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

Optimal one-dimensional structures for the principal eigenvalue of two-dimensional domains

Giuseppe Buttazzo and Francesco Paolo Maiale
Abstract

A shape optimization problem arising from the optimal reinforcement of a membrane by means of one-dimensional stiffeners or from the fastest cooling of a two-dimensional object by means of “conducting wires” is considered. The criterion we consider is the maximization of the first eigenvalue and the admissible classes of choices are the one of one-dimensional sets with prescribed total length, or the one where the constraint of being connected (or with an a priori bounded number of connected components) is added. The corresponding relaxed problems and the related existence results are described.


Keywords: Optimal reinforcement, eigenvalues of the Laplacian, stiffeners, fastest cooling.

2010 Mathematics Subject Classification: 49J45, 35R35, 35J25, 49Q10.

1 Introduction

The problem of finding the vibration modes of an elastic membrane Ω2\Omega\subset\mathbb{R}^{2}, fixed at its boundary Ω\partial\Omega, is known to reduce to the PDE

{Δu=λuin Ω,u=0on Ω.\begin{cases}-\Delta u=\lambda u&\text{in }\Omega,\\ u=0&\text{on }\partial\Omega.\end{cases}

The eigenvalues λ\lambda for which the PDE above has nonzero solutions are all strictly positive with no finite limit, hence they can be ordered as

0<λ1λ2λk+.0<\lambda_{1}\leq\lambda_{2}\leq\dots\leq\lambda_{k}\leq\dots\to+\infty.

We are interested in the behavior of the first eigenvalue, which can be also characterized via the variational problem

min{Ω|u|2𝑑xΩ|u|2𝑑x:uH01(Ω),u0}.\min\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx}{\int_{\Omega}|u|^{2}\,dx}\ :\ u\in H_{0}^{1}(\Omega),\ u\neq 0\right\}.

Our goal is to see how the value λ1\lambda_{1} above modifies when we attach to the membrane a one-dimensional stiffener SS, modeled by a one-dimensional rectifiable set SΩS\subset\Omega. In this case the first eigenvalue depends on SS and is given by

λ1(S)=inf{Ω|u|2𝑑x+mS|τu|2𝑑1Ω|u|2𝑑x:uCc(Ω),u0},\lambda_{1}(S)=\inf\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{S}|\nabla_{\tau}u|^{2}\,d\mathcal{H}^{1}}{\int_{\Omega}|u|^{2}\,dx}\ :\ u\in C_{c}^{\infty}(\Omega),\ u\neq 0\right\}, (1.1)

where τ\nabla_{\tau} is the tangential derivative and the parameter mm indicates the stiffness coefficient of the material of which SS is made of.

A similar problem arises in the heat diffusion when a two-dimensional heat conductor, with zero temperature at the boundary and initial temperature u0u_{0}, has to be cooled as fast as possible by adding one-dimensional strongly conducting wires SS. The corresponding second order operator in presence of the structure SS is given in the weak form by

𝒜Su,ϕ=Ωuϕdx+mSτuτϕd1,\langle\mathcal{A}_{S}u,\phi\rangle=\int_{\Omega}\nabla u\nabla\phi\,dx+m\int_{S}\nabla_{\tau}u\nabla_{\tau}\phi\,d\mathcal{H}^{1},

where uu and ϕ\phi vary in the Sobolev space H01(Ω)H1(S)H^{1}_{0}(\Omega)\cap H^{1}(S). By the Fourier analysis we may write the solution of the heat equation

{tu+𝒜Su=0in ]0,T[×Ωu=0on ]0,T[×Ωu=u0on Ω for t=0\begin{cases}\partial_{t}u+\mathcal{A}_{S}u=0&\text{in }]0,T[\times\Omega\\ u=0&\text{on }]0,T[\times\partial\Omega\\ u=u_{0}&\text{on $\Omega$ for }t=0\end{cases}

as

u(t,x)=k1ckuk(x)etλk(S),ck=Ωu0uk𝑑xu(t,x)=\sum_{k\geq 1}c_{k}u_{k}(x)e^{-t\lambda_{k}(S)},\qquad c_{k}=\int_{\Omega}u_{0}u_{k}\,dx

where λk(S)\lambda_{k}(S) are the eigenvalues of the operator 𝒜S\mathcal{A}_{S} and uku_{k} the corresponding eigenfunctions (normalized with unitary L2L^{2} norm). The fastest cooling then reduces to searching the structure SS providing the maximal first eigenvalue among the class of admissible choices for SS.

In the present paper we consider the shape optimization problem related to the functional λ1(S)\lambda_{1}(S) defined in (1.1) on the following two classes of admissible choices for the stiffener SS, where (S)\mathcal{L}(S) denotes the length of SS:

𝒜L={SΩ¯,S rectifiable, (S)L};𝒜Lc={SΩ¯,S rectifiable, S connected, (S)L}.\begin{split}&\mathcal{A}_{L}=\big{\{}S\subset\bar{\Omega},\ S\text{ rectifiable, }\mathcal{L}(S)\leq L\big{\}};\\ &\mathcal{A}^{c}_{L}=\big{\{}S\subset\bar{\Omega},\ S\text{ rectifiable, }S\text{ connected, }\mathcal{L}(S)\leq L\big{\}}.\end{split}

Similarly, we could consider the admissible class of stiffeners having at most NN connected components

𝒜Lc,N={SΩ¯,S rectifiable, S has most N connected components, (S)L}.\mathcal{A}^{c,N}_{L}=\big{\{}S\subset\bar{\Omega},\ S\text{ rectifiable, }S\text{ has most $N$ connected components, }\mathcal{L}(S)\leq L\big{\}}.

We do not consider this last situation since there are no essential differences between the cases N>1N>1 and N=1N=1. For a general presentation of shape optimization problem we refer to the books [6] and [19]

In Section 2 we give the precise formulation of the two optimization problems involving the admissible classes 𝒜L\mathcal{A}_{L} and 𝒜Lc\mathcal{A}^{c}_{L} and their corresponding relaxed formulations. We will show that the relaxed problems admit a solution, which will be in both cases a measure μ\mu supported in Ω\Omega in the first case and on a rectifiable set SS in the second one. Our main results are that these measures do not have singular parts; more precisely, in the case 𝒜L\mathcal{A}_{L} it is a function θLp(Ω)\theta\in L^{p}(\Omega), while in the case 𝒜Lc\mathcal{A}^{c}_{L} it is a measure of the form θ  S\theta\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S where SS is a suitable connected set and θL1(S)\theta\in L^{1}(S).

Section 3 contains the proofs of the results. In Section 4 we consider the case when Ω\Omega is a disk, in which some explicit calculations can be made for the relaxed optimization problem related to the choice 𝒜L\mathcal{A}_{L} of admissible sets. Section 5 deals with the case 𝒜Lc\mathcal{A}^{c}_{L} in which the admissible sets SS are connected. Finally, in Section 6 we collected some open questions that in our opinion merit some further investigation.

2 Formulation of the problem and main results

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded Lipschitz domain. The two optimization problems we consider are

max{λ1(S):S𝒜L}\max\big{\{}\lambda_{1}(S)\ :\ S\in\mathcal{A}_{L}\big{\}} (2.1)
max{λ1(S):S𝒜Lc}\max\big{\{}\lambda_{1}(S)\ :\ S\in\mathcal{A}^{c}_{L}\big{\}} (2.2)

where λ1(S)\lambda_{1}(S) is defined in (1.1). We now deduce in the two cases the corresponding relaxed problems, obtained by means of the possible limits of admissible SnS_{n}. In the following we use the notation:

  • |A||A| for the (two-dimensional) Lebesgue measure of AA;

  • (S)\mathcal{L}(S) for the length of a one-dimensional set SS;

  • μ\|\mu\| for the mass of a measure μ\mu.

Let (Sn)(S_{n}) be a sequence of admissible stiffeners for problem (2.1); considering the measures μn=1  Sn\mu_{n}=\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S_{n} we have that μnL\|\mu_{n}\|\leq L, hence a subsequence (that we still indicate by μn\mu_{n}) weakly* converges to a suitable measure μ\mu. It is then convenient to define λ1(μ)\lambda_{1}(\mu) for every measure μ\mu on Ω\Omega by setting

λ1(μ)=inf{Ω|u|2𝑑x+m|u|2𝑑μΩ|u|2𝑑x:uCc(Ω),u0}.\lambda_{1}(\mu)=\inf\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\ :\ u\in C^{\infty}_{c}(\Omega),\ u\neq 0\right\}.

Note that, in general the infimum above is not attained on Cc(Ω)C^{\infty}_{c}(\Omega) and minimizing sequences converge, strongly in L2(Ω)L^{2}(\Omega) and weakly in a suitably defined Sobolev space Hμ1H^{1}_{\mu}, to solutions of the relaxed problem

min{Ω|u|2𝑑x+m|μu|2𝑑μΩ|u|2𝑑x:uH01(Ω)Hμ1,u0}.\min\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int|\nabla_{\mu}u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\ :\ u\in H^{1}_{0}(\Omega)\cap H^{1}_{\mu},\ u\neq 0\right\}.

Here μ\nabla_{\mu} represents a kind of tangential gradient that was defined in [4] for every measure μ\mu. In this way, when μ=1  S\mu=\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S the tangential gradient μ\nabla_{\mu} coincides with the usual tangential gradient to SS, so that the definition above of λ1(μ)\lambda_{1}(\mu) reduces to λ1(S)\lambda_{1}(S). The relaxed version of the optimization problem (2.1) then reads

max{λ1(μ):μ𝒜L}\max\big{\{}\lambda_{1}(\mu)\ :\ \mu\in\mathscr{A}_{L}\big{\}} (2.3)

where 𝒜L\mathscr{A}_{L} is the class of nonnegative measures μ\mu on Ω\Omega with μL\|\mu\|\leq L.

Proposition 2.1.

The relaxed optimization problem (2.3) admits a solution.

Proof.

For every fixed uCc(Ω)u\in C^{\infty}_{c}(\Omega) the map

μΩ|u|2𝑑x+m|u|2𝑑μΩ|u|2𝑑x\mu\mapsto\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}

is weakly* continuous. Hence λ1(μ)\lambda_{1}(\mu) is upper semicontinuous for the weak* convergence, being the infimum of continuous functions. The existence result then follows from the fact that, thanks to the bound μL\|\mu\|\leq L, the class 𝒜L\mathscr{A}_{L} is weakly* compact. ∎

Remark 2.2.

Following the theory developed in [4] concerning variational integrals with respect to a general measure, the expression of λ1(μ)\lambda_{1}(\mu) can be equivalently given as

λ1(μ)=inf{Ω|u|2𝑑x+m|μu|2𝑑μΩ|u|2𝑑x:uH01(Ω)Hμ1,u0},\lambda_{1}(\mu)=\inf\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int|\nabla_{\mu}u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\ :\ u\in H^{1}_{0}(\Omega)\cap H^{1}_{\mu},\ u\neq 0\right\},

where the Sobolev space Hμ1H^{1}_{\mu} and the “tangential gradient” μ\nabla_{\mu} are suitably defined. We refer the interested reader to [4], where the precise definitions and all the details are explained. We will see that for our purposes we do not need these fine tools, since we will obtain that optimal measures for problem (2.3) are actually LpL^{p} functions, for which the tangential gradient reduces to the usual gradient and the Sobolev space H01(Ω)Hμ1H^{1}_{0}(\Omega)\cap H^{1}_{\mu} reduces to the usual Sobolev space H01(Ω)H^{1}_{0}(\Omega).

Before stating our main result, we introduce a slightly technical assumption which ensures a bound on the LL^{\infty}-norm of the gradient on the boundary of Ω\Omega.

Definition 2.3 (External Ball Condition).

A subset Ωd\Omega\subset\mathbb{R}^{d} satisfies the uniform external ball condition with radius ρ>0\rho>0 if

x0Ω,y0d:B(y0,ρ)dΩ and x0B(y0,ρ).\forall x_{0}\in\partial\Omega,\ \exists y_{0}\in\mathbb{R}^{d}\ :\ B(y_{0},\rho)\subset\mathbb{R}^{d}\setminus\Omega\text{ and }x_{0}\in\partial B(y_{0},\rho).

We will always require Ω\Omega connected to work with the “unique” eigenfunction which is positive on all Ω\Omega (see [18, Theorem 1.2.5]) and with fixed L2L^{2}-norm. Our main result concerning optimization problem (2.3) is below.

Theorem 2.4.

Let Ω\Omega be a connected subset of 2\mathbb{R}^{2} with Lipschitz boundary satisfying the uniform external ball condition. Then the optimization problem (2.3) admits a solution of the form μ=θdx\mu=\theta\,dx where θ\theta is a function belonging to Lp(Ω)L^{p}(\Omega) for every p<+p<+\infty, equal to zero almost everywhere on the set

{xΩ:|uθ|(x)<uθL(Ω)}\big{\{}x\in\Omega\ :\ |\nabla u_{\theta}|(x)<\|\nabla u_{\theta}\|_{L^{\infty}(\Omega)}\big{\}}

and satisfying the identity

maxμ𝒜Lλ1(μ)=minuH01(Ω){0}{Ω|u|2𝑑x+mLuL(Ω)2Ω|u|2𝑑x}.\max_{\mu\in\mathscr{A}_{L}}\lambda_{1}(\mu)=\min_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|_{L^{\infty}(\Omega)}^{2}}{\int_{\Omega}|u|^{2}\,dx}\right\}. (2.4)

Furthermore,

  1. (i)

    if Ω\Omega is convex we have θL(Ω)\theta\in L^{\infty}(\Omega);

  2. (ii)

    if ΩC2,α\partial\Omega\in C^{2,\alpha}, then there exists β=β(α)(0,1)\beta=\beta(\alpha)\in(0,1) such that θC1,β(Ω¯)\theta\in C^{1,\beta}(\bar{\Omega});

The proof of theorem above is given in Section 3. We now consider the relaxation of the optimization problem (2.2) where the connectedness constraint is imposed. In this case, if (Sn)(S_{n}) is sequence in 𝒜Lc\mathcal{A}_{L}^{c}, the limit of (a subsequence of) μn=1  Sn\mu_{n}=\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S_{n} is still a measure μ\mu supported by a suitable set SS. Since the sequence (Sn)(S_{n}) is compact in the Hausdorff convergence the set SS is closed and connected. In addition, thanks to the Gołab theorem (see [17], and the books [3], [15]), we have μ1  S\mu\geq\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S, hence the set SS verifies (S)L\mathcal{L}(S)\leq L, so that S𝒜LcS\in\mathcal{A}_{L}^{c}. Then, introducing the class

𝒜Lc={μ measure on Ω,μL,sptμ=S connected, μ1  S},\mathscr{A}_{L}^{c}=\big{\{}\mu\hbox{ measure on }\Omega,\ \|\mu\|\leq L,\ \mathrm{spt}\mu=S\hbox{ connected, }\mu\geq\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S\big{\}},

the relaxed version of the optimization problem (2.2) reads

max{λ1(μ):μ𝒜Lc}.\max\big{\{}\lambda_{1}(\mu)\ :\ \mu\in\mathscr{A}^{c}_{L}\big{\}}. (2.5)
Proposition 2.5.

