We discuss two optimization problems related to the fractional p-Laplacian. First, we prove the existence of at least one minimizer for the principal eigenvalue of the fractional p-Laplacian with Dirichlet conditions, with a bounded weight function varying in a rearrangement class. Then, we investigate the minimization of the energy functional for general nonlinear equations driven by the same operator, as the reaction varies in a rearrangement class. In both cases, we provide a pointwise relation between the optimizing datum and the corresponding solution.
Citation: Antonio Iannizzotto, Giovanni Porru. Optimization problems in rearrangement classes for fractional p-Laplacian equations[J]. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002
[1] | Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic $ p $-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015 |
[2] | Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the $ p $-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017 |
[3] | Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082 |
[4] | María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715 |
[5] | Daniela De Silva, Giorgio Tortone . Improvement of flatness for vector valued free boundary problems. Mathematics in Engineering, 2020, 2(4): 598-613. doi: 10.3934/mine.2020027 |
[6] | Peter Bella, Mathias Schäffner . Local boundedness for $ p $-Laplacian with degenerate coefficients. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023081 |
[7] | Catharine W. K. Lo, José Francisco Rodrigues . On the obstacle problem in fractional generalised Orlicz spaces. Mathematics in Engineering, 2024, 6(5): 676-704. doi: 10.3934/mine.2024026 |
[8] | Nicola Abatangelo, Sven Jarohs, Alberto Saldaña . Fractional Laplacians on ellipsoids. Mathematics in Engineering, 2021, 3(5): 1-34. doi: 10.3934/mine.2021038 |
[9] | Stefano Borghini . Symmetry results for Serrin-type problems in doubly connected domains. Mathematics in Engineering, 2023, 5(2): 1-16. doi: 10.3934/mine.2023027 |
[10] | David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036 |
We discuss two optimization problems related to the fractional p-Laplacian. First, we prove the existence of at least one minimizer for the principal eigenvalue of the fractional p-Laplacian with Dirichlet conditions, with a bounded weight function varying in a rearrangement class. Then, we investigate the minimization of the energy functional for general nonlinear equations driven by the same operator, as the reaction varies in a rearrangement class. In both cases, we provide a pointwise relation between the optimizing datum and the corresponding solution.
The present paper deals with some optimization problems related to elliptic equations of nonlinear, nonlocal type, with data varying in rearrangement classes. For the reader's convenience, we recall here the basic definition, referring to Section 2 for details. Given a bounded smooth domain Ω⊂RN and a non-negative function g0∈L∞(Ω), we say that g∈L∞(Ω) lies in the rearrangement class of g0, denoted G, if for all t⩾0
|{g>t}|=|{g0>t}|, |
where we denote by |⋅| the N-dimensional Lebesgue measure of sets. We may define several functionals Φ:G→R corresponding to variational problems, and study the optimization problems
ming∈GΦ(g),maxg∈GΦ(g). |
We note that, since G is not a convex set, the problems above do not fall in the familiar case of convex optimization, whatever the nature of Φ. The following is a classical example. For all g∈G consider the Dirichlet problem
{−Δu=g(x)in Ωu=0on ∂Ω, |
which, by classical results in the calculus of variations, admits a unique weak solution ug∈H10(Ω). So set
Φ(g)=∫Ωgugdx. |
The existence of a maximizer for Φ, i.e., of a datum ˆg∈G s.t. for all g∈G
Φ(ˆg)⩾Φ(g), |
was proved in [3,4], while the existence of a minimizer was investigated in [5]. One challenging feature of such problem is that, in general, the functional Φ turns out to be continuous (in a suitable sense) but the class G fails to be compact. Therefore, a possible strategy consists in optimizing Φ over the closure ¯G of G in the sequential weak* topology of L∞(Ω) (a much larger, and convex, set), and then proving that the maximizers and minimizers actually lie in G (which is far from being trivial).
In addition, due to the nature of the rearrangement equivalence and some functional inequalities, the maximizer/minimizer g may show some structural connection to the solution ug of the corresponding variational problem. This has interesting consequences, for instance let g0 be the characteristic function of some subdomain D0⊂Ω, then the optimal g is as well the characteristic function of some D⊂Ω with |D|=|D0|. Moreover, it is proved that g=η∘ug in Ω for some nondecreasing η, while by the Dirichlet condition we have ug=0 on ∂Ω. Therefore, any optimal domain D has a positive distance from ∂Ω.
