{\infty}^{1}\quad\text{in $B_{1/2}.$}

Since gk1g^{1}_{k} is a sequence of supersolutions to the scalar one-phase problem (6.1), while |Gk|+Cσk|y|1a|G_{k}|+C\sigma_{k}|y|^{1-a} is a sequence of subsolutions to the same problem, we conclude by the arguments in [18, Lemma 5.5.] (see also the proof of Step 2 of [12, Lemma 7.4.] for the case s=1/2s=1/2) that |G|~g~1\widetilde{|G_{\infty}|}\equiv\tilde{g}_{\infty}^{1} satisfies (in the viscosity sense) the linearized problem

(7.21) {La(Unw)=0in B1P,|rw|=0on B1L,\begin{cases}L_{a}(U_{n}w)=0&\text{in $B_{1}\setminus P,$}\\ |\nabla_{r}w|=0&\text{on $B_{1}\cap L$},\end{cases}

where

|rw|(X0)=lim(xn,y)(0,0)w(x0,xn,y)w(x0,0,0)r,r=|(xn,y)|,X0=(x0,0,0).|\nabla_{r}w|(X_{0})=\lim_{(x_{n},y)\to(0,0)}\frac{w(x_{0}^{\prime},x_{n},y)-w(x_{0}^{\prime},0,0)}{r},\quad r=|(x_{n},y)|,\,X_{0}=(x_{0}^{\prime},0,0).

In particular, since (g~k1)k(\tilde{g}^{1}_{k})_{k} and (|Gk|~)k(\widetilde{|{G}_{k}|})_{k} are uniformly bounded in B1B_{1}, we get a uniform bound on g~1|G|~\tilde{g}^{1}_{\infty}\equiv\widetilde{|{G}_{\infty}|}, hence by [18, Theorem 6.2.] (see [12, Lemma 4.2] for s=1/2s=1/2), since g~1(0)=0,\tilde{g}^{1}_{\infty}(0)=0, we deduce that for C0C_{0} universal,

(7.22) |g~1(X)ν,x|C0ρ1+γin B2ρ,\big{\lvert}{\tilde{g}^{1}_{\infty}(X)-\langle\nu^{\prime},x^{\prime}\rangle}\big{\rvert}\leq C_{0}\rho^{1+\gamma}\quad\text{in $B_{2\rho}$},

for some vector νn1\nu^{\prime}\in\mathbb{R}^{n-1} and γ>0\gamma>0 universal constant. Details are omitted as we reduced to the scalar case, hence the arguments of [18, Theorem 5.1.] (see [12, Theorem 7.1.] for s=1/2s=1/2) apply directly.

Step 2 - Improvement of flatness. In view of (7.22) (see also [18, Corollary 3.3.] and [12, Theorem 8.2.]), for ρ<1/(8C0)\rho<1/(8C_{0}) small enough, we get

ν,x18ρg~1(X)ν,x+18ρin B2ρ\langle\nu^{\prime},x^{\prime}\rangle-\frac{1}{8}\rho\leq\tilde{g}^{1}_{\infty}(X)\leq\langle\nu^{\prime},x^{\prime}\rangle+\frac{1}{8}\rho\quad\text{in $B_{2\rho}$}

and, for kk sufficiently large, we deduce from the uniform convergence of aka_{k} to aa_{\infty} and of AkA_{k} to AA_{\infty} that

(7.23) ν,x14ρg~k1(X)|Gk|~(X)ν,x+14ρin B2ρP.\langle\nu^{\prime},x^{\prime}\rangle-\frac{1}{4}\rho\leq\tilde{g}^{1}_{k}(X)\leq\widetilde{|G_{k}|}(X)\leq\langle\nu^{\prime},x^{\prime}\rangle+\frac{1}{4}\rho\quad\text{in $B_{2\rho}\setminus P$}.

The argument of the proof of [12, Lemma 7.2.] then gives

(7.24) U(x,νεk4ρ,y)gk1|Gk|U(x,ν+εk4ρ,y)in B32ρ,U\left(\langle x,\nu\rangle-\frac{\varepsilon_{k}}{4}\rho,y\right)\leq g^{1}_{k}\leq|G_{k}|\leq U\left(\langle x,\nu\rangle+\frac{\varepsilon_{k}}{4}\rho,y\right)\quad\text{in $B_{\frac{3}{2}\rho}$},

for a unit vector ν\nu with |νen|Cεk.|\nu-e_{n}|\leq C\varepsilon_{k}.