The relaxed optimization problem (2.5) admits a solution.

Proof.

The proof is similar to the one of Proposition 2.1. The map λ1(μ)\lambda_{1}(\mu) is upper semicontinuous for the weak* convergence, and the existence result then follows from the compactness, with respect to the weak* convergence, of the class 𝒜Lc\mathscr{A}_{L}^{c}. This is a consequence of the compactness, for the Hausdorff convergence, of the class of closed and connected sets, and of the Gołab theorem, which gives the inequality μ1  S\mu\geq\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S for a weak* limit μ\mu of a sequence μn=1  Sn\mu_{n}=\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S_{n} with SnS_{n} connected and converging to SS in the Hausdorff sense. ∎

Again, the question if the optimization problem (2.5) admits as a solution a measure that is actually a function on a set SS arises; this would avoid the use of the delicate theory of variational integrals with respect to a general measure and of the related Sobolev spaces. This is indeed the case and our main result concerning optimization problem (2.5) is below.

Theorem 2.6.

The optimization problem (2.5) admits a solution of the form μ=θ1  S\mu=\theta\,\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S, where SS is a closed connected subset of Ω¯\bar{\Omega} with (S)L\mathcal{L}(S)\leq L and θL1(S)\theta\in L^{1}(S) with θ1\theta\geq 1 on SS.

Remark 2.7.

If we introduce the class of measures

𝒜Lc,N={μ measure on Ω,μL,sptμ=S N-connected, μ1  S},\mathscr{A}_{L}^{c,N}=\big{\{}\mu\hbox{ measure on }\Omega,\ \|\mu\|\leq L,\ \mathrm{spt}\mu=S\hbox{ $N$-connected, }\mu\geq\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S\big{\}},

where NN-connected means that is has at most NN connected components, then the proof of Theorem 2.6 easily generalize to the maximization problem

max{λ1(μ):μ𝒜Lc,N}.\max\big{\{}\lambda_{1}(\mu)\ :\ \mu\in\mathscr{A}^{c,\,N}_{L}\big{\}}.

In particular, there exist a solution of the form μN=j=1θj1  Sj\mu_{N}=\sum_{j=1}^{\ell}\theta_{j}\,\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S_{j} where the SjS_{j} are closed connected subsets of Ω¯\bar{\Omega} of total length L\leq L. Moreover, θL1(1jSj)\theta\in L^{1}(\cup_{1\leq j\leq\ell}S_{j}) and θ1\theta\geq 1 on its support.

3 Proof of the results

We start to consider the optimization problem (2.3), which is a max-min problem:

supμ𝒜LinfuCc(Ω)Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑x.\sup_{\mu\in\mathscr{A}_{L}}\inf_{u\in C^{\infty}_{c}(\Omega)}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\;.

Proposition 2.1 gives the existence of an optimal relaxed solution, which is a measure μ\mu on Ω\Omega with μL\|\mu\|\leq L. Of course, since the cost above is monotone increasing with respect to μ\mu, optimal measures will saturate the constraint, so we will have μ=L\|\mu\|=L.

The main result of this section asserts that, under mild assumptions on the boundary of Ω\Omega, the optimal measures μ\mu are actually of the form θdx\theta\,dx, where θ\theta is a function that solves the optimization problem

max{λ1(θ):θ𝒜L}\max\big{\{}\lambda_{1}(\theta)\ :\ \theta\in\mathscr{A}_{L}\big{\}} (3.1)

and λ1(θ)\lambda_{1}(\theta) is defined by

λ1(θ)=min{Ω(1+mθ)|u|2𝑑xΩ|u|2𝑑x:uH01(Ω)}.\lambda_{1}(\theta)=\min\left\{\frac{\int_{\Omega}(1+m\theta)|\nabla u|^{2}\,dx}{\int_{\Omega}|u|^{2}\,dx}\ :\ u\in H^{1}_{0}(\Omega)\right\}.

Furthermore, we will see that the optimal densities θ\theta satisfy some higher-integrability properties and, if Ω\Omega is convex, belong to L(Ω)L^{\infty}(\Omega).

In order to obtain better properties of the optimal measure μ\mu provided by the existence result seen in Proposition 2.1 it is convenient to consider the optimization problem (2.3) under the stronger constraint that μ=θdx\mu=\theta\,dx with Ωθp𝑑xLp\int_{\Omega}\theta^{p}\,dx\leq L^{p} and p>1p>1. In other words, we consider the class

𝒜L,p={θLp(Ω):θ0 and Ωθp(x)𝑑xLp}\mathscr{A}_{L,p}=\left\{\theta\in L^{p}(\Omega)\ :\ \theta\geq 0\text{ and }\int_{\Omega}\theta^{p}(x)\,dx\leq L^{p}\right\}

and the optimization problem

max{λ1(θ):θ𝒜L,p}\max\big{\{}\lambda_{1}(\theta)\ :\ \theta\in\mathscr{A}_{L,p}\big{\}} (3.2)

We still have a max-min problem:

supθ𝒜L,pinfuH01(Ω)Ω(1+mθ)|u|2𝑑xΩ|u|2𝑑x.\sup_{\theta\in\mathscr{A}_{L,p}}\inf_{u\in H^{1}_{0}(\Omega)}\frac{\int_{\Omega}(1+m\theta)|\nabla u|^{2}\,dx}{\int_{\Omega}|u|^{2}\,dx}\;.
Proposition 3.1.

For every p>1p>1 there exist a unique solution θp\theta_{p} of the optimization problem (3.2), given by

θp(x)=L|up|2/(p1)|up|2/(p1)Lp.\theta_{p}(x)=L\frac{|\nabla u_{p}|^{2/(p-1)}}{\||\nabla u_{p}|^{2/(p-1)}\|_{L^{p}}}\;.

where upu_{p} is the unique positive solution with upL2(Ω)=1\|u_{p}\|_{L^{2}(\Omega)}=1 of the auxiliary problem

minuH01(Ω){0}Ω|u|2𝑑x+mL|u|2Lq(Ω)Ω|u|2𝑑x,\min_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\||\nabla u|^{2}\|_{L^{q}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}\;,

where q=p/(p1)q=p/(p-1) is the dual exponent of pp. Furthermore, the function upu_{p} belongs to L(Ω)W01,2q(Ω)L^{\infty}(\Omega)\cap W_{0}^{1,2q}(\Omega) and if in addition ΩC2,α\partial\Omega\in C^{2,\alpha}, then there exists β=β(α)(0,1)\beta=\beta(\alpha)\in(0,1) such that upC2,β(Ω¯)u_{p}\in C^{2,\beta}(\bar{\Omega}). In particular, θpC1,β\theta_{p}\in C^{1,\beta} up to the boundary.

Proof.

Let us denote by E(θ,u)E(\theta,u) and by (u)\mathcal{E}(u) the functionals

E(θ,u)=Ω(1+mθ)|u|2𝑑xΩ|u|2𝑑x,(u)=Ω|u|2𝑑x+mL|u|2Lq(Ω)Ω|u|2𝑑x.\begin{split}&E(\theta,u)=\frac{\int_{\Omega}(1+m\theta)|\nabla u|^{2}\,dx}{\int_{\Omega}|u|^{2}\,dx}\;,\\ &\mathcal{E}(u)=\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\||\nabla u|^{2}\|_{L^{q}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}\;.\end{split}

Then problem (3.2) is written as

maxθ𝒜L,pminuH01(Ω){0}E(θ,u).\max_{\theta\in\mathscr{A}_{L,p}}\ \min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}E(\theta,u).

Interchanging the max and the min above gives the inequality

maxθ𝒜L,pminuH01(Ω){0}E(θ,u)minuH01(Ω){0}maxθ𝒜L,pE(θ,u).\max_{\theta\in\mathscr{A}_{L,p}}\ \min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}E(\theta,u)\leq\min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}\ \max_{\theta\in\mathscr{A}_{L,p}}E(\theta,u). (3.3)

The maximum with respect to θ𝒜L,p\theta\in\mathscr{A}_{L,p} at the right-hand side above is easily computed and for every fixed uH01(Ω){0}u\in H^{1}_{0}(\Omega)\setminus\{0\} this maximum is reached at

θ=L|u|2/(p1)|u|2/(p1)Lp.\theta=L\frac{|\nabla u|^{2/(p-1)}}{\||\nabla u|^{2/(p-1)}\|_{L^{p}}}\;.

Then the right-hnd side in (3.3) becomes the auxiliary minimization problem

minuH01(Ω){0}(u).\min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}\mathcal{E}(u)\;.

A straightforward application of the direct methods of the calculus of variations gives the existence of an optimal solution upu_{p} of the auxiliary problem above. Setting

θp=L|up|2/(p1)|up|2/(p1)Lp\theta_{p}=L\frac{|\nabla u_{p}|^{2/(p-1)}}{\||\nabla u_{p}|^{2/(p-1)}\|_{L^{p}}}

by (3.3) we obtain

minuH01(Ω){0}E(θp,u)maxθ𝒜L,pminuH01(Ω){0}E(θ,u)minuH01(Ω){0}(u)=(up).\min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}E(\theta_{p},u)\leq\max_{\theta\in\mathscr{A}_{L,p}}\ \min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}E(\theta,u)\leq\min_{u\in H^{1}_{0}(\Omega)\setminus\{0\}}\mathcal{E}(u)=\mathcal{E}(u_{p})\;.

The minimum problem at the left-hand side above has upu_{p} as a solution, as it can be easily verified by performing the corresponding Euler-Lagrange equation. In addition, we have

E(θp,up)=(up),E(\theta_{p},u_{p})=\mathcal{E}(u_{p}),

so that finally we obtain the equality

maxθ𝒜L,pλ(θ)=λ(θp)\max_{\theta\in\mathscr{A}_{L,p}}\lambda(\theta)=\lambda(\theta_{p})

which proves the first assertion. In addition, the function upu_{p} verifies the PDE

div((1+mθp)up)=λ(θp)up,upH01(Ω).-\mathrm{div}\big{(}(1+m\theta_{p})\nabla u_{p}\big{)}=\lambda(\theta_{p})u_{p},\qquad u_{p}\in H^{1}_{0}(\Omega). (3.4)

The fact that upL(Ω)u_{p}\in L^{\infty}(\Omega) is standard (see Remark 3.6). To prove the Hölder-regularity, notice that, if Ω\partial\Omega is of class C2,αC^{2,\alpha}, then by [21] there exists β=β(α)(0,1)\beta=\beta(\alpha)\in(0,1) such that

upC1,β(Ω¯).u_{p}\in C^{1,\beta}(\bar{\Omega}).

Next, we can apply Theorem 1.2.12 of [18] to conclude that upC2,β(Ω¯)u_{p}\in C^{2,\beta}(\bar{\Omega}). The reason is that the coefficient 1+mθp1+m\theta_{p} of the PDE in (3.4) is Hölder-continuous with a parameter β\beta as a consequence of the definition of θp\theta_{p} itself. ∎

The next result shows that we can estimate λ1(θp)\lambda_{1}(\theta_{p}) uniformly with respect to pp.

Lemma 3.2.

Let p1p\geq 1. Then there exists a positive constant cc depending only on Ω\Omega, its volume and LL such that

λ1(θp)cfor all p1.\lambda_{1}(\theta_{p})\leq c\qquad\text{for all }p\geq 1.
Proof.

Since Ω\Omega is bounded, for uW1,2q(Ω)W1,(Ω)u\in W^{1,2q}(\Omega)\cap W^{1,\infty}(\Omega) we have

|u|2Lq(Ω)|Ω|1/q|u|2L(Ω)max{1,|Ω|}uL(Ω)2.\||\nabla u|^{2}\|_{L^{q}(\Omega)}\leq|\Omega|^{1/q}\||\nabla u|^{2}\|_{L^{\infty}(\Omega)}\leq\max\{1,|\Omega|\}\|\nabla u\|_{L^{\infty}(\Omega)}^{2}.

Therefore, we can bound the eigenvalue as follows

λ1(θp)=minuH01(Ω){0}Ω|u|2𝑑x+mL|u|2Lq(Ω)Ω|u|2𝑑xminuH01(Ω){0}Ω|u|2𝑑x+mLmax{1,|Ω|}uL(Ω)2Ω|u|2𝑑x.\begin{split}\lambda_{1}(\theta_{p})&=\min_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\||\nabla u|^{2}\|_{L^{q}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}\\ &\leq\min_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\max\{1,|\Omega|\}\|\nabla u\|_{L^{\infty}(\Omega)}^{2}}{\int_{\Omega}|u|^{2}\,dx}\;.\end{split}

The latter does not depend on pp, and a straightforward application of the direct methods of the calculus of variations shows that the minimum is achieved and it is finite. It follows that

λ1(θp)c(Ω,L)\lambda_{1}(\theta_{p})\leq c(\Omega,L)

for all p1p\geq 1, and this concludes the proof. ∎

Before passing to prove higher summability and regularity properties of the solutions of the optimization problem (3.1) we need to prove some uniform (with respect to pp) estimates of the solutions upu_{p} and θp\theta_{p}. An important step is a Γ\Gamma-convergence of the related functionals.

3.1 Γ\Gamma-convergence as p1p\to 1 of the functions upu_{p}

Let p>1p>1 and q=pq=p^{\prime} as before. Consider the family of functionals defined on L2(Ω)L^{2}(\Omega)

Fp(u):={Ω|u|2𝑑x+mL|u|2Lq(Ω)Ω|u|2𝑑xif uW01,2q(Ω),+otherwise,F_{p}(u):=\begin{cases}\displaystyle\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\||\nabla u|^{2}\|_{L^{q}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}&\text{if }u\in W_{0}^{1,2q}(\Omega),\\ +\infty&\text{otherwise,}\end{cases}

and for p=1p=1

F1(u):={Ω|u|2𝑑x+mLuL(Ω)2Ω|u|2𝑑xif uW01,(Ω),+otherwise.F_{1}(u):=\begin{cases}\displaystyle\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|^{2}_{L^{\infty}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}&\text{if }u\in W_{0}^{1,\infty}(\Omega),\\ +\infty&\text{otherwise.}\end{cases}

For p1p\geq 1 denote by upu_{p} the unique positive minimizer and with unitary L2L^{2} norm of the corresponding functional FpF_{p}.

Definition 3.3 (Γ\Gamma-convergence).