A similar approach applies to several variational problems and functionals. For instance, in [8] the authors consider the following p-Laplacian equation with p>1, q∈[0,p):
{−Δpu=g(x)uq−1in Ωu=0on ∂Ω, |
which admits a unique non-negative solution ug∈W1,p0(Ω), and study the maximum and minimum over G of the functional
Φ(g)=∫Ωguqgdx. |
In [7], the following weighted eigenvalue problem is considered:
{−Δpu=λg(x)|u|p−2uin Ωu=0on ∂Ω. |
It is well known that the problem above admits a principal eigenvalue λ(g)>0 (see [20]), and the authors prove that the functional λ(g) has a minimizer in G.
In recent years, several researchers have studied optimization problems related to elliptic equations of fractional order (see [23] for a general introduction to such problems and the related variational methods). In the linear framework, the model operator is the s-fractional Laplacian with s∈(0,1), defined by
(−Δ)su(x)=CN,slimε→0+∫Bcε(x)u(x)−u(y)|x−y|N+2sdy, |
where CN,s>0 is a normalization constant. In [26], the following problem is examined:
{(−Δ)su+h(x,u)=g(x)in Ωu=0in Ωc, |
where h(x,⋅) is nondecreasing and grows sublinearly in the second variable. The solution ug∈Hs0(Ω) is unique and is the unique minimizer of the energy functional
Φ(g)=12∬RN×RN|u(x)−u(y)|2|x−y|N+2sdxdy+∫Ω[H(x,u)−gu]dx |
(where H(x,⋅) denotes the primitive of h(x,⋅)), so the authors investigate the minimization of Φ(g) over G. Besides, in [1], the following nonlocal eigenvalue problem is considered:
{(−Δ)su=λg(x)uin Ωu=0in Ωc. |
The authors prove, among other results, the existence of a minimizer g∈G for the principal eigenvalue λ(g). Optimization of the principal eigenvalue of fractional operators has significant applications in biomathematics, see [25]. In all the aforementioned problems, optimization in G also yields representation formulas and qualitative properties (e.g., Steiner symmetry over convenient domains) of the optimal data.
In the present paper, we focus on the following nonlinear, nonlocal operator:
LKu(x)=limε→0+∫Bcε(x)|u(x)−u(y)|p−2(u(x)−u(y))K(x,y)dy. |
Here N⩾2, p>1, s∈(0,1), and K:RN×RN→R is a measurable kernel s.t. for a.e. x,y∈RN,
(K1) K(x,y)=K(y,x);
(K2) C1⩽K(x,y)|x−y|N+ps⩽C2 (0<C1⩽C2).
If C1=C2=CN,p,s>0 (a normalization constant varying from one reference to the other), LK reduces to the s-fractional p-Laplacian
(−Δ)spu(x)=CN,p,slimε→0+∫Bcε(x)|u(x)−u(y)|p−2(u(x)−u(y))|x−y|N+psdy, |
which in turn coincides with the s-fractional Laplacian seen above for p=2. The nonlinear operator LK arises from problems in game theory (see [2,6]). Besides, the special case (−Δ)sp can be seen as either an approximation of the classical p-Laplace operator for fixed p and s→1− (see [19]), or an approximation of the fractional ∞-Laplacian for fixed s and p→∞, with applications to the problem of Hölder continuous extensions of functions (see [22]). Equations driven by the fractional p-Laplacian are the subject of a vast literature, dealing with existence, qualitative properties, and regularity of the solutions (see for instance [13,14,15,24]).