On the other hand, by (7.18)-(7.19) we conclude that, up to a subsequence, gki/εk3/4gig^{i}_{k}/\varepsilon_{k}^{3/4}\to g^{i}_{*} uniformly to some gig^{i}_{*} such that

{Lagi=0in B1/2Pgi=0on PB1/2.\begin{cases}L_{a}g^{i}_{*}=0&\mbox{in }B_{1/2}\setminus P\\ g^{i}_{*}=0&\mbox{on }P\cap B_{1/2}.\end{cases}

for every i=2,,mi=2,\ldots,m. Thus, for k>0k>0 sufficiently large

|gkiMiUεk3/4|Cεk3/4ργUin B32ρ,|g^{i}_{k}-M_{i}U\varepsilon_{k}^{3/4}|\leq C\varepsilon_{k}^{3/4}\rho^{\gamma}U\quad\text{in $B_{\frac{3}{2}\rho}$},

with γ,C>0\gamma,C>0 universal and |Mi|M|M_{i}|\leq M universal. From the properties of the function UU and (7.18), we deduce that

(7.25) |gkiMigk1εk3/4|Cεk3/4(ργU+εks)(εk8)3/4ρsin B32ρ,|g^{i}_{k}-M_{i}g^{1}_{k}\varepsilon_{k}^{3/4}|\leq C\varepsilon_{k}^{3/4}(\rho^{\gamma}U+\varepsilon_{k}^{s})\leq\left(\frac{\varepsilon_{k}}{8}\right)^{3/4}\rho^{s}\quad\text{in $B_{\frac{3}{2}\rho}$},

by choosing ρρ0\rho\leq\rho_{0} small enough and then kk sufficiently large.

Now, inspired by the construction of Step 2 of [19, Lemma 7.1.], set

ξk1:=f1+εk3/4i1Mifi,f¯k1:=ξk1|ξk1|.\xi_{k}^{1}:=f^{1}+\varepsilon_{k}^{3/4}\sum_{i\neq 1}M_{i}f^{i},\quad\bar{f}^{1}_{k}:=\frac{\xi^{1}_{k}}{|\xi_{k}^{1}|}.

Notice that,

(7.26) f¯k1=ξk1+O(εk3/2).\bar{f}^{1}_{k}=\xi_{k}^{1}+O(\varepsilon_{k}^{3/2}).

We claim that

(7.27) U(x,νεk2ρ,y)Gk,f¯k1|Gk|U(x,ν+εk2ρ,y)in Bρ,U\left(\langle x,\nu\rangle-\frac{\varepsilon_{k}}{2}\rho,y\right)\leq\langle G_{k},\bar{f}^{1}_{k}\rangle\leq|G_{k}|\leq U\left(\langle x,\nu\rangle+\frac{\varepsilon_{k}}{2}\rho,y\right)\quad\text{in $B_{\rho}$},

and

(7.28) |GkGk,f¯k1f¯k1|(εk2)3/4ρsin Bρ,|G_{k}-\langle G_{k},\bar{f}^{1}_{k}\rangle\bar{f}^{1}_{k}|\leq\left(\frac{\varepsilon_{k}}{2}\right)^{3/4}\rho^{s}\quad\text{in $B_{\rho}$},

thus reaching a contradiction. First, the upper bound in (7.27) is a direct consequence of (7.24). For the lower bound, we observe that by (7.18), (7.19) and (7.26),

|GkUf¯k1|0,as k,|G_{k}-U\bar{f}^{1}_{k}|\to 0,\quad\text{as $k\to\infty$},

while

(7.29) |Gk|0in 1/2{xnεk}.|G_{k}|\equiv 0\quad\text{in $\mathcal{B}_{1/2}\cap\{x_{n}\leq-\varepsilon_{k}\}$}.

As pointed out in Remark 7.1, we know that

(7.30) Gk,f¯k1>0in B12+(Gk),\langle G_{k},\bar{f}^{1}_{k}\rangle>0\quad\text{in $B^{+}_{\frac{1}{2}}(G_{k})$},

for kk sufficiently large. Moreover, by the definition of f¯k1\bar{f}^{1}_{k}, (7.26) and (7.30),

(7.31) Gk,f¯k1(U(x,νεk4ρ,y)Cεk3/2)+in B32ρ.\langle G_{k},\bar{f}^{1}_{k}\rangle\geq\left(U\left(\langle x,\nu\rangle-\frac{\varepsilon_{k}}{4}\rho,y\right)-C\varepsilon_{k}^{3/2}\right)^{+}\quad\text{in $B_{\frac{3}{2}\rho}$}.