Let (X,d)(X,d) be a metric space. We say that a sequence of functionals

Fε:X{+}F_{\varepsilon}:X\longrightarrow\mathbb{R}\cup\{+\infty\}

Γ\Gamma-converges to a functional F:X{+}F:X\to\mathbb{R}\cup\{+\infty\} if the following hold:

  1. (i)

    for every sequence (xn)(x_{n}) in XX converging to some xXx\in X we have (often called Γlim inf\Gamma-\liminf inequality)

    F(x)lim infnFn(xn);F(x)\leq\liminf_{n\to\infty}F_{n}(x_{n});
  2. (ii)

    for every xXx\in X there is a sequence (xn)(x_{n}) in XX converging to xx such that (often called Γlim sup\Gamma-\limsup inequality)

    F(x)lim supnFn(xn).F(x)\geq\limsup_{n\to\infty}F_{n}(x_{n}).
Proposition 3.4.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded set with |Ω|<+|\Omega|<+\infty. Then

  1. (a)

    the sequence of functionals FpF_{p} Γ\Gamma-converges to F1F_{1} in L2(Ω)L^{2}(\Omega);

  2. (b)

    the sequence of minima upu_{p} converges strongly in H1(Ω)H^{1}(\Omega) to u1u_{1} and

    limp1+upL2q(Ω)=u1L(Ω).\lim_{p\to 1^{+}}\|\nabla u_{p}\|_{L^{2q}(\Omega)}=\|\nabla u_{1}\|_{L^{\infty}(\Omega)}.
Proof.

The same proof given in [8, Proposition 3.3] works if we set f:=λ(θp)upf:=\lambda(\theta_{p})u_{p} because we can estimate it in H1(Ω)H^{1}(\Omega) uniformly with respect to pp. In particular, we have

λ(θp)upL2(Ω)=λ(θp)c\|\lambda(\theta_{p})u_{p}\|_{L^{2}(\Omega)}=\lambda(\theta_{p})\leq c

as a consequence of Lemma 3.2 and

Ω|up|2𝑑x=λ(θp)upL2(Ω)mL|up|2Lq(Ω)upL2(Ω)c.\int_{\Omega}|\nabla u_{p}|^{2}\,dx=\|\lambda(\theta_{p})u_{p}\|_{L^{2}(\Omega)}-mL\||\nabla u_{p}|^{2}\|_{L^{q}(\Omega)}\implies\|\nabla u_{p}\|_{L^{2}(\Omega)}\leq c.

in a similar way. ∎

3.2 Uniform estimate of upu_{p}

To find a uniform estimate of upu_{p} and θp\theta_{p} we need to assume that a certain geometric condition is satisfied by Ω\Omega, namely that there is a uniform ρ>0\rho>0 such that the external ball condition (see Definition 2.3) holds with ρ\rho at all xΩx\in\Omega. Following closely the method developed in [8], we obtain an almost uniform estimate on the LL^{\infty}-norm of the gradient up\nabla u_{p} on the boundary of Ω\Omega since the LL^{\infty}-norm of upu_{p} must be taken into account. We first recall a few regularity properties.

Lemma 3.5.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded Lipschitz domain and let p1p\geq 1. Then

upW2,r(Ω)for all r>d.u_{p}\in W^{2,r}(\Omega)\qquad\text{for all }r>d.

In particular, if Ω\Omega is of class C1,1C^{1,1} then upC1(Ω¯)u_{p}\in C^{1}(\bar{\Omega}).

Proof.

It follows from [16, Theorem 9.15] and a standard bootstrap argument. ∎

Remark 3.6.

The Sobolev embedding theorem [14] gives us another proof of the fact that upL(Ω)u_{p}\in L^{\infty}(\Omega) and, more precisely, that when Ω\Omega is of class C2,αC^{2,\alpha},

upC1,β(Ω¯)for some 0<β=β(α)<1.u_{p}\in C^{1,\beta}(\bar{\Omega})\qquad\text{for some }0<\beta=\beta(\alpha)<1.

The main ingredients behind the uniform estimate of up\nabla u_{p} are two weak comparison principles for eigenfunctions and the fact that on a radially symmetric domain the optimal pair (θp,up)(\theta_{p},u_{p}) is radial (see Lemma 4.1).

Lemma 3.7 (Weak comparison principle).

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded connected open set and let G:[0,+)[0,+)G:[0,+\infty)\to[0,+\infty) be a convex function such that G(0)>0G^{\prime}(0)>0.

  1. (i)

    Denote by uΩu_{\Omega} the unique positive solution of

    {div(G(|u|2)u(x))=λ1(Ω,G)u(x)if xΩ,u(x)=0if xΩ,\begin{cases}-\mathrm{div}\big{(}G^{\prime}(|\nabla u|^{2})\nabla u(x)\big{)}=\lambda_{1}(\Omega,G)u(x)&\text{if }x\in\Omega,\\ u(x)=0&\text{if }x\in\partial\Omega,\end{cases}

    with unitary L2L^{2} norm. Then, for any ωΩ\omega\subset\Omega bounded open subset, it turns out that uΩuωu_{\Omega}\geq u_{\omega}.

  2. (ii)

    Let uΩu_{\Omega} be as above. If u¯\bar{u} is the unique positive solution of

    {div(G(|u|2)u(x))=λ1(Ω,G)uΩL(Ω)if xΩ,u(x)=0if xΩ,\begin{cases}-\mathrm{div}\big{(}G^{\prime}(|\nabla u|^{2})\nabla u(x)\big{)}=\lambda_{1}(\Omega,G)\|u_{\Omega}\|_{L^{\infty}(\Omega)}&\text{if }x\in\Omega,\\ u(x)=0&\text{if }x\in\partial\Omega,\end{cases}

    with unitary L2L^{2} norm, then u¯uΩ\bar{u}\geq u_{\Omega}.

Remark 3.8.

The assumption Ω\Omega connected ensures that we can choose uΩu_{\Omega} and uωu_{\omega} to be the unique positive solutions with fixed L2L^{2}-norm.

Lemma 3.9.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open set with boundary in Cloc1,αC_{\mathrm{loc}}^{1,\alpha} for some α>0\alpha>0. Suppose that Ω\Omega satisfies the external ball condition at some x0Ωx_{0}\in\partial\Omega with radius ρ>0\rho>0. Then there is a constant c=c(|Ω|,L)c=c(|\Omega|,L) such that

(1+Cp|up|2(q1)(x0))upL(Ω)c(1+diam(Ω)ρ)d1diam(Ω),\frac{\left(1+C_{p}|\nabla u_{p}|^{2(q-1)}(x_{0})\right)}{\|u_{p}\|_{L^{\infty}(\Omega)}}\leq c\left(1+\frac{\mathrm{diam}(\Omega)}{\rho}\right)^{d-1}\mathrm{diam}(\Omega), (3.5)

where

Cp:=mL(Ω|up|2q(x)𝑑x)1/p.C_{p}:=mL\left(\int_{\Omega}|\nabla u_{p}|^{2q}(x)\,dx\right)^{-1/p}.
Proof.

Introduce the auxiliary function

G(t):=t+CpqtqG(t):=t+\frac{C_{p}}{q}t^{q}

and notice that it satisfies the assumptions of Lemma 3.7. We can assume without loss of generality that the center of the external ball at x0x_{0} is the origin (i.e., y0=0y_{0}=0) so that, setting R:=diam(Ω)R:=\mathrm{diam}(\Omega), we obtain the inclusion

ΩBR+ρB¯ρ=:CR,ρ.\Omega\subset B_{R+\rho}\setminus\bar{B}_{\rho}=:C_{R,\rho}.

Let u~\tilde{u} be the solution of

div(G(|u|2)u)=λ1(θp)upL(Ω)-\mathrm{div}\big{(}G^{\prime}(|\nabla u|^{2})\nabla u\big{)}=\lambda_{1}(\theta_{p})\|u_{p}\|_{L^{\infty}(\Omega)}

with uW01,2q(Ω)u\in W_{0}^{1,2q}(\Omega) and L2L^{2}-norm equal to one, and let UU be the solution of

div(G(|u|2)u)=λ1(CR,ρ,θp)upL(Ω)-\mathrm{div}\big{(}G^{\prime}(|\nabla u|^{2})\nabla u\big{)}=\lambda_{1}(C_{R,\rho},\theta_{p})\|u_{p}\|_{L^{\infty}(\Omega)}

with uW01,2q(CR,ρ)u\in W_{0}^{1,2q}(C_{R,\rho}) and the same L2L^{2}-norm as above. Using (i)(i) and (ii)(ii) of Lemma 3.7 we find that

upu¯Uu_{p}\leq\bar{u}\leq U

since Ω\Omega is contained in the annulus CR,ρC_{R,\rho}. The function UU is radially symmetric, and therefore we can rewrite the equation in polar coordinates:

{r1dr(rd1G(|U|2)U)=λ1(CR,ρ,θp)upL(Ω)if r(ρ,ρ+R),U(ρ)=U(ρ+R)=0.\begin{cases}-r^{1-d}\partial_{r}(r^{d-1}G(|U^{\prime}|^{2})U^{\prime})=\lambda_{1}(C_{R,\rho},\theta_{p})\|u_{p}\|_{L^{\infty}(\Omega)}&\text{if }r\in(\rho,\rho+R),\\ U(\rho)=U(\rho+R)=0.\end{cases}

Let ρ1\rho_{1} be the point where UU attains its maximum value and integrate the previous equation to deduce that

ρd1(1+Cp|U|2(q1))U(ρ)=λ1(CR,ρ,θp)ρρ1rd1upL(Ω)𝑑rλ1(CR,ρ,θp)upL(Ω)R(R+ρ)dd,\begin{split}\rho^{d-1}(1+C_{p}|U^{\prime}|^{2(q-1)})U^{\prime}(\rho)&=\lambda_{1}(C_{R,\rho},\theta_{p})\int_{\rho}^{\rho_{1}}r^{d-1}\|u_{p}\|_{L^{\infty}(\Omega)}\,dr\\ &\leq\lambda_{1}(C_{R,\rho},\theta_{p})\|u_{p}\|_{L^{\infty}(\Omega)}\frac{R(R+\rho)^{d}}{d},\end{split}

and next we estimate the left-hand side using that the radial derivative rU\partial_{r}U is positive. Finally, the inclusion property of eigenvalues,

ΩCR,ρλ1(CR,ρ,θp)λ1(θp),\Omega\subset C_{R,\rho}\implies\lambda_{1}(C_{R,\rho},\theta_{p})\leq\lambda_{1}(\theta_{p}),

together with Lemma 3.2, allows us to infer that (3.5) holds. ∎

If now Ω\Omega satisfies the uniform external ball condition, we can extend the estimate (3.5) to hold for upL(Ω)\|\nabla u_{p}\|_{L^{\infty}(\partial\Omega)} provided that we replace ρ\rho with the largest possible radius.

Lemma 3.10.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open set with boundary in Cloc1,αC_{\mathrm{loc}}^{1,\alpha} for some α>0\alpha>0. The following assertions hold:

  1. (i)

    If Ω\Omega is convex, then

    (1+CpupL(Ω)2(q1))upL(Ω)cdiam(Ω),\frac{\left(1+C_{p}\|\nabla u_{p}\|_{L^{\infty}(\partial\Omega)}^{2(q-1)}\right)}{\|u_{p}\|_{L^{\infty}(\Omega)}}\leq c\mathrm{diam}(\Omega),

    where cc is the constant given in Lemma 3.2.

  2. (ii)

    If Ω\Omega satisfies the uniform external boundary condition with radius ρ\rho, then

    (1+CpupL(Ω)2(q1))upL(Ω)c(1+diam(Ω)ρ)d1diam(Ω).\frac{\left(1+C_{p}\|\nabla u_{p}\|_{L^{\infty}(\partial\Omega)}^{2(q-1)}\right)}{\|u_{p}\|_{L^{\infty}(\Omega)}}\leq c\left(1+\frac{\mathrm{diam}(\Omega)}{\rho}\right)^{d-1}\mathrm{diam}(\Omega).

The estimate obtained when Ω\Omega is convex is more precise, but it is not enough to infer that the optimal density belongs to L(Ω)L^{\infty}(\Omega) through the method presented in this section. In any case, the LL^{\infty}-norm of upu_{p} appears at the denominator: we will see later that this does not lead to any additional problem.

3.3 Uniform estimate of θp\theta_{p}

The next step is to find a uniform estimate for the LrL^{r}-norm of θp\theta_{p}; more precisely, we prove that

θpLr(Ω)C,\|\theta_{p}\|_{L^{r}(\Omega)}\leq C,

where CC is a positive constant that depends on rr but not on pp. It is worth remarking that eigenfunctions are bounded in LL^{\infty}, so it is convenient to work with their LL^{\infty} norm only.

Remark 3.11.

Let r1r\geq 1. Then

upLr(Ω)upL(Ω)|Ω|1/rmax{|Ω|,1}upL(Ω),\|u_{p}\|_{L^{r}(\Omega)}\leq\|u_{p}\|_{L^{\infty}(\Omega)}|\Omega|^{1/r}\leq\max\{|\Omega|,1\}\|u_{p}\|_{L^{\infty}(\Omega)},

which means that we can always find a uniform estimate of the LrL^{r}-norm using the LL^{\infty}-norm. This will be particularly important in the uniform estimate of θp\theta_{p}.

We are now ready to prove the a priori estimate for θp\theta_{p}. The key lemma below is due to De Pascale-Evans-Pratelli in [11] and the proof is adapted to the eigenvalue case in the spirit of [8].

Lemma 3.12 (De Pascale-Evans-Pratelli).

Let Ωd\Omega\subset\mathbb{R}^{d} be a bounded open set with smooth boundary and let r2r\geq 2 be arbitrary. Let G:[0,+)[0,)G:[0,+\infty)\to[0,\infty) be a convex function with G(0)>0G^{\prime}(0)>0 and let

uC1(Ω¯)Hloc2(Ω)u\in C^{1}(\bar{\Omega})\cap H_{\mathrm{loc}}^{2}(\Omega)

be the unique positive solution of

div(G(|u|2)u)=λ1u,-\mathrm{div}(G^{\prime}(|\nabla u|^{2})\nabla u)=\lambda_{1}u, (3.6)

where λ1\lambda_{1} depends on Ω\Omega and GG only, satisfying u=0u=0 on Ω\partial\Omega and with unitary L2L^{2}-norm. Then for every ε(0,1)\varepsilon\in(0,1) we have the following estimate:

Ω|G(|u|2)|r|u|2𝑑x3εG(|u|2)Lr(Ω)r+|Ω|((r1)rεr1+1ε2r1)λ1ruL(Ω)2r(r1)2uL(Ω)2εΩHG(|u|2)r|u|2𝑑d1,\begin{split}\int_{\Omega}|G^{\prime}(|\nabla u|^{2})|^{r}|\nabla u|^{2}\,dx\leq&3\varepsilon\|G^{\prime}(|\nabla u|^{2})\|_{L^{r}(\Omega)}^{r}+|\Omega|\left(\frac{(r-1)^{r}}{\varepsilon^{r-1}}+\frac{1}{\varepsilon^{2r-1}}\right)\lambda_{1}^{r}\|u\|_{L^{\infty}(\Omega)}^{2r}\\ &-\frac{(r-1)^{2}\|u\|_{L^{\infty}(\Omega)}^{2}}{\varepsilon}\int_{\partial\Omega}HG^{\prime}(|\nabla u|^{2})^{r}|\nabla u|^{2}\,d\mathcal{H}^{d-1},\end{split}

where HH is the mean curvature of Ω\partial\Omega with respect to the outer normal and |Ω||\Omega| is the volume of the set Ω\Omega.