Inspired by the cited references, we will examine two variational problems driven by LK, set on a bounded domain Ω with C1,1-smooth boundary, with a datum g varying in a rearrangement class G, and optimize the corresponding functionals. First, we consider the following nonlinear, nonlocal eigenvalue problem:
{LKu=λg(x)|u|p−2uin Ωu=0in Ωc. | (1.1) |
Let λ(g) be the principal eigenvalue of (1.1), defined by
λ(g)=infu≠0∬RN×RN|u(x)−u(y)|pK(x,y)dxdy∫Ωg|u|pdx, |
and ug be the (unique) associated eigenfunction s.t. ug>0 in Ω and
∫Ωgupgdx=1. |
With such definitions, we will study the following optimization problem:
ming∈Gλ(g). |
Precisely, we will prove that such problem admits at least one solution, that any solution actually minimizes λ(g) over the larger set ¯G, while all minimizers over ¯G lie in G, and finally that any minimal weight can be represented as a nondecreasing function of the corresponding eigenfunction:
Theorem 1.1. Let Ω⊂RN be a bounded domain with C1,1-boundary, p>1, s∈(0,1), K:RN×RN→R be measurable satisfying (K1), (K2), g0∈L∞(Ω)+∖{0}, G be the rearrangement class of g0. For all g∈¯G, let λ(g) be the principal eigenvalue of (1.1). Then,
(i) there exists ˆg∈G s.t. λ(ˆg)⩽λ(g) for all g∈G;
(ii) for all ˆg as in (i) and g∈¯G∖G, λ(ˆg)<λ(g);
(iii) for all ˆg as in (i) there exists a nondecreasing map η:R→R s.t. ˆg=η∘uˆg in Ω.
Then, we will focus on the following general nonlinear Dirichlet problem:
{LKu+h(x,u)=g(x)in Ωu=0in Ωc, | (1.2) |
where, in addition to the previous hypotheses, we assume that h:Ω×R→R+ is a Carathéodory mapping satisfying the following conditions:
(h1) h(x,⋅) is nondecreasing in R for a.e. x∈Ω;
(h2) h(x,t)⩽C0(1+|t|q−1) for a.e. x∈Ω and all t∈R, with C0>0 and q∈(1,p).
For all g∈¯G problem (1.2) has a unique solution ug, with associated energy
Ψ(g)=1p∬RN×RN|ug(x)−ug(y)|pK(x,y)dxdy+∫Ω[H(x,ug)−gug]dx |
(where H(x,⋅) denotes the primitive of h(x,⋅)). Our second result deals with following optimization problem:
ming∈GΨ(g), |
and is stated as follows:
Theorem 1.2. Let Ω⊂RN be a bounded domain with C1,1-boundary, p>1, s∈(0,1), K:RN×RN→R be measurable satisfying (K1), (K2), h:Ω×R→R+ be a Carathéodory mapping satisfying (h1), (h2), g0∈L∞(Ω)+∖{0}, G be the rearrangement class of g0. For all g∈¯G, let ug be the solution of (1.2) and Ψ(g) be the associated energy. Then,
(i) there exists ˆg∈G s.t. Ψ(ˆg)⩽Ψ(g) for all g∈G;
(ii) for all ˆg as in (i) and g∈¯G∖G, Ψ(ˆg)<Ψ(g);
(iii) for all ˆg as in (i) there exists a nondecreasing map η:R→R s.t. ˆg=η∘uˆg in Ω.
Theorem 1.1 above extends [1, Theorem 1.1] to the nonlinear framework, which requires some delicate arguments due to the non-Hilbertian structure of the problem. Similarly, Theorem 1.2 extends [26, Theorem 3.2], also introducing the structure property of minimizers in (iii).
The dual problems, i.e., maximization of λ(g) and Ψ(g) respectively, remain open for now. The reason is easily understood, as soon as we recall that both λ(g) and Ψ(g) admit variational characterizations as minima of convenient functions on the Sobolev space Ws,p0(Ω), so further minimizing with respect to g conjures a 'double minimization' problem. On the contrary, maximizing λ(g), Ψ(g), respectively, would result in a min-max problem, which requires a different approach.
The structure of the paper is the following: In Section 2 we recall some preliminaries on rearrangement classes and fractional order equations; in Section 3 we deal with the eigenvalue problem (1.1); and in Section 4 we deal with the general Dirichlet problem (1.2).