Set

h(X):=(U(x,νεk4ρ,y)Cεk3/2)+h(X):=\left(U\left(\langle x,\nu\rangle-\frac{\varepsilon_{k}}{4}\rho,y\right)-C\varepsilon_{k}^{3/2}\right)^{+}

and call with HH the LaL_{a}-harmonic function in B32ρ{(x,0),νεk4ρ}B_{\frac{3}{2}\rho}\setminus\{\langle(x,0),\nu\rangle\leq\frac{\varepsilon_{k}}{4}\rho\} satisfying

H=hon B32ρ,H=0on 32ρ{x,ν=εk4ρ}.H=h\quad\text{on $\partial B_{\frac{3}{2}\rho}$},\quad H=0\quad\text{on $\mathcal{B}_{\frac{3}{2}\rho}\cap\{\langle x,\nu\rangle=\frac{\varepsilon_{k}}{4}\rho\}$.}

Then, by (7.24)-(7.30)-(7.31) and the comparison principle, we conclude that

Gk,f¯k1Hin B32ρ.\langle G_{k},\bar{f}^{1}_{k}\rangle\geq H\quad\text{in $B_{\frac{3}{2}\rho}$}.

On the other hand, by applying the Boundary Harnack

H(1Cεk3/2)U(x,νεk4ρ,y)on Bρ,H\geq(1-C\varepsilon_{k}^{3/2})U\left(\langle x,\nu\rangle-\frac{\varepsilon_{k}}{4}\rho,y\right)\quad\text{on $B_{\rho},$}

for C>0C>0 universal, from which the required lower bound follows for kk sufficiently large.

We are left with the proof of (7.28). By the definition of f¯k1\bar{f}^{1}_{k} and (7.26), it is sufficient to show that

|GkGk,ξk1ξk1|(εk4)3/4ρsin Bρ.|G_{k}-\langle G_{k},\xi^{1}_{k}\rangle\xi^{1}_{k}|\leq\left(\frac{\varepsilon_{k}}{4}\right)^{3/4}\rho^{s}\quad\text{in $B_{\rho}$}.

Set G¯k:=GkGk,ξk1ξk1\bar{G}_{k}:=G_{k}-\langle G_{k},\xi^{1}_{k}\rangle\xi^{1}_{k}, then by (7.19) we immediately get

|g¯k1|=ε3/4|i1Migki|Cεk3/2.|\bar{g}_{k}^{1}|=\varepsilon^{3/4}\bigg{|}\sum_{i\neq 1}M_{i}g^{i}_{k}\bigg{|}\leq C\varepsilon_{k}^{3/2}.

Instead, for the remaining components, by using (7.25) we have

|g¯ki|=|gkiεk3/4Migk1εk3/2Mij1Mjgkj|(εk8)3/4ρs+Cεk9/4,|\bar{g}^{i}_{k}|=\bigg{|}g_{k}^{i}-\varepsilon_{k}^{3/4}M_{i}g^{1}_{k}-\varepsilon_{k}^{3/2}M_{i}\sum_{j\neq 1}M_{j}g^{j}_{k}\bigg{|}\leq\left(\frac{\varepsilon_{k}}{8}\right)^{3/4}\rho^{s}+C\varepsilon_{k}^{9/4},

and the desired bound follows for kk large enough. ∎

Finally, we are able to conclude the Section with the proof of the Theorem 1.2.

Proof of Theorem 1.2.

The statements about nn^{*} and the fact that {|G|>0}{y=0}\{|G|>0\}\cap\{y=0\} has locally finite perimeter follow exactly as in the scalar case (see [14, Section 5] and [21, Theorem 1.2.]). Moreover, the estimate on the Hausdorff dimension of Sing(F(G))\mathrm{Sing}(F(G)) follows combining Theorem 1.1, Proposition 5.7, Definition 5.10, Proposition 5.12.
Lastly, let us deal with the regularity result for Reg(F(G))\mathrm{Reg}(F(G)). First, by Proposition 6.5, the vector of normalized eigenfunctions associated to a shape optimizer of (1.1) is a viscosity solution of (6.10).
Fix λ=(λ1s(Ω),,λms(Ω)\lambda=(\lambda_{1}^{s}(\Omega),\dots,\lambda_{m}^{s}(\Omega), then by Remark 6.6, for every X0Reg(F(G)),r>0X_{0}\in\mathrm{Reg}(F(G)),r>0 the function