Proof.

The assumption Ω\Omega smooth implies that the solution uu to (3.6) is smooth up to the boundary, that is, uC(Ω¯)u\in C^{\infty}(\bar{\Omega}). Set

σ:=G(|u|2),\sigma:=G^{\prime}(|\nabla u|^{2}),

and use σr1uH01(Ω)\sigma^{r-1}u\in H_{0}^{1}(\Omega) as a test function for the equation (3.6). Using the integration by parts formula on the left-hand side, we find the identity

Ωσr|u|2𝑑x+(r1)Ωuσr1uσdx=λ1Ωσr1|u|2𝑑x.\int_{\Omega}\sigma^{r}|\nabla u|^{2}\,dx+(r-1)\int_{\Omega}u\sigma^{r-1}\nabla u\cdot\nabla\sigma\,dx=\lambda_{1}\int_{\Omega}\sigma^{r-1}|u|^{2}\,dx. (3.7)

The right-hand side of the equality can be easily estimated via the Hölder inequality and Remark 3.11 as:

λ1Ωσr1|u|2𝑑xλ1|Ω|1/ruL(Ω)2σLr(Ω)r1.\lambda_{1}\int_{\Omega}\sigma^{r-1}|u|^{2}\,dx\leq\lambda_{1}|\Omega|^{1/r}\|u\|_{L^{\infty}(\Omega)}^{2}\|\sigma\|_{L^{r}(\Omega)}^{r-1}.

To estimate the integral Ωuσr1uσdx\int_{\Omega}u\sigma^{r-1}\nabla u\cdot\nabla\sigma\,dx we can use φ:=div(σr1u)\varphi:=\mathrm{div}(\sigma^{r-1}\nabla u) as test function for the equation (3.6); it turns out that

Ωdiv(σu)div(σr1u)dx=λ1Ωdiv(σr2σu)u𝑑x=λ1Ω(σr2div(σu)u+(r2)σr2(uσ)u)𝑑xλ1Ω(λ1σr2|u|2+(r2)σr2|uσ||u|)𝑑x.\begin{split}\int_{\Omega}\mathrm{div}(\sigma\nabla u)&\mathrm{div}(\sigma^{r-1}\nabla u)\,dx=-\lambda_{1}\int_{\Omega}\mathrm{div}(\sigma^{r-2}\sigma\nabla u)u\,dx\\ &=-\lambda_{1}\int_{\Omega}\Big{(}\sigma^{r-2}\mathrm{div}(\sigma\nabla u)u+(r-2)\sigma^{r-2}(\nabla u\cdot\nabla\sigma)u\Big{)}\,dx\\ &\leq\lambda_{1}\int_{\Omega}\Big{(}\lambda_{1}\sigma^{r-2}|u|^{2}+(r-2)\sigma^{r-2}|\nabla u\cdot\nabla\sigma||u|\Big{)}\,dx.\end{split} (3.8)

Integrating by parts twice the left-hand side leads to the following chain of equalities,

Ωdiv(σu)div(σr1u)dx=Ωσu[div(σr1u)]dx+Ωσdiv(σr1u)uν𝑑d1=Ω(σui)j(σr1uj)i𝑑x+Ωσr(uνΔuuiuijνj)𝑑d1=Ω(σui)j(σr1uj)i𝑑x+Ωσr(Δuuνν)uν𝑑d1=Ω(σui)j(σr1uj)i𝑑x+Ωσruν(Δuuνν)𝑑d1,\begin{split}\int_{\Omega}\mathrm{div}&(\sigma\nabla u)\mathrm{div}(\sigma^{r-1}\nabla u)\,dx\\ &=-\int_{\Omega}\sigma\nabla u\cdot\nabla\left[\mathrm{div}(\sigma^{r-1}\nabla u)\right]\,dx+\int_{\partial\Omega}\sigma\mathrm{div}(\sigma^{r-1}\nabla u)u_{\nu}\,d\mathcal{H}^{d-1}\\ &=\int_{\Omega}(\sigma u_{i})_{j}(\sigma^{r-1}u_{j})_{i}\,dx+\int_{\partial\Omega}\sigma^{r}\left(u_{\nu}\Delta u-u_{i}u_{ij}\nu_{j}\right)\,d\mathcal{H}^{d-1}\\ &=\int_{\Omega}(\sigma u_{i})_{j}(\sigma^{r-1}u_{j})_{i}\,dx+\int_{\partial\Omega}\sigma^{r}\left(\Delta u-u_{\nu\nu}\right)u_{\nu}\,d\mathcal{H}^{d-1}\\ &=\int_{\Omega}(\sigma u_{i})_{j}(\sigma^{r-1}u_{j})_{i}\,dx+\int_{\partial\Omega}\sigma^{r}u_{\nu}(\Delta u-u_{\nu\nu})\,d\mathcal{H}^{d-1},\end{split}

where uν=uνu_{\nu}=\nabla u\cdot\nu and uνν=Hess(u)ννu_{\nu\nu}=\mathrm{Hess}(u)\nu\cdot\nu are, respectively, the first-order and second-order derivatives in the direction of the exterior normal ν\nu to Ω\partial\Omega and we introduce the short notation uj=juu_{j}=\partial_{j}u. It follows that

Ωdiv(σu)div(σr1u)dx=ΩσrHess(u)22𝑑x+(r1)Ωσr2|uσ|2𝑑x+rΩσr1σjuiuij𝑑x+Ωσruν(Δuuνν)𝑑d1\begin{split}\int_{\Omega}\mathrm{div}(\sigma\nabla u)&\mathrm{div}(\sigma^{r-1}\nabla u)\,dx=\int_{\Omega}\sigma^{r}\|\mathrm{Hess}(u)\|_{2}^{2}\,dx+(r-1)\int_{\Omega}\sigma^{r-2}|\nabla u\cdot\nabla\sigma|^{2}\,dx\\ &+r\int_{\Omega}\sigma^{r-1}\sigma_{j}u_{i}u_{ij}\,dx+\int_{\partial\Omega}\sigma^{r}u_{\nu}(\Delta u-u_{\nu\nu})\,d\mathcal{H}^{d-1}\end{split}

where the 22-norm associated to the Hessian matrix is given by

Hess(u)22:=i,j=1duij2.\|\mathrm{Hess}(u)\|_{2}^{2}:=\sum_{i,j=1}^{d}u_{ij}^{2}.

Since uu is smooth up to the boundary of Ω\Omega, it is easy to verify that the Laplace operator can be decomposed as

Δu=uνν+Huν(x)for all xΩ,\Delta u=u_{\nu\nu}+Hu_{\nu}(x)\quad\text{for all $x\in\partial\Omega$},

where HH denotes the mean curvature of Ω\partial\Omega. This immediately implies the estimate

(r1)Ωσr2|uσ|2𝑑x+ΩHσr|u|2𝑑d1Ωdiv(σu)div(σr1u)𝑑x.\begin{split}(r-1)\int_{\Omega}\sigma^{r-2}|\nabla u\cdot\nabla\sigma|^{2}\,dx+&\int_{\partial\Omega}H\sigma^{r}|\nabla u|^{2}\,d\mathcal{H}^{d-1}\\ &\leq\int_{\Omega}\mathrm{div}(\sigma\nabla u)\mathrm{div}(\sigma^{r-1}\nabla u)\,dx.\end{split} (3.9)

In fact, it is easy to verify that

rΩσr1σjuiuij𝑑x+ΩσrHess(u)22𝑑x0r\int_{\Omega}\sigma^{r-1}\sigma_{j}u_{i}u_{ij}\,dx+\int_{\Omega}\sigma^{r}\|\mathrm{Hess}(u)\|_{2}^{2}\,dx\geq 0

since the 22-norm is positive by definition and σjuiuij0\sigma_{j}u_{i}u_{ij}\geq 0 follows from the convexity assumption on GG. We now plug (3.9) into (3.8) and use the Hölder inequality with p=r/2p=r/2 and q=r/(r2)q=r/(r-2) to obtain the following estimate:

Ωσr2|uσ|2𝑑xλ12Ω|u|2σr2𝑑xΩHσr|u|2𝑑d1λ12uLr(Ω)2σLr(Ω)r2ΩHσr|u|2𝑑d1.\begin{split}\int_{\Omega}\sigma^{r-2}|\nabla u\cdot\nabla\sigma|^{2}\,dx&\leq\lambda_{1}^{2}\int_{\Omega}|u|^{2}\sigma^{r-2}\,dx-\int_{\partial\Omega}H\sigma^{r}|\nabla u|^{2}\,d\mathcal{H}^{d-1}\\ &\leq\lambda_{1}^{2}\|u\|_{L^{r}(\Omega)}^{2}\|\sigma\|_{L^{r}(\Omega)}^{r-2}-\int_{\partial\Omega}H\sigma^{r}|\nabla u|^{2}\,d\mathcal{H}^{d-1}.\end{split}

The conclusion follows by combining the inequalities discovered so far with the identity (3.7) and applying repeatedly the Young inequality

AαBβεαA+εα/ββB,A^{\alpha}B^{\beta}\leq\varepsilon\alpha A+\varepsilon^{\alpha/\beta}\beta B,

which is valid for α+β=1\alpha+\beta=1. In particular, it turns out that

Ωσr|u|2𝑑xuL(Ω)Ωσr1|uσ|𝑑x+λ1uL(Ω)2|Ω|1/rσLr(Ω)r1εσLr(Ω)r+(r1)2uL(Ω)2εΩσr2|uσ|2𝑑x+λ1uL(Ω)2|Ω|1rσLr(Ω)r1εσLr(Ω)r+λ12(r1)2uL(Ω)2εuLr(Ω)2σLr(Ω)r2+λ1uL(Ω)2|Ω|1rσLr(Ω)r1(r1)2uL(Ω)2εΩHσr|u|2𝑑d13εσLr(Ω)r+|Ω|((r1)rεr1+1ε2r1)λ1ruL(Ω)2r(r1)2uL(Ω)2εΩHσr|u|2𝑑d1,\begin{split}\int_{\Omega}\sigma^{r}|\nabla u|^{2}\,dx&\leq\|u\|_{L^{\infty}(\Omega)}\int_{\Omega}\sigma^{r-1}|\nabla u\cdot\nabla\sigma|\,dx+\lambda_{1}\|u\|_{L^{\infty}(\Omega)}^{2}|\Omega|^{1/r}\|\sigma\|_{L^{r}(\Omega)}^{r-1}\\ &\leq\varepsilon\|\sigma\|_{L^{r}(\Omega)}^{r}+\frac{(r-1)^{2}\|u\|_{L^{\infty}(\Omega)}^{2}}{\varepsilon}\int_{\Omega}\sigma^{r-2}|\nabla u\cdot\nabla\sigma|^{2}\,dx\\ &\quad+\lambda_{1}\|u\|_{L^{\infty}(\Omega)}^{2}|\Omega|^{\frac{1}{r}}\|\sigma\|_{L^{r}(\Omega)}^{r-1}\\ &\leq\varepsilon\|\sigma\|_{L^{r}(\Omega)}^{r}+\lambda_{1}^{2}\frac{(r-1)^{2}\|u\|_{L^{\infty}(\Omega)}^{2}}{\varepsilon}\|u\|_{L^{r}(\Omega)}^{2}\|\sigma\|_{L^{r}(\Omega)}^{r-2}\\ &\quad+\lambda_{1}\|u\|_{L^{\infty}(\Omega)}^{2}|\Omega|^{\frac{1}{r}}\|\sigma\|_{L^{r}(\Omega)}^{r-1}-\frac{(r-1)^{2}\|u\|_{L^{\infty}(\Omega)}^{2}}{\varepsilon}\int_{\partial\Omega}H\sigma^{r}|\nabla u|^{2}\,d\mathcal{H}^{d-1}\\ &\leq 3\varepsilon\|\sigma\|_{L^{r}(\Omega)}^{r}+|\Omega|\left(\frac{(r-1)^{r}}{\varepsilon^{r-1}}+\frac{1}{\varepsilon^{2r-1}}\right)\lambda_{1}^{r}\|u\|_{L^{\infty}(\Omega)}^{2r}\\ &\quad-\frac{(r-1)^{2}\|u\|_{L^{\infty}(\Omega)}^{2}}{\varepsilon}\int_{\partial\Omega}H\sigma^{r}|\nabla u|^{2}\,d\mathcal{H}^{d-1},\end{split}

and this concludes the proof of the lemma. ∎

The De Pascale-Evans-Pratelli lemma holds for domains Ω\Omega with smooth boundary, so before passing to the uniform estimate of θp\theta_{p} we need to present an approximation argument that allows us to use domains with Lipschitz boundary that only satisfy the uniform external ball condition.

Lemma 3.13.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open set satisfying the uniform external ball condition and let Ωn\Omega_{n} be a sequence of open sets, with |Ωn|<|\Omega_{n}|<\infty and such that

ΩΩnand|ΩnΩ|0 as n.\Omega\subseteq\Omega_{n}\qquad\text{and}\qquad|\Omega_{n}\setminus\Omega|\to 0\hbox{ as }n\to\infty.

Fix p[1,)p\in[1,\infty) and let upu_{p} be the minimizer of FpF_{p} on Ω\Omega and upnu_{p}^{n} the minimizers of FpF_{p} on Ωn\Omega_{n}, all positive with fixed L2L^{2}-norm equal to one and extended by zero outside Ω\Omega and Ωn\Omega_{n} respectively. Then

upnupu_{p}^{n}\to u_{p}

strongly in both H1(2)H^{1}(\mathbb{R}^{2}) and W1,2q(2)W^{1,2q}(\mathbb{R}^{2}).

This is proved in [8, Lemma 3.9] for the energy problem, but the same arguments apply to the eigenvalue case. We finally have all the ingredients we need to obtain a uniform estimate of θp\theta_{p}:

Proposition 3.14.