Notation. For all Ω⊂RN, we denote by |Ω| the N-dimensional Lebesgue measure of Ω and Ωc=RN∖Ω. For all x∈RN, r>0 we denote by Br(x) the open ball centered at x with radius r. When we say that g⩾0 in Ω, we mean g(x)⩾0 for a.e. x∈Ω, and similar expressions. Whenever X is a function space on the domain Ω, X+ denotes the positive order cone of X. In any Banach space we denote by → strong (or norm) convergence, by ⇀ weak convergence, and by ∗⇀ weak* convergence. For all q∈[1,∞], we denote by ‖ the norm of L^q(\Omega) . Finally, C denotes several positive constants, varying from line to line.
In this section we collect some necessary preliminary results on rearrangement classes and fractional Sobolev spaces.
Let \Omega\subset {\mathbb R}^N ( N \geqslant 2 ) be a bounded domain, g_0\in L^\infty(\Omega) be s.t. 0 \leqslant g_0 \leqslant M in \Omega ( M > 0 ), and g_0 > 0 on some subset of \Omega with positive measure. We say that a function g\in L^\infty(\Omega) is a rearrangement of g_0 , denoted g\sim g_0 , if for all t \geqslant 0
\big|\{x\in\Omega:\, g(x) > t\}\big| = \big|\{x\in\Omega:\, g_0(x) > t\}\big|. |
Also, we define the rearrangement class
\mathcal{G} = \big\{g\in L^\infty(\Omega):\, g\sim g_0\big\}. |
Clearly, 0 \leqslant g \leqslant M in \Omega for all g\in \mathcal{G} . Recalling that L^\infty(\Omega) is the topological dual of L^1(\Omega) , we can endow such space with the weak* topology, characterized by the following type of convergence:
g_n \overset{*}{\rightharpoonup} g \ \Longleftrightarrow \ \lim\limits_n\, \int_\Omega g_nh\, dx = \int_\Omega gh\, dx \ \text{for all } h\in L^1(\Omega) . |
We denote by \overline{\mathcal{G}} the closure of \mathcal{G} in L^\infty(\Omega) with respect to such topology. It is proved in [3,4] that \overline{\mathcal{G}} is a sequentially weakly* compact convex set, and that 0 \leqslant g \leqslant M in \Omega for all g\in \overline{\mathcal{G}} . Therefore, given a sequentially weakly* continuous functional \Phi: \overline{\mathcal{G}}\to {\mathbb R} , there exist \check g, \hat g\in \overline{\mathcal{G}} s.t. for all g\in \overline{\mathcal{G}}
\Phi(\check g) \leqslant \Phi(g) \leqslant \Phi(\hat g). |
In general, the extrema are not attained at points of \mathcal{G} . As usual, we say that \Phi is Gâteaux differentiable at g\in \overline{\mathcal{G}} , if there exists a linear functional \Phi'(g)\in L^\infty(\Omega)^* s.t. for all h\in \overline{\mathcal{G}}
\lim\limits_{\tau\to 0^+}\, \frac{\Phi(g+\tau(h-g))-\Phi(g)}{\tau} = \langle\Phi'(g), h-g\rangle. |
We remark that g\in \overline{\mathcal{G}} being a minimizer (or maximizer) of \Phi does not imply \Phi'(g) = 0 in general. Nevertheless, if \Phi is convex, then for all h\in \overline{\mathcal{G}}
\Phi(h) \geqslant \Phi(g)+\langle\Phi'(g), h-g\rangle, |
with strict inequality if \Phi is strictly convex and h\neq g (see [27] for an introduction to convex functionals and variational inequalities). Finally, let us recall a technical lemma on optimization of linear functionals over \overline{\mathcal{G}} , which also provides a representation formula:
Lemma 2.1. Let h\in L^1(\Omega) . Then,
(i) there exists \hat g\in \mathcal{G} s.t. for all g\in \overline{\mathcal{G}}
\int_\Omega \hat g h\, dx \geqslant \int_\Omega gh\, dx; |
(ii) if \hat g is unique, then there exists a nondecreasing map \eta: {\mathbb R}\to {\mathbb R} s.t. \hat g = \eta\circ h in \Omega .