G2(X)=Γ(1+s)ΛrsG(X0+rX)G_{2}(X)=\frac{\Gamma(1+s)}{\sqrt{\Lambda}r^{s}}G(X_{0}+rX)

is a viscosity solution to (6.8) with α=1,σ=rsλ,0F(G2)\alpha=1,\sigma=r^{s}\lambda,0\in F(G_{2}). Moreover, by Proposition 5.7 and Definition 5.10, given ε>0\varepsilon>0 there exists r>0r>0 such that G2G_{2} is ε\varepsilon-flat solution of (6.8) in B1B_{1}, in some directions (f,ν)(f,\nu) with α=1\alpha=1 and σε2(n+2)\sigma\leq\varepsilon^{2(n+2)}.
Then, by a standard iteration of Lemma 3.3., we get that Reg(F(G2))C1,a\mathrm{Reg}(F(G_{2}))\in C^{1,a} in B1/2B_{1/2}. ∎

References

  • [1] M. Allen and A. Petrosyan. A two-phase problem with a lower-dimensional free boundary. Interfaces Free Bound., 14(3):307–342, 2012.
  • [2] H. W. Alt, L. Caffarelli, and A. Friedman. Variational problems with two phases and their free boundaries. Trans. Amer. Math. Soc., 282(2):431–461, 1984.
  • [3] J. F. Bonder, A. Ritorto, and A. M. Salort. A class of shape optimization problems for some nonlocal operators. Advances in Calculus of Variations, 11(4):373 – 386, 01 Oct. 2018.
  • [4] L. Brasco, E. Lindgren, and E. Parini. The fractional cheeger problem. Interfaces Free Bound., 16(3):419–458, 2014.
  • [5] L. Brasco and E. Parini. The second eigenvalue of the fractional p-laplacian. Advances in Calculus of Variations, 9(4):323–355, 2016.
  • [6] T. Briançon and J. Lamboley. Regularity of the optimal shape for the first eigenvalue of the laplacian with volume and inclusion constraints. Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 26(4):1149–1163, 2009.
  • [7] D. Bucur. Minimization of the k-th eigenvalue of the dirichlet laplacian. Archive for Rational Mechanics and Analysis, 206(3):1073–1083, Dec 2012.
  • [8] D. Bucur, D. Mazzoleni, A. Pratelli, and B. Velichkov. Lipschitz regularity of the eigenfunctions on optimal domains. Archive for Rational Mechanics and Analysis, 216(1):117–151, Apr 2015.
  • [9] G. Buttazzo and G. Dal Maso. An existence result for a class of shape optimization problems. Archive for Rational Mechanics and Analysis, 122(2):183–195, Jun 1993.
  • [10] L. Caffarelli and L. Silvestre. An extension problem related to the fractional Laplacian. Comm. Partial Differential Equations, 32(7-9):1245–1260, 2007.
  • [11] L. A. Caffarelli, J.-M. Roquejoffre, and Y. Sire. Variational problems for free boundaries for the fractional Laplacian. J. Eur. Math. Soc. (JEMS), 12(5):1151–1179, 2010.
  • [12] D. De Silva and J. M. Roquejoffre. Regularity in a one-phase free boundary problem for the fractional Laplacian. Ann. Inst. H. Poincaré Anal. Non Linéaire, 29(3):335–367, 2012.
  • [13] D. De Silva and O. Savin. C2,αC^{2,\alpha} regularity of flat free boundaries for the thin one-phase problem. Journal of Differential Equations, 253(8):2420 – 2459, 2012.
  • [14] D. De Silva and O. Savin. Regularity of lipschitz free boundaries for the thin one-phase problem. Journal of the European Mathematical Society, 017(6):1293–1326, 2015.
  • [15] D. De Silva and O. Savin. CC^{\infty} regularity of certain thin free boundaries. Indiana University Mathematics Journal, 64(5):1575–1608, 2015.
  • [16] D. De Silva and O. Savin. Almost minimizers of the one-phase free boundary problem. Communications in Partial Differential Equations, 45(8):913–930, 2020.
  • [17] D. De Silva and O. Savin. Thin one-phase almost minimizers. Nonlinear Analysis, 193:111507, 2020. Nonlocal and Fractional Phenomena.
  • [18] D. De Silva, O. Savin, and Y. Sire. A one-phase problem for the fractional Laplacian: regularity of flat free boundaries. Bull. Inst. Math. Acad. Sin. (N.S.), 9(1):111–145, 2014.
  • [19] D. De Silva and G. Tortone. A vectorial problem with thin free boundary. arXiv e-prints, page arXiv:2010.05782, Oct. 2020.