Let Ω2\Omega\subset\mathbb{R}^{2} be a set of finite perimeter satisfying the uniform external ball condition with radius RR. For every rdr\geq d, there are constants

δ(Ω):=δandC(r,d,Per(Ω),λ1,diam(Ω),R,u1L(Ω))=C,\delta(\Omega):=\delta\qquad\text{and}\qquad C(r,d,\mathrm{Per}(\Omega),\lambda_{1},\mathrm{diam}(\Omega),R,\|u_{1}\|_{L^{\infty}(\Omega)})=C,

satisfying the inequality

θpLr(Ω)Cfor all p(1,1+δ).\|\theta_{p}\|_{L^{r}(\Omega)}\leq C\qquad\text{for all }p\in(1,1+\delta). (3.10)
Proof.

We first suppose that Ω\Omega is smooth. Set

Gp(t):=t+Cpqtq.G_{p}(t):=t+\frac{C_{p}}{q}t^{q}.

and notice that upu_{p} is the minimizer of the functional

H01(Ω)uΩGp(|u|2)𝑑xH_{0}^{1}(\Omega)\ni u\longmapsto\int_{\Omega}G_{p}(|\nabla u|^{2})\,dx

with L2L^{2}-norm equal to one. It is easy to verify that θp\theta_{p} is the derivative, that is,

θp=Gp(|up|2)1,\theta_{p}=G^{\prime}_{p}(|\nabla u_{p}|^{2})-1,

and, since HΩ1/RH_{\partial\Omega}\geq-1/R by smoothness of Ω\Omega, it follows from De Pascale-Evans-Pratelli (Lemma 3.12) that

Ωσpr|up|2𝑑x3εσpLr(Ω)r+|Ω|((r1)rεr1+ε12r)λ1(Gp)rupL(Ω)2r(r1)2upL(Ω)2εRΩσpr|up|2𝑑d1.\begin{split}\int_{\Omega}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,dx\leq&3\varepsilon\|\sigma_{p}\|_{L^{r}(\Omega)}^{r}+|\Omega|\left(\frac{(r-1)^{r}}{\varepsilon^{r-1}}+\varepsilon^{1-2r}\right)\lambda_{1}(G_{p})^{r}\|u_{p}\|_{L^{\infty}(\Omega)}^{2r}\\ &-\frac{(r-1)^{2}\|u_{p}\|_{L^{\infty}(\Omega)}^{2}}{\varepsilon R}\int_{\partial_{\Omega}}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,d\mathcal{H}^{d-1}.\end{split}

It follows from the definition of GpG_{p} that λ1\lambda_{1} is equal to λ1(θp)\lambda_{1}(\theta_{p}), which is uniformly bounded (Lemma 3.2) by a positive constant Λ\Lambda. We Denote by SS and BB the sets

S={xΩ¯:|up(x)|upL2q(Ω)},B={xΩ¯:|up(x)|>upL2q(Ω)},\begin{split}&S=\big{\{}x\in\bar{\Omega}\ :\ |\nabla u_{p}(x)|\leq\|\nabla u_{p}\|_{L^{2q}(\Omega)}\big{\}},\\ &B=\big{\{}x\in\bar{\Omega}\ :\ |\nabla u_{p}(x)|>\|\nabla u_{p}\|_{L^{2q}(\Omega)}\big{\}},\end{split}

and notice that

σp:=Gp(|up|2)1+mLon S.\sigma_{p}:=G_{p}(|\nabla u_{p}|^{2})\leq 1+mL\qquad\hbox{on }S.

We now estimate separately the three terms on the right-hand side of the inequality above. The first one gives

σpLr(Ω)r=σpLr(S)r+σpLr(B)r(1+mL)r|Ω|+upL2q(Ω)2Bσpr|up|2𝑑x.\begin{split}\|\sigma_{p}\|_{L^{r}(\Omega)}^{r}&=\|\sigma_{p}\|_{L^{r}(S)}^{r}+\|\sigma_{p}\|_{L^{r}(B)}^{r}\\ &\leq(1+mL)^{r}|\Omega|+\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{-2}\int_{B}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,dx.\end{split}

For the third term we can use the estimate (ii)(ii) given in Lemma 3.10 to obtain

Ωσpr|up|2dd1=SΩσpr|up|2𝑑d1+BΩσpr|up|2𝑑d1(1+mL)r2SΩσp2|up|2𝑑d1+upL2q(Ω)2rBΩσpr|up|r𝑑d1c2(1+mL)r2(1+diam(Ω)ρ)2(d1)Per(Ω)diam2(Ω)upL(Ω)2+crupL2q(Ω)2r(1+diam(Ω)ρ)r(d1)Per(Ω)diamr(Ω)upL(Ω)r.\begin{split}\int_{\partial\Omega}\sigma_{p}^{r}|\nabla u_{p}|^{2}&\,d\mathcal{H}^{d-1}=\int_{S\cap\partial\Omega}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,d\mathcal{H}^{d-1}+\int_{B\cap\partial\Omega}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,d\mathcal{H}^{d-1}\\ \leq&(1+mL)^{r-2}\int_{S\cap\partial\Omega}\sigma_{p}^{2}|\nabla u_{p}|^{2}\,d\mathcal{H}^{d-1}\\ &+\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2-r}\int_{B\cap\partial\Omega}\sigma_{p}^{r}|\nabla u_{p}|^{r}\,d\mathcal{H}^{d-1}\\ \leq&c^{2}(1+mL)^{r-2}\left(1+\frac{\mathrm{diam}(\Omega)}{\rho}\right)^{2(d-1)}\mathrm{Per}(\Omega)\mathrm{diam}^{2}(\Omega)\|u_{p}\|_{L^{\infty}(\Omega)}^{2}\\ &+c^{r}\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2-r}\left(1+\frac{\mathrm{diam}(\Omega)}{\rho}\right)^{r(d-1)}\mathrm{Per}(\Omega)\mathrm{diam}^{r}(\Omega)\|u_{p}\|_{L^{\infty}(\Omega)}^{r}.\end{split}

Now take ε=16upL2q(Ω)2\varepsilon=\frac{1}{6}\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2} in the initial inequality and rearrange the terms in such a way that the following holds:

12Ωσpr|up|2𝑑x12upL2q(Ω)2(1+mL)r|Ω|+62r1|Ω|((r1)rupL2q(Ω)2r2+1upL2q(Ω)4r2)ΛrupL(Ω)2r+6(r1)2upL(Ω)2upL2q(Ω)2R[(1+mL)r2C2(Ω)uL(Ω)2+Cr(Ω)upL2q(Ω)2rupL(Ω)r],\begin{split}\frac{1}{2}\int_{\Omega}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,dx\leq&\frac{1}{2}\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2}(1+mL)^{r}|\Omega|\\ &+6^{2r-1}|\Omega|\left(\frac{(r-1)^{r}}{\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2r-2}}+\frac{1}{\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{4r-2}}\right)\Lambda^{r}\|u_{p}\|_{L^{\infty}(\Omega)}^{2r}\\ &+\frac{6(r-1)^{2}\|u_{p}\|_{L^{\infty}(\Omega)}^{2}}{\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2}R}\Big{[}(1+mL)^{r-2}C_{2}(\Omega)\|u\|_{L^{\infty}(\Omega)}^{2}\\ &+C_{r}(\Omega)\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2-r}\|u_{p}\|_{L^{\infty}(\Omega)}^{r}\Big{]},\end{split}

where

Cs(Ω):=cs(1+diam(Ω)R)s(d1)Per(Ω)diams(Ω).C_{s}(\Omega):=c^{s}\left(1+\frac{\mathrm{diam}(\Omega)}{R}\right)^{s(d-1)}\mathrm{Per}(\Omega)\mathrm{diam}^{s}(\Omega).

Now notice that the same inequality holds for any Ω\Omega as in the statement since we can approximate it by a sequence of smooth sets Ωn\Omega_{n} satisfying the assumptions of Lemma 3.13 and also

Per(Ωn)2Per(Ω)anddiam(Ωn)2diam(Ω).\mathrm{Per}(\Omega_{n})\leq 2\,\mathrm{Per}(\Omega)\qquad\text{and}\qquad\mathrm{diam}(\Omega_{n})\leq 2\,\mathrm{diam}(\Omega).

Furthermore, since u1u_{1} is the limit of upu_{p}, we can always find a small positive number δ>0\delta>0 such that the following holds:

12u1L(Ω)upL(Ω)32u1L(Ω)for all p(1,1+δ1).\frac{1}{2}\|u_{1}\|_{L^{\infty}(\Omega)}\leq\|u_{p}\|_{L^{\infty}(\Omega)}\leq\frac{3}{2}\|u_{1}\|_{L^{\infty}(\Omega)}\qquad\text{for all }p\in(1,1+\delta_{1}).

Then the estimate above shows that

σpLr(Ω)Cfor all p(1,1+δ).\|\sigma_{p}\|_{L^{r}(\Omega)}\leq C\qquad\text{for all }p\in(1,1+\delta).

Finally, from the inequality Ωθpr𝑑xΩσpr𝑑x\int_{\Omega}\theta_{p}^{r}\,dx\leq\int_{\Omega}\sigma_{p}^{r}\,dx and applying once again the estimate

Ωσpr𝑑x(1+mL)r|Ω|+upL2q(Ω)2Bσpr|up|2𝑑x.\int_{\Omega}\sigma_{p}^{r}\,dx\leq(1+mL)^{r}|\Omega|+\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{-2}\int_{B}\sigma_{p}^{r}|\nabla u_{p}|^{2}\,dx.

the conclusion follows ∎

Remark 3.15.

The estimate (3.10) is not uniform with respect to rr since the constant CC depends on rr and, while it is true that we can write

u1Lr(Ω)(1+|Ω|)u1L(Ω),\|u_{1}\|_{L^{r}(\Omega)}\leq(1+|\Omega|)\|u_{1}\|_{L^{\infty}(\Omega)},

it is also easy to show that

limrC(r)=.\lim_{r\to\infty}C(r)=\infty.

The reason is that there is a term multiplied by a positive constant that is linear with respect to rr, namely

C(r)[62r1|Ω|Λr(r1)rupL2q(Ω)2r2upL(Ω)2r]1rΛras r.C(r)\simeq\left[6^{2r-1}|\Omega|\frac{\Lambda^{r}(r-1)^{r}}{\|\nabla u_{p}\|_{L^{2q}(\Omega)}^{2r-2}}\|u_{p}\|_{L^{\infty}(\Omega)}^{2r}\right]^{\frac{1}{r}}\simeq\Lambda r\qquad\text{as }r\to\infty.

Thus, even if we assume that Ω\Omega is convex, we cannot get rid of this term because it does not come out of the boundary part. In Section 3.5, we show an alternative approach to the problem that allows us to achieve r=r=\infty when Ω\Omega is convex, but it requires regularity results in optimal transport theory.

3.4 Proof of Theorem 2.4: LpL^{p}-regularity for 1p<1\leq p<\infty

We are now ready to use all the results we collected so far to give a proof of the main theorem except for (i) that will be proved in the next section.

Proof of Theorem 2.4.

For p>1p>1, let upW01,2q(Ω)u_{p}\in W_{0}^{1,2q}(\Omega) be the positive minimizer of FpF_{p} with fixed L2L^{2}-norm (equal to one) and let θp\theta_{p} be given by

θp(x)=L[|up|2(q1)(x)(Ω|up|2q𝑑x)1p].\theta_{p}(x)=L\left[|\nabla u_{p}|^{2(q-1)}(x)\left(\int_{\Omega}|\nabla u_{p}|^{2q}\,dx\right)^{-\frac{1}{p}}\right].

Since θp\theta_{p} is admissible, using Proposition 3.14 we can find a constant C>0C>0 such that for p>1p>1 small enough we have

mθpLp(Ω)=mLandθpLr(Ω)Cm\|\theta_{p}\|_{L^{p}(\Omega)}=mL\qquad\text{and}\qquad\|\theta_{p}\|_{L^{r}(\Omega)}\leq C

for some rdr\geq d. It follows that θp\theta_{p} is uniformly bounded in L2(Ω)L^{2}(\Omega) and hence, up to subsequences, it converges weakly to a nonnegative function θ¯L2(Ω)\bar{\theta}\in L^{2}(\Omega). Since

Ωθ¯𝑑x=limp1+Ωθp𝑑xlim infp1+θpLp(Ω)|Ω|1/q=L,\int_{\Omega}\bar{\theta}\,dx=\lim_{p\to 1^{+}}\int_{\Omega}\theta_{p}\,dx\leq\liminf_{p\to 1^{+}}\|\theta_{p}\|_{L^{p}(\Omega)}|\Omega|^{1/q}=L,

we easily infer that θ¯\bar{\theta} is admissible. On the other hand, Proposition 3.4 asserts that upu_{p} converges strongly in H01(Ω)H_{0}^{1}(\Omega) to the minimum u1u_{1} of F1F_{1}, which means that

(1+mθp)upφdxΩ(1+mθ¯)u1φdxas p1+\int(1+m\theta_{p})\nabla u_{p}\cdot\nabla\varphi\,dx\to\int_{\Omega}(1+m\bar{\theta})\nabla u_{1}\cdot\nabla\varphi\,dx\qquad\hbox{as }p\to 1^{+}

for every φCc(Ω)\varphi\in C_{c}^{\infty}(\Omega). Moreover, it is easy to check that

λ1(θp)Ωupφ𝑑xλ1(θ¯)Ωu1φ𝑑xas p1+,\lambda_{1}(\theta_{p})\int_{\Omega}u_{p}\varphi\,dx\to\lambda_{1}(\bar{\theta})\int_{\Omega}u_{1}\varphi\,dx\qquad\hbox{as }p\to 1^{+},

which implies that u1u_{1} is a solution of the equation

div((1+θ¯)u1)=λ1(θ¯)u1-\mathrm{div}((1+\bar{\theta})\nabla u_{1})=\lambda_{1}(\bar{\theta})u_{1}

for xΩx\in\Omega, with Dirichlet boundary condition on Ω\partial\Omega. An application of the integration by parts formula shows that

E(θ¯,u1):=Ω(1+θ¯)|u1|2𝑑x=λ(θ¯)Ω|u1|2𝑑x=1.E(\bar{\theta},u_{1}):=\int_{\Omega}(1+\bar{\theta})|\nabla u_{1}|^{2}\,dx=\lambda(\bar{\theta})\underbrace{\int_{\Omega}|u_{1}|^{2}\,dx}_{=1}.

The strong converge of upu_{p} to u1u_{1} in L2(Ω)L^{2}(\Omega) and Proposition 3.4 implies that

E(θ¯,u1)=limp1+λ(θp)=limp1+E(θp,up)=limp1+Fp(up)=F1(u1),\begin{split}E(\bar{\theta},u_{1})&=\lim_{p\to 1^{+}}\lambda(\theta_{p})\\ &=\lim_{p\to 1^{+}}E(\theta_{p},u_{p})=\lim_{p\to 1^{+}}F_{p}(u_{p})=F_{1}(u_{1}),\end{split}

which means that

E(θ¯,u1)=minuH01(Ω){0}F1(u).E(\bar{\theta},u_{1})=\min_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}F_{1}(u).

Finally the general min-max inequality shows that

supθ𝒜Lλ1(θ)=supθ𝒜LminuH01(Ω)E(θ,u)minuH01(Ω)supθ𝒜LE(θ,u)=minuH01(Ω)F1(u)=E(θ¯,u1),\begin{split}\sup_{\theta\in\mathscr{A}_{L}}\lambda_{1}(\theta)&=\sup_{\theta\in\mathscr{A}_{L}}\min_{u\in H_{0}^{1}(\Omega)}E(\theta,u)\\ &\leq\min_{u\in H_{0}^{1}(\Omega)}\sup_{\theta\in\mathscr{A}_{L}}E(\theta,u)=\min_{u\in H_{0}^{1}(\Omega)}F_{1}(u)=E(\bar{\theta},u_{1}),\end{split}

and this is enough to conclude that θ¯\bar{\theta} is a solution of the maximization problem (2.3) since it is an admissible competitor.

The regularity of the optimal density θ¯\bar{\theta} follows from the fact that θ¯Lr(Ω)\bar{\theta}\in L^{r}(\Omega) for rr arbitrarily large (since Ω\Omega is a bounded-volume set). It is now trivial to show that θ¯\bar{\theta} is equal to zero almost everywhere on the set

{xΩ:|uθ¯|(x)<uθ¯L(Ω)},\big{\{}x\in\Omega\ :\ |\nabla u_{\bar{\theta}}|(x)<\|\nabla u_{\bar{\theta}}\|_{L^{\infty}(\Omega)}\big{\}},

while (2.4) follows from the min-max inequality which gives an equality evaluated at the optimal couple. Finally, the assertion (ii) follows from the regularity of the minimizer u1u_{1} of F1F_{1} as in Proposition 3.1. ∎

3.5 Proof of Theorem 2.4: LL^{\infty}-regularity for Ω\Omega convex

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open set which is either convex or with a boundary of class C2,αC^{2,\alpha}. Instead of relaxing the maximization problem to (3.1), we can study

maxμ𝒜LinfuCc1(Ω){0}{Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑x},\max_{\mu\in\mathscr{A}_{L}}\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\right\}, (3.11)

where 𝒜L\mathscr{A}_{L} is the class defined in (2.3). Set

J(u,μ):={Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑x},J(u,\,\mu):=\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\right\},

and consider the functional

J(μ):=infuCc1(Ω){0}{Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑x}.J(\mu):=\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\right\}. (3.12)

Notice that there are no measures in 𝒜L\mathscr{A}_{L} such that J(μ)=J(\mu)=-\infty. This is a significant advantage over the energy problem and can be checked easily since

Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑xΩ|u|2𝑑xΩ|u|2𝑑xλ1(Ω)>0.\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\geq\frac{\int_{\Omega}|\nabla u|^{2}\,dx}{\int_{\Omega}|u|^{2}\,dx}\geq\lambda_{1}(\Omega)>0.
Proposition 3.16.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded open set. Then the maximization problem (3.11) admits a solution μ¯𝒜L\bar{\mu}\in\mathscr{A}_{L} with

sptμ{xΩ:|u¯(x)|=u¯L},\mathrm{spt}\mu\subset\{x\in\Omega\ :\ |\nabla\bar{u}(x)|=\|\nabla\bar{u}\|_{L^{\infty}}\}, (3.13)

where u¯\bar{u} is the unique positive function with fixed L2L^{2}-norm achieving the minimum in the functional (3.12). Furthermore, we have the identity

J(μ¯)=minuW01,(Ω¯){0}Ω|u|2𝑑x+mLuL(Ω)2Ω|u|2𝑑x.J(\bar{\mu})=\min_{u\in W_{0}^{1,\infty}(\bar{\Omega})\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|^{2}_{L^{\infty}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}. (3.14)
Proof.

First, we notice that for every fixed uCc1(Ω)u\in C_{c}^{1}(\Omega) the map μJ(u,μ)\mu\mapsto J(u,\mu) is continuous with respect to the weak* convergence. Hence, being J(μ)J(\mu) the infimum of continuous maps, it is weakly* upper semi-continuous. Since we observed already that 𝒜L\mathscr{A}_{L} is weakly* compact and nonempty we infer that (3.12) admits a solution μ¯\bar{\mu}. To prove the other claims, observe that

maxμ𝒜LinfuCc1(Ω){0}J(u,μ)infuCc1(Ω){0}supμ𝒜LJ(u,μ)\max_{\mu\in\mathscr{A}_{L}}\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}J(u,\mu)\leq\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}\sup_{\mu\in\mathscr{A}_{L}}J(u,\mu)

is always true, but the equality does not hold a priori since we lack concavity of JJ with respect to the first variable uu. Nevertheless, we have

supμ𝒜LJ(u,μ)=Ω|u|2𝑑x+mLuL(Ω)2Ω|u|2𝑑x,\sup_{\mu\in\mathscr{A}_{L}}J(u,\mu)=\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|^{2}_{L^{\infty}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx},

and it is easy to check that this is achieved using any measure μ\mu with total mass equal to LL and support satisfying (3.13). Denote any one of them by μ¯\bar{\mu} and notice that μ¯𝒜L\bar{\mu}\in\mathscr{A}_{L} implies

maxμ𝒜LinfuCc1(Ω){0}J(u,μ)infuCc1(Ω){0}Ω|u|2𝑑x+mΩ|u|2𝑑μ¯Ω|u|2𝑑x=infuCc1(Ω){0}Ω|u|2𝑑x+mLuL(Ω)2Ω|u|2𝑑x=infuCc1(Ω){0}supμ𝒜LJ(u,μ).\begin{split}\max_{\mu\in\mathscr{A}_{L}}\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}J(u,\mu)&\geq\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\bar{\mu}}{\int_{\Omega}|u|^{2}\,dx}\\ &=\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|_{L^{\infty}(\Omega)}^{2}}{\int_{\Omega}|u|^{2}\,dx}\\ &=\inf_{u\in C_{c}^{1}(\Omega)\setminus\{0\}}\sup_{\mu\in\mathscr{A}_{L}}J(u,\,\mu).\end{split}

This shows that we can interchange infimum and supremum and therefore the claim (3.14) holds, concluding the proof. ∎

Now that we know the existence of the optimal measure μ¯\bar{\mu}, we can equivalently investigate the minimization problem associated with the functional

J1(u)=Ω|u|2𝑑x+mLuL(Ω)2J_{1}(u)=\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|_{L^{\infty}(\Omega)}^{2}

satisfying the additional constraint Ω|u|2𝑑x=1\int_{\Omega}|u|^{2}\,dx=1. The proof of the next result follows immediately from [13]. Notice that what we find out here is compatible with the investigation we carried out in Section 3.

Theorem 3.17.

The optimization problem

min{J1(u):uH01(Ω),uL2(Ω)=1}\min\big{\{}J_{1}(u)\ :\ u\in H_{0}^{1}(\Omega),\ \|u\|_{L^{2}(\Omega)}=1\big{\}}

admits a unique solution u¯W2,p(Ω)\bar{u}\in W^{2,p}(\Omega) for all p>dp>d. If, in addition, Ω\Omega is convex, then u¯W2,(Ω)\bar{u}\in W^{2,\infty}(\Omega).

We can now show that the optimal measure μ¯\bar{\mu} belongs to LpL^{p} spaces for p<p<\infty arbitrary, as in Section 3.4, and show that we can obtain a uniform LL^{\infty} estimate if Ω\Omega convex concluding the proof of Theorem 2.4.

Proof of Theorem 2.4 (i).

Let μ¯\bar{\mu} be the optimal measure given by Proposition 3.16. A standard result in elliptic regularity theory (e.g, [5]) implies

u¯argmin(J1())u¯C2,β(Ω)\bar{u}\in\mathrm{argmin}(J_{1}(\cdot))\implies\bar{u}\in C^{2,\,\beta}(\Omega)

for some β>0\beta>0. Note that here we use the fact that Ω\Omega is either regular (C2,αC^{2,\,\alpha} and thus β\beta depends on α\alpha) or convex. By Theorem 3.17, we infer that

ΔuLp(Ω)for all p1.\Delta u\in L^{p}(\Omega)\qquad\text{for all $p\geq 1$}.

Thus u¯\bar{u} and μ¯\bar{\mu} solve the problem

{div(μ¯u¯)=Δu¯+λ1u¯,if xΩ,u¯|Ω0,|u¯(x)|=u¯L(Ω),if xspt(μ¯).\begin{cases}-\mathrm{div}(\bar{\mu}\nabla\bar{u})=\Delta\bar{u}+\lambda_{1}\bar{u},&\text{if $x\in\Omega$},\\[5.0pt] \bar{u}\,\big{|}_{\partial\Omega}\equiv 0,\\[5.0pt] |\nabla\bar{u}(x)|=\|\nabla\bar{u}\|_{L^{\infty}(\Omega)},&\text{if $x\in\mathrm{spt}(\bar{\mu})$}.\end{cases}

The right-hand side of the equation belongs to Lp(Ω)L^{p}(\Omega) so we deduce that also μ¯Lp(Ω)\bar{\mu}\in L^{p}(\Omega) for all p>dp>d using the regularity results for the Monge-Kantorovich problem given in [11, 12, 22]. Similarly, we have

Ω convexΔu¯L(Ω),\Omega\text{ convex}\implies\Delta\bar{u}\in L^{\infty}(\Omega),

and we can apply once more [11, 12, 22] to conclude that μ¯L(Ω)\bar{\mu}\in L^{\infty}(\Omega). ∎

Remark 3.18.

The approach via Γ\Gamma-convergence and the approach presented in this section are both needed to prove Theorem 2.4. Indeed, the LL^{\infty} estimate on θ¯\bar{\theta} is impossible to obtain via Γ\Gamma-convergence, even if we assume that Ω\Omega is convex. On the other hand, the Monge-Kantorovich approach requires Ω\Omega to be either convex or C2,αC^{2,\alpha} even for the higher integrability result θLp(Ω)\theta\in L^{p}(\Omega), which only requires the uniform external ball condition with Γ\Gamma-convergence.

4 The radial case

Let Ω\Omega be the unit disc of 2\mathbb{R}^{2}. In this section, we exploit the symmetries of the domain to show that the solution of (2.3) is a radially symmetric function with an explicit formula. First, we prove a technical lemma which allows us to use polar coordinates to deal with the min-max problem.

Lemma 4.1.

The optimal density θ¯\bar{\theta} solution of (2.3) and the corresponding optimal profile u¯:=uθ¯\bar{u}:=u_{\bar{\theta}} are both radially symmetric functions.

Proof.

Let upu_{p} be the function given in Proposition 3.1. Then upu_{p} is the unique solution (with fixed L2L^{2}-norm) of the minimization problem

minuH01(Ω){Ω|u|2𝑑x+mLuL2q(Ω)2Ω|u|2𝑑x}.\min_{u\in H_{0}^{1}(\Omega)}\left\{\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\|\nabla u\|_{L^{2q}(\Omega)}^{2}}{\int_{\Omega}|u|^{2}\,dx}\right\}.

Now recall that the Steiner symmetrization [20, Chapter 7] of a function uH01(Ω)u\in H_{0}^{1}(\Omega), denoted by uu^{\ast}, satisfies the Pólya-Szegö’s inequality

Ω|u|p𝑑xΩ|u|p𝑑x.\int_{\Omega^{\ast}}|\nabla u^{\ast}|^{p}\,dx\leq\int_{\Omega}|\nabla u|^{p}\,dx.

for all 1p<1\leq p<\infty (see [18, Theorem 2.2.4]). The unit ball is symmetric so it coincides with its symmetrization Ω\Omega^{\ast} and the L2L^{2}-norm of uu coincide with the one of uu^{\ast} so from the inequality

Ω|u|2𝑑x+mL|u|2Lq(Ω)Ω|u|2𝑑xΩ|u|2𝑑x+mL|u|2Lq(Ω)Ω|u|2𝑑x\frac{\int_{\Omega}|\nabla u^{\ast}|^{2}\,dx+mL\||\nabla u^{\ast}|^{2}\|_{L^{q}(\Omega)}}{\int_{\Omega}|u^{\ast}|^{2}\,dx}\leq\frac{\int_{\Omega}|\nabla u|^{2}\,dx+mL\||\nabla u|^{2}\|_{L^{q}(\Omega)}}{\int_{\Omega}|u|^{2}\,dx}

we infer that each upu_{p} is radially symmetric. On the other hand, we proved in Proposition 3.4 that upu_{p} converges strongly in L2L^{2} to u¯\bar{u} as p1p\to 1 and therefore, up to subsequences, we can assume that upu_{p} converges almost everywhere to u¯\bar{u}. Thus

u¯(x)=limp1+up(x)=limp1+up(|x|)u¯ radially symmetric.\bar{u}(x)=\lim_{p\to 1^{+}}u_{p}(x)=\lim_{p\to 1^{+}}u_{p}(|x|)\implies\text{$\bar{u}$ radially symmetric.}

Finally, if we choose u¯\bar{u} to be the solution with L2L^{2}-norm equal to one, then θ¯\bar{\theta} is the unique maximizer of the functional

θΩ(1+θ)u¯dx.\theta\longmapsto\int_{\Omega}(1+\theta)\nabla\bar{u}\,dx.

The function u¯\bar{u} is radial so we can apply (iv) of [18, Theorem 2.2.4] and obtain

Ω(1+θ)u¯dxΩ(1+θ)u¯dx,\int_{\Omega}(1+\theta)\nabla\bar{u}\,dx\leq\int_{\Omega}(1+\theta)^{\ast}\nabla\bar{u}\,dx,

which allows us to conclude that θ¯\bar{\theta} is also radially symmetric. ∎

Now fix d=2d=2 for simplicity and notice that the optimal profile u¯\bar{u} and the optimal density θ¯\bar{\theta} satisfy the elliptic equation

div((1+θ¯)u¯)=λ1u¯-\mathrm{div}((1+\bar{\theta})\nabla\bar{u})=\lambda_{1}\bar{u} (4.1)

so, exploiting the fact that they are both radially symmetric, we can write

1rr(r(1+θ¯(r))u¯(r))=λ1u¯(r).-\frac{1}{r}\partial_{r}\big{(}r(1+\bar{\theta}(r))\bar{u}^{\prime}(r)\big{)}=\lambda_{1}\bar{u}(r). (4.2)

Since by Theorem 2.4 the support of θ¯\bar{\theta} is contained in the set where |u¯||\nabla\bar{u}| achieves its maximum value, it is easy to see that there exists a¯(0, 1)\bar{a}\in(0,\,1) such that

u¯(r)=1rfor all r[a¯, 1].\bar{u}(r)=1-r\quad\text{for all $r\in[\bar{a},\,1]$}.