  • [20] D. De Silva and G. Tortone. Improvement of flatness for vector valued free boundary problems. Mathematics in Engineering, 2(mine-02-04-027):598, 2020.
  • [21] M. Engelstein, A. Kauranen, M. Prats, G. Sakellaris, and Y. Sire. Minimizers for the thin one-phase free boundary problem. arXiv e-prints, page arXiv:1907.11604, July 2019.
  • [22] E. Fabes, C. Kenig, and R. Serapioni. The local regularity of solutions of degenerate elliptic equations. Comm. Partial Differential Equations, 7(1):77–116, 1982.
  • [23] E. B. Fabes, C. E. Kenig, and D. Jerison. Boundary behavior of solutions to degenerate elliptic equations. In Conference on harmonic analysis in honor of Antoni Zygmund, Vol. I, II (Chicago, Ill., 1981), Wadsworth Math. Ser., pages 577–589. Wadsworth, Belmont, CA, 1983.
  • [24] X. Fernández-Real and X. Ros-Oton. Stable cones in the thin one-phase problem. arXiv e-prints, page arXiv:2009.11626, Sept. 2020.
  • [25] R. L. Frank and L. Geisinger. Refined semiclassical asymptotics for fractional powers of the laplace operator. Journal für die reine und angewandte Mathematik (Crelles Journal), 2016(712):1–37, 2016.
  • [26] D. Kriventsov and F.-H. Lin. Regularity for shape optimizers: the nondegenerate case. Comm. Pure Appl. Math., 71(8):1535–1596, 2018.
  • [27] D. Kriventsov and F.-H. Lin. Regularity for shape optimizers: the degenerate case. Comm. Pure Appl. Math., 72(8):1678–1721, 2019.
  • [28] M. Kwaśnicki. Eigenvalues of the fractional laplace operator in the interval. Journal of Functional Analysis, 262(5):2379–2402, 2012.
  • [29] E. Lindgren and P. Lindqvist. Fractional eigenvalues. Calculus of Variations and Partial Differential Equations, 49(1):795–826, Jan 2014.
  • [30] D. Mazzoleni and A. Pratelli. Existence of minimizers for spectral problems. Journal de Mathématiques Pures et Appliquées, 100(3):433–453, 2013.
  • [31] D. Mazzoleni, S. Terracini, and B. Velichkov. Regularity of the optimal sets for some spectral functionals. Geom. Funct. Anal., 27(2):373–426, 2017.
  • [32] D. Mazzoleni, S. Terracini, and B. Velichkov. Regularity of the free boundary for the vectorial Bernoulli problem. Analysis & PDE, 13(3):741 – 764, 2020.
  • [33] D. Mazzoleni, B. Trey, and B. Velichkov. Regularity of the optimal sets for the second Dirichlet eigenvalue. arXiv e-prints, page arXiv:2010.00441, Oct. 2020.
  • [34] E. Parini and A. Salort. Compactness and dichotomy in nonlocal shape optimization. Mathematische Nachrichten, 293(11):2208–2232, 2020.
  • [35] Y. Sire, S. Terracini, and G. Tortone. On the nodal set of solutions to degenerate or singular elliptic equations with an application to s-harmonic functions. Journal de Mathématiques Pures et Appliquées, 2020.
  • [36] S. Terracini, G. Tortone, and S. Vita. On s-harmonic functions on cones. Anal. PDE, 11(7):1653–1691, 2018.
  • [37] S. Terracini and S. Vita. On the asymptotic growth of positive solutions to a nonlocal elliptic blow-up system involving strong competition. Ann. Inst. H. Poincaré Anal. Non Linéaire, 35(3):831–858, 2018.
  • [38] G. Tortone and A. Zilio. Regularity results for segregated configurations involving fractional laplacian. Nonlinear Analysis, 193:111532, 2020. Nonlocal and Fractional Phenomena.
  • [39] B. Trey. Lipschitz continuity of the eigenfunctions on optimal sets for functionals with variable coefficients. ESAIM: COCV, 26:89, 2020.
  • [40] B. Trey. Regularity of optimal sets for some functional involving eigenvalues of an operator in divergence form. Annales de l’Institut Henri Poincaré C, Analyse non linéaire, 2020.
  • [41] K. Wang and J. Wei. On the uniqueness of solutions of a nonlocal elliptic system. Mathematische Annalen, 365(1-2):105–153, 2016.
  • [42] G. Weiss. A homogeneity improvement approach to the obstacle problem. Invent. Math., 138(1):23–50, 1999.