Notice that we can avoid placing a constant in front of (1r)(1-r) because we are using that u¯\bar{u} is unique when the L2L^{2}-norm is fixed (although we do not care about the actual value here). It follows from (4.2) that

1rr(r(1+θ¯(r)))=λ1(1r)for all r[a¯, 1],\frac{1}{r}\partial_{r}\big{(}r(1+\bar{\theta}(r))\big{)}=\lambda_{1}(1-r)\quad\text{for all $r\in[\bar{a},\,1]$},

and this leads to a ordinary differential equation in θ\theta which admits an explicit solution that depends on a¯\bar{a} and λ1\lambda_{1}, that is,

θ(r)={0if r[0,a¯],λ13r2+λ12r1+a¯r(1+λ13a¯2λ12a¯)if r[a¯, 1].\theta(r)=\begin{cases}0&\text{if $r\in[0,\,\bar{a}]$},\\[6.00006pt] -\frac{\lambda_{1}}{3}r^{2}+\frac{\lambda_{1}}{2}r-1+\frac{\bar{a}}{r}\left(1+\frac{\lambda_{1}}{3}\bar{a}^{2}-\frac{\lambda_{1}}{2}\bar{a}\right)&\text{if $r\in[\bar{a},\,1]$}.\end{cases}

Suppose that the value of λ1\lambda_{1} is known. We can find the optimal value a¯\bar{a} by exploiting the integral condition on θ\theta. Namely, we know that Ωθ𝑑x=L\int_{\Omega}\theta\,dx=L so

a¯1θ(r)r𝑑r=L2πλ1(a¯)=12(L2π+12(a¯1)216a¯2+8a¯33a¯4).\int_{\bar{a}}^{1}\theta(r)r\,dr=\frac{L}{2\pi}\implies\lambda_{1}(\bar{a})=12\left(\frac{\frac{L}{2\pi}+\frac{1}{2}(\bar{a}-1)^{2}}{1-6\bar{a}^{2}+8\bar{a}^{3}-3\bar{a}^{4}}\right).

This leads to

λ1=(L2π+12(a¯1)216a¯2+8a¯33a¯4)a¯=f1(λ1),\lambda_{1}=\left(\frac{\frac{L}{2\pi}+\frac{1}{2}(\bar{a}-1)^{2}}{1-6\bar{a}^{2}+8\bar{a}^{3}-3\bar{a}^{4}}\right)\implies\bar{a}=f^{-1}(\lambda_{1}), (4.3)

which admits a unique solution in the interval (0, 1)(0,\,1) provided that λ1\lambda_{1} is bigger than or equal to the minimum of ff, which is true for the one given in (4.1).

Remark 4.2.

In the energy problem with f=1f=1, one can prove (see Example 5.1 of [8]) that a¯\bar{a} is the unique solution of the polynomial equation

ad+1(d+1)(1+mLωd)a+d=0.a^{d+1}-(d+1)\left(1+\frac{mL}{\omega_{d}}\right)a+d=0.

Observe that for L=0L=0 the unique solution of the equation is a¯=1\bar{a}=1, and this is compatible with the fact that there is no reinforcement at all. The same is true in the eigenvalue case, but it cannot be inferred from (4.3) since the relation holds only when a density appears.

We can now recover u¯\bar{u}, the optimal profile, completely using the boundary condition naturally arising from the decomposition and the Neumann condition at the origin. Namely, it is easy to verify that

u¯(r)=c1J0(λ1r)+c2Y0(λ1r)for r[0,a¯],\bar{u}(r)=c_{1}J_{0}(\sqrt{\lambda_{1}}r)+c_{2}Y_{0}(\sqrt{\lambda_{1}}r)\quad\text{for $r\in[0,\,\bar{a}]$},

where J0J_{0} and Y0Y_{0} are the first Bessel functions of first and second kind respectively. To find the constants we simply notice that by continuity

c1J0(a¯λ1)+c2Y0(a¯λ1)=1a¯c_{1}J_{0}(\bar{a}\sqrt{\lambda_{1}})+c_{2}Y_{0}(\bar{a}\sqrt{\lambda_{1}})=1-\bar{a}

and, similarly, the Neumann condition gives

limr0+[c1J1(λ1r)+c2Y1(λ1r)]=0.\lim_{r\to 0^{+}}\left[c_{1}J_{1}(\sqrt{\lambda_{1}}r)+c_{2}Y_{1}(\sqrt{\lambda_{1}}r)\right]=0.

Since limr0+Y1(r)=\lim_{r\to 0^{+}}Y_{1}(r)=-\infty and limr0+J1(r)=0\lim_{r\to 0^{+}}J_{1}(r)=0, the second condition is satisfied if and only if c2=0c_{2}=0. This immediately shows that

c1=1a¯J0(a¯λ1),c_{1}=\frac{1-\bar{a}}{J_{0}(\bar{a}\sqrt{\lambda_{1}})},

and therefore the optimal profile is given by

u¯(r)={1a¯J0(a¯λ1)J0(λ1r)if r[0,a¯],1rif r[a¯, 1].\bar{u}(r)=\begin{cases}\frac{1-\bar{a}}{J_{0}(\bar{a}\sqrt{\lambda_{1}})}J_{0}(\sqrt{\lambda_{1}}r)&\text{if $r\in[0,\,\bar{a}]$},\\[6.00006pt] 1-r&\text{if $r\in[\bar{a},\,1]$}.\end{cases}

We now show the shape of the optimal density θ¯\bar{\theta} (see Figure 1) via a numerical analysis. The general idea is to fix λ1\lambda_{1} admissible (bigger than j0,02j_{0,0}^{2}) and recover the unique a¯(0, 1)\bar{a}\in(0,\,1) from the minimum problem

mina(0, 1)[λ11a¯J1(a¯λ1)]20ar[J0(λ1r)]2𝑑r+12(1a2)+mL2π[1a¯J0(a¯λ1)]20ar[J0(λ1r)]2𝑑r+a1r(1r)2𝑑r.\min_{a\in(0,\,1)}\frac{\left[\sqrt{\lambda_{1}}\frac{1-\bar{a}}{J_{1}(\bar{a}\sqrt{\lambda_{1}})}\right]^{2}\int_{0}^{a}r\left[J_{0}(\sqrt{\lambda_{1}}r)\right]^{2}\,dr+\frac{1}{2}(1-a^{2})+\frac{mL}{2\pi}}{\left[\frac{1-\bar{a}}{J_{0}(\bar{a}\sqrt{\lambda_{1}})}\right]^{2}\int_{0}^{a}r\left[J_{0}(\sqrt{\lambda_{1}}r)\right]^{2}\,dr+\int_{a}^{1}r(1-r)^{2}\,dr}.

Finally the length LL is determined starting from the identity (4.3). The numerical simulation confirms the regularity result in Theorem 2.4 because a¯(0, 1)\bar{a}\in(0,\,1) turns out to be the unique one for which

(1a¯)λJ0(a¯λ1)J1(λ1r)=limra¯u¯(r)=limra¯+u¯(r)=1-\frac{(1-\bar{a})\sqrt{\lambda}}{J_{0}(\bar{a}\sqrt{\lambda_{1}})}J_{1}(\sqrt{\lambda_{1}}r)=\lim_{r\to\bar{a}^{-}}\bar{u}^{\prime}(r)=\lim_{r\to\bar{a}^{+}}\bar{u}^{\prime}(r)=-1

holds. We would like to point out the main difference with the energy problem. In Example 5.1 of [8] it was proved that the optimal density is linear,

θ¯f(r)=(ra¯1)+r[0,1],\bar{\theta}_{f}(r)=\Big{(}\frac{r}{\bar{a}}-1\Big{)}^{+}\qquad\forall r\in[0,1],

while in our case the optimal density is not linear (it depends on r2r^{2} and r1r^{-1}) and, coherently with this dependence, it is not strictly increasing but rather

r¯(a¯, 1):θ|(a¯,r¯) increasing and θ|(r¯, 1) decreasing.\exists\,\bar{r}\in(\bar{a},\,1)\>:\>\text{$\theta\,\big{|}_{(\bar{a},\,\bar{r})}$ increasing and $\theta\,\big{|}_{(\bar{r},\,1)}$ decreasing}.
Refer to caption
Refer to caption
Figure 1: The picture on the left represents the levels sets of θ\theta on B(0, 1)B(0,\,1). We used λ1=10\lambda_{1}=10 and m=5m=5 and obtained a¯0.244419\bar{a}\approx 0.244419 and L0.424242L\approx 0.424242. The function θ\theta is increasing up to r0.751491r\approx 0.751491 and then it starts decreasing up to the border.

5 The connected case

We now consider the maximization problem (2.2), in which SS ranges in the class of closed, connected, one-dimensional subsets of Ω¯\bar{\Omega}. We follow closely the method introduced in [2] for the same optimal reinforcement problem when an external force acts on the membrane Ω\Omega, and we show that a small modification of the main proof is enough to get the same conclusion in the eigenvalues’ problem.

5.1 Proof of Theorem 2.6

In Proposition 2.5 we proved that there exists a solution μ\mu, in the class 𝒜Lc\mathscr{A}_{L}^{c}, to the maximization problem (2.5). It only remains to show that there is some θL1(Ω)\theta\in L^{1}(\Omega) such that μ=θ1  S\mu=\theta\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S. The following technical result was proved in [2, Lemma 3.3].

Lemma 5.1.

Let KK be a compact set in 2\mathbb{R}^{2} with |K|=0|K|=0. For all ε>0\varepsilon>0 there exists a function ϕε\phi_{\varepsilon} of class CC^{\infty} satisfying the following properties:

  1. (1)

    ϕε\phi_{\varepsilon} is locally constant on KK;

  2. (2)

    |ϕε(x)x|ε|\phi_{\varepsilon}(x)-x|\leq\varepsilon at all points x2x\in\mathbb{R}^{2};

  3. (3)

    |ϕε(x)|1|\nabla\phi_{\varepsilon}(x)|\leq 1 for all x2x\in\mathbb{R}^{2} and |ϕε(x)|=1|\nabla\phi_{\varepsilon}(x)|=1 everywhere except in an open set AεA_{\varepsilon} of measure less than ε\varepsilon containing KK.

We shall now prove that the optimal measure is absolutely continuous with respect to 1\mathcal{H}^{1}. In [2, Lemma 3.4] the same result is obtained in the energy problem, so it is sufficient to estimate the denominator and conclude in the same way.

Proof of Theorem 2.6.

The main difference with the mentioned paper is that we will show that for all vCc(Ω)v\in C_{c}^{\infty}(\Omega) and ε>0\varepsilon>0 there exists uH01(Ω)u\in H_{0}^{1}(\Omega) such that

Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑xΩ|v|2𝑑x+mΩ|v|2𝑑μa+C1εΩ|v|2𝑑xC2ε,\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\leq\frac{\int_{\Omega}|\nabla v|^{2}\,dx+m\int_{\Omega}|\nabla v|^{2}\,d\mu^{a}+C_{1}\varepsilon}{\int_{\Omega}|v|^{2}\,dx-C_{2}\varepsilon},

where μa\mu^{a} is the absolutely continuous part of μ\mu. Indeed, when ε>0\varepsilon>0 is small enough we can use Taylor approximation theorem to infer that

Ω|u|2𝑑x+mΩ|u|2𝑑μΩ|u|2𝑑xΩ|v|2𝑑x+mΩ|v|2𝑑μaΩ|v|2𝑑x+Cε,\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{\Omega}|\nabla u|^{2}\,d\mu}{\int_{\Omega}|u|^{2}\,dx}\leq\frac{\int_{\Omega}|\nabla v|^{2}\,dx+m\int_{\Omega}|\nabla v|^{2}\,d\mu^{a}}{\int_{\Omega}|v|^{2}\,dx}+C^{\prime}\varepsilon,

where C=C1/C2C^{\prime}=C_{1}/C_{2} is a positive constant. Now let

u(x):=(1θε(d(x)))v(x)+θε(d(x))v(ϕε(x)),u(x):=(1-\theta_{\varepsilon}(d(x)))v(x)+\theta_{\varepsilon}(d(x))v(\phi_{\varepsilon}(x)),

where d(x):=d(x,S)d(x):=d(x,S) and θ\theta is a cut-off function identically one on B(S,ε)B(S,\varepsilon) and zero on the complement of B(S,2ε)B(S,2\varepsilon). In [2, Lemma 3.4] it was proved that

12Ω|u|2𝑑x+m2Ω|u|2𝑑μC12ε+12Ω|v|2𝑑x+m2Ω|v|2𝑑μa,\frac{1}{2}\int_{\Omega}|\nabla u|^{2}\,dx+\frac{m}{2}\int_{\Omega}|\nabla u|^{2}\,d\mu\leq\frac{C_{1}}{2}\varepsilon+\frac{1}{2}\int_{\Omega}|\nabla v|^{2}\,dx+\frac{m}{2}\int_{\Omega}|\nabla v|^{2}\,d\mu^{a},

so we only have to deal with the denominator. A straightforward computation, assuming that vL2(Ω)2=1\|v\|_{L^{2}(\Omega)}^{2}=1, shows that

uL2(Ω)2=Ωθε2(d(x))[v(x)(v(x)v(ϕε(x)))+v(ϕε(x))(v(ϕε(x))v(x))]𝑑x++2Ωθε(d(x))v(x)(v(ϕε(x))v(x))𝑑x+1,\begin{split}\|u\|_{L^{2}(\Omega)}^{2}&=\int_{\Omega}\theta_{\varepsilon}^{2}(d(x))\left[v(x)(v(x)-v(\phi_{\varepsilon}(x)))+v(\phi_{\varepsilon}(x))(v(\phi_{\varepsilon}(x))-v(x))\right]\,dx+\\ &+2\int_{\Omega}\theta_{\varepsilon}(d(x))v(x)(v(\phi_{\varepsilon}(x))-v(x))\,dx+1,\end{split}

which immediately leads to the following estimate. If we denote A=|uL2(Ω)21|A=\left|\|u\|_{L^{2}(\Omega)}^{2}-1\right|, then we find that

AΩθε2(d(x))[|v(x)||v(x)v(ϕε(x))|+|v(ϕε(x))||v(ϕε(x))v(x)|]𝑑x+2Ω|θε(d(x))||v(x)||v(ϕε(x))v(x)|𝑑xCεΩ(|v(x)|+|v(ϕε(x))|)𝑑x+2CεΩ|v(x)|𝑑x4CvL1(Ω)ε.\begin{split}A&\leq\int_{\Omega}\theta_{\varepsilon}^{2}(d(x))\big{[}|v(x)|\left|v(x)-v(\phi_{\varepsilon}(x))\right|+|v(\phi_{\varepsilon}(x))|\left|v(\phi_{\varepsilon}(x))-v(x)\right|\big{]}\,dx\\ &\qquad+2\int_{\Omega}|\theta_{\varepsilon}(d(x))||v(x)|\left|v(\phi_{\varepsilon}(x))-v(x)\right|\,dx\\ &\leq C\varepsilon\int_{\Omega}(|v(x)|+|v(\phi_{\varepsilon}(x))|)\,dx+2C\varepsilon\int_{\Omega}|v(x)|\,dx\\ &\leq 4C\|v\|_{L^{1}(\Omega)}\,\varepsilon.\end{split}

Therefore, to conclude the proof, it suffices to set C2:=4CvL1(Ω)C_{2}:=4C\|v\|_{L^{1}(\Omega)} and continue as in [2, Lemma 3.4]. ∎

5.2 Indirect method and boundary points

Let S𝒜LcS\in\mathcal{A}_{L}^{c} and let uu be a solution of the minimization problem

λ1(S):=minuH01(Ω){0}Ω|u|2𝑑x+mSθ|τu|2𝑑1Ω|u|2𝑑x.\lambda_{1}(S):=\min_{u\in H_{0}^{1}(\Omega)\setminus\{0\}}\frac{\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{S}\theta|\nabla_{\tau}u|^{2}\,d\mathcal{H}^{1}}{\int_{\Omega}|u|^{2}\,dx}. (5.1)

Then for each vH01(Ω)v\in H_{0}^{1}(\Omega) there results

ddε|ε=0Ω|(u+εv)|2𝑑x+mSθ|τ(u+εv)|2𝑑1Ω|u+εv|2𝑑x=0,\frac{d}{d\varepsilon}\Big{|}_{\varepsilon=0}\frac{\int_{\Omega}|\nabla(u+\varepsilon v)|^{2}\,dx+m\int_{S}\theta|\nabla_{\tau}(u+\varepsilon v)|^{2}\,d\mathcal{H}^{1}}{\int_{\Omega}|u+\varepsilon v|^{2}\,dx}=0,

which is easily seen to be equivalent to

Ωu2𝑑x[Ωuvdx+mSθ(τuτv)𝑑1]Ωuv𝑑x[Ω|u|2𝑑x+mSθ|τu|2𝑑1]=0.\begin{split}\int_{\Omega}u^{2}\,dx&\left[\int_{\Omega}\nabla u\cdot\nabla v\,dx+m\int_{S}\theta(\nabla_{\tau}u\cdot\nabla_{\tau}v)\,d\mathcal{H}^{1}\right]\\ &-\int_{\Omega}uv\,dx\left[\int_{\Omega}|\nabla u|^{2}\,dx+m\int_{S}\theta|\nabla_{\tau}u|^{2}\,d\mathcal{H}^{1}\right]=0.\end{split}

Since uu minimizes the functional in (5.1) we can substitute it with λ1(S)\lambda_{1}(S) to obtain the following identity which is valid for all vH01(Ω)v\in H_{0}^{1}(\Omega),

Ωuvdx+mSθ(τuτv)𝑑1λ1(S)Ωuv𝑑x=0.\int_{\Omega}\nabla u\cdot\nabla v\,dx+m\int_{S}\theta(\nabla_{\tau}u\cdot\nabla_{\tau}v)\,d\mathcal{H}^{1}-\lambda_{1}(S)\int_{\Omega}uv\,dx=0.

The integration by parts formula shows that

Ωuvdx=ΩΔuv𝑑x+S[uν]v𝑑1,\int_{\Omega}\nabla u\cdot\nabla v\,dx=-\int_{\Omega}\Delta uv\,dx+\int_{S}\left[\frac{\partial u}{\partial\nu}\right]v\,d\mathcal{H}^{1},

where [uν]:=+u+u\left[\frac{\partial u}{\partial\nu}\right]:=\partial_{+}u+\partial_{-}u does not depend on the choice of an orientation and ±u\partial_{\pm}u are, respectively, the positive and negative derivatives of uu on SS. Finally, since

τuτv=(uτ)τ(vτ)τ=(uτ)(vτ),\nabla_{\tau}u\cdot\nabla_{\tau}v=(\nabla u\cdot\tau)\tau\cdot(\nabla v\cdot\tau)\tau=(\nabla u\cdot\tau)\cdot(\nabla v\cdot\tau),

we can integrate by parts the second term and obtain

mSθτuτvd1=mSdivτ(θτu)v𝑑1+m[vθτu]S#,m\int_{S}\theta\nabla_{\tau}u\cdot\nabla_{\tau}v\,d\mathcal{H}^{1}=-m\int_{S}\mathrm{div}_{\tau}(\theta\nabla_{\tau}u)v\,d\mathcal{H}^{1}+m\left[v\theta\nabla_{\tau}u\right]_{S^{\char 35\relax}},

where divτ(τ)-\mathrm{div}_{\tau}(-\nabla_{\tau}) is the Laplace-Beltrami operator on SS, and S#S^{\char 35\relax} is the set of terminal-type and branching-type points of SS.

Proposition 5.2.

If uH01(Ω)C2(Ω¯)u\in H_{0}^{1}(\Omega)\cap C^{2}(\bar{\Omega}) is a minimum point of (5.1), then it solves the following boundary-value problem:

{Δu=λ1(S)uif xΩS,u=0if xΩ,[uν]mdivτ(θτu)=0if xS,[θτu]S#=0.\begin{cases}-\Delta u=\lambda_{1}(S)u&\text{if $x\in\Omega\setminus S$},\\ u=0&\text{if $x\in\partial\Omega$,}\\ \left[\frac{\partial u}{\partial\nu}\right]-m\,\mathrm{div}_{\tau}(\theta\nabla_{\tau}u)=0&\text{if $x\in S$},\\[6.00006pt] \big{[}\theta\nabla_{\tau}u\big{]}_{S^{\char 35\relax}}=0.\end{cases}

For points in S#S^{\char 35\relax} the following three situations are all possible and, in the general case, we expect all of them to appear:

  1. (i)

    Dirichlet. If xS#Ωx\in S^{\char 35\relax}\cap\partial\Omega, then u(x)=0u(x)=0.

  2. (ii)

    Neumann. If xS#x\in S^{\char 35\relax} is a terminal point of SS, then τu(x)=0\nabla_{\tau}u(x)=0.

  3. (iii)

    Kirchhoff. If xS#x\in S^{\char 35\relax} is a branching point of SS, then

    iτiui(x)=0,\sum_{i}\nabla_{\tau_{i}}u^{i}(x)=0,

    where uiu^{i} is the trace of uu over the ii-th branch of SS ending at xx and τi\tau_{i} the corresponding tangent vector.

As a consequence, using Proposition 5.2 and Theorem 2.6, it is easy to verify that [2, Proposition 4.1] can be proved in the same way.

Proposition 5.3.

Let μ\mu be a solution of (2.5) and let uu be the unique positive solution of the associated minimization problem with L2L^{2}-norm fixed. Then there exists a positive constant cc such that

{|τu|=cfor 1-a.e. x{θ(x)>1},|τu|cfor 1-a.e. x{θ(x)=1}.\begin{cases}|\nabla_{\tau}u|=c&\text{for $\mathcal{H}^{1}$-a.e. $x\in\{\theta(x)>1\}$},\\ |\nabla_{\tau}u|\leq c&\text{for $\mathcal{H}^{1}$-a.e. $x\in\{\theta(x)=1\}$}.\end{cases}

6 Conclusion and open problems

In this section we present and discuss some open questions related to the optimization problems we considered.

Problem 1.

A first problem is related to the regularity of solutions. We have shown (Theorem 2.4) that when Ω\Omega is regular enough the optimization problem (2.3) admits a solution μ\mu that is indeed a function θC1,β(Ω¯)\theta\in C^{1,\beta}(\bar{\Omega}) for a suitable β(0,1)\beta\in(0,1). It would be interesting to know if additional regularity properties on θ\theta hold in general. Similarly, the optimization problem (2.5) admits a solution μ\mu which is of the form θ1  S\theta\mathcal{H}^{1}\mathbin{\vrule height=6.88889pt,depth=0.0pt,width=0.55974pt\vrule height=0.55974pt,depth=0.0pt,width=5.59721pt}S for a suitable closed connected set SΩ¯S\subset\bar{\Omega} and a function θL1(S)\theta\in L^{1}(S), with (S)L\mathcal{L}(S)\leq L and θ1\theta\geq 1 on SS. Even if we could expect that SS and θ\theta are regular enough, at the moment these regularity results are not available and seem rather difficult. In particular, it would be interesting to prove (or disprove) the regularity of the optimal set SS up to a finite number of branching points, where the Kirchhoff rule iτiu=0\sum_{i}\nabla_{\tau_{i}}u=0 holds.

Problem 2.

For the optimal set SS of problem (2.5) several necessary conditions of optimality merit to be investigated, for instance we list the following ones, that look similar to other problems studied in the fields of optimal transport and of structural mechanics (see [7], [9], [10]).

  • (a)

    Does SS contain closed loops (i.e. subsets homeomorphic to the circle S1S^{1})? This should not be the case, even if a complete proof is missing.

  • (b)

    The branching points of SS (if any) do have only three branches or a higher number of branches is possible?

  • (c)

    Does the optimal set SS intersect always the boundary Ω\partial\Omega?

  • (d)

    Is it possible that (S)=L\mathcal{L}(S)=L or we always have (S)<L\mathcal{L}(S)<L and hence θ>1\theta>1 somewhere on SS?

Problem 3.

As stated in the Introduction, passing from a single connected set SS to sets with at most a NN connected components (with NN a priori fixed) does not introduce essential differences in the statements and in the proofs. However, it would be interesting to establish if, in the case when NN connected components are allowed, the optimal set SS has actually exactly NN components.

Problem 4.

Finally, the numerical treatment of the optimization problems we considered, present several difficulties, essentially due to the fact that a very large number of local maxima are possible and global optimization algorithms are usually too slow for this kind of problems. In the case of energy optimization, considered in [2] and in [8], some efficient optimization methods have been implemented, but the eigenvalue optimization considered in the present paper seems to present a higher level of complexity.

Acknowledgements

The work of the first author is part of the project 2017TEXA3H “Gradient flows, Optimal Transport and Metric Measure Structures” funded by the Italian Ministry of Research and University. The first author is members of the “Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni” (GNAMPA) of the “Istituto Nazionale di Alta Matematica” (INDAM).

References

  • [1]
  • [2] G. Alberti, G. Buttazzo, S. Guarino Lo Bianco, E. Oudet: Optimal reinforcing networks for elastic membranes. Netw. Heterog. Media, 14 (3) (2019), 589–615.
  • [3] L. Ambrosio, P. Tilli: Selected topics on analysis in metric spaces. Oxford Lecture Series in Mathematics and its Applications 25, Oxford University Press, Oxford, 2004.
  • [4] G. Bouchitté, G. Buttazzo, P. Seppecher: Energies with respect to a measure and applications to low dimensional structures. Calc. Var. Partial Differential Equations, 5 (1996), 37–54.
  • [5] H. Brezis, G. Stampacchia: Sur la régularité de la solution d’inéquations elliptiques. Bull. Soc. Math. France, 96 (1968), 153–180.
  • [6] D. Bucur, G. Buttazzo: Variational methods in shape optimization problems. Progress in Nonlinear Differential Equations and Their Applications 65, Birkhäuser, Basel, 2005.
  • [7] G. Buttazzo, E. Oudet, E. Stepanov: Optimal transportation problems with free Dirichlet regions. In “Variational Methods for Discontinuous Structures”, Progr. Nonlinear Differential Equations Appl. 51, Birkhäuser, Basel, 2002, 41–65.
  • [8] G. Buttazzo, E. Oudet, B. Velichkov: A free boundary problem arising in PDE optimization. Calc. Var. Partial Differential Equations, 54 (2015), 3829–3856.
  • [9] G. Buttazzo, E. Stepanov: Optimal transportation networks as free Dirichlet regions for the Monge-Kantorovich problem. Ann. Sc. Norm. Super. Pisa Cl. Sci., 2 (4) (2003), 631–678.
  • [10] A. Chambolle, J. Lamboley, A. Lemenant, E. Stepanov: Regularity for the optimal compliance problem with length penalization. SIAM J. Math. Anal., 49 (2) (2017), 1166–1224.
  • [11] L. De Pascale, L.C. Evans, A. Pratelli: Integral estimates for transport densities. Bull. London Math. Soc., 36 (2004), 383–395.
  • [12] L. De Pascale, A. Pratelli: Regularity properties for the Monge transport density and for solutions of some shape optimization problem. Calc. Var. Partial Differential Equations, 14 (2002), 249–274.
  • [13] L.C. Evans: A second order elliptic equation with gradient constraint. Calc. Var. Partial Differential Equations, 4 (1979), 555–572.
  • [14] L.C. Evans: Partial Differential Equations. Graduate Studies in Mathematics 19, American Mathematical Society, Providence, 2010.
  • [15] K.J. Falconer: The geometry of fractal sets. Cambridge Tracts in Mathematics 85, Cambridge University Press, Cambridge, 1986.
  • [16] D. Gilbarg, N.S. Trudinger: Elliptic partial differential equations of second order. Classics in Mathematics, Springer-Verlag, Berlin, 2001.
  • [17] S. Gołab: Sur quelques points de la théorie de la longueur (On some points of the theory of the length). Ann. Soc. Polon. Math., 7 (1929), 227–241.
  • [18] A. Henrot: Extremum Problems for Eigenvalues of Elliptic Operators. Frontiers in Mathematics, Birkhäuser, Basel, 2006.
  • [19] A. Henrot, M. Pierre: Shape Variation and Optimization: A Geometric Analysis. EMS Tracts in Mathematics 28, European Mathematical Society, Zürich, 2018.
  • [20] S.G. Krantz, H.R. Parks: The Geometry of Domains in Space. Birkhäuser Advanced Texts, Birkhäuser, Basel, 1999.
  • [21] G.M. Lieberman: Boundary regularity for solutions of degenerate elliptic equations. Nonlinear Anal. Theory Methods Appl., 12 (1988), 1203–1219.
  • [22] F. Santambrogio: Absolute continuity and summability of transport densities: simpler proofs and new estimates. Calc. Var. Partial Differential Equations, 36 (2009), 343–354.

Giuseppe Buttazzo: Dipartimento di Matematica, Università di Pisa
Largo B. Pontecorvo 5, 56127 Pisa - ITALY
giuseppe.buttazzo@unipi.it
http://www.dm.unipi.it/pages/buttazzo/


Francesco Paolo Maiale: Scuola Normale Superiore
Piazza dei Cavalieri 7, 56126 Pisa - ITALY
francesco.maiale@sns.it
https://poisson.phc.dm.unipi.it/~fpmaiale/