Proof. By [3, Theorems 1, 4], there exists \hat g\in \mathcal{G} which maximizes the linear functional
g\mapsto\int_\Omega gh\, dx |
over \mathcal{G} . Given g\in \overline{\mathcal{G}}\setminus \mathcal{G} , we can find a sequence (g_n) in \mathcal{G} s.t. g_n \overset{*}{\rightharpoonup} g . For all n\in {\mathbb N} we have
\int_\Omega\hat gh\, dx \geqslant \int_\Omega g_nh\, dx, |
so passing to the limit we get
\int_\Omega\hat gh\, dx \geqslant \int_\Omega gh\, dx, |
thus proving (i). From [3, Theorem 5] we have (ii).
We recall some basic notions about the variational formulations of problems (1.1) and (1.2). For p > 1 , s\in(0, 1) , all open \Omega\subseteq {\mathbb R}^N , and all measurable u:\Omega\to {\mathbb R} we define the Gagliardo seminorm
[u]_{s, p, \Omega} = \Big[\iint_{\Omega\times\Omega}\frac{|u(x)-u(y)|^p}{|x-y|^{N+ps}}\, dx\, dy\Big]^\frac{1}{p}. |
The corresponding fractional Sobolev space is defined by
W^{s, p}(\Omega) = \big\{u\in L^p(\Omega):\, [u]_{s, p, \Omega} < \infty\big\}. |
If \Omega is bounded and with a C^{1, 1} -smooth boundary, we incorporate the Dirichlet conditions by defining the space
W^{s, p}_0(\Omega) = \big\{u\in W^{s, p}( {\mathbb R}^N):\, u = 0 \ \text{in } \Omega^c \big\}, |
endowed with the norm \|u\|_{ W^{s, p}_0(\Omega)} = [u]_{s, p, {\mathbb R}^N} . This is a uniformly convex, separable Banach space with dual W^{-s, p'}(\Omega) , s.t. C^\infty_c(\Omega) is a dense subset of W^{s, p}_0(\Omega) , and the embedding W^{s, p}_0(\Omega)\hookrightarrow L^q(\Omega) is compact for all q\in[1, p^*_s) , where
p^*_s = \begin{cases} \frac{Np}{N-ps} & \text{if } ps < N \\ \infty & \text{if } ps \geqslant N . \end{cases} |
For a detailed account on fractional Sobolev spaces, we refer the reader to [10,21]. Now let K: {\mathbb R}^n\times {\mathbb R}^N\to {\mathbb R} be a measurable kernel satisfying (K_1) and (K_2) . We introduce an equivalent norm on W^{s, p}_0(\Omega) by setting
[u]_K = \Big[\iint_{ {\mathbb R}^N\times {\mathbb R}^N}|u(x)-u(y)|^pK(x, y)\, dx\, dy\Big]^\frac{1}{p}. |
We can now rephrase more carefully the definitions given in Section 1, by defining the operator \mathcal{L}_K: W^{s, p}_0(\Omega)\to W^{-s, p'}(\Omega) as the gradient of the C^1 -functional
u \mapsto \frac{[u]_K^p}{p}. |
Equivalently, for all u, \varphi\in W^{s, p}_0(\Omega) we set
\langle \mathcal{L}_K u, \varphi\rangle = \iint_{ {\mathbb R}^N\times {\mathbb R}^N}|u(x)-u(y)|^{p-2}(u(x)-u(y))(\varphi(x)-\varphi(y))K(x, y)\, dx\, dy. |
Both problems that we are going to study belong to the following class of nonlinear, nonlocal Dirichlet problems:
\begin{equation} \begin{cases} \mathcal{L}_K u = f(x, u) & \text{in } \Omega \\ u = 0 & \text{in } \Omega^c , \end{cases} \end{equation} | (2.1) |
where f:\Omega\times {\mathbb R}\to {\mathbb R} is a Carathéodory mapping subject to the following subcritical growth conditions: there exist C > 0 , r\in(1, p^*_s) s.t. for a.e. x\in\Omega and all t\in {\mathbb R}
\begin{equation} |f(x, t)| \leqslant C(1+|t|^{r-1}). \end{equation} | (2.2) |
We say that u\in W^{s, p}_0(\Omega) is a weak solution of (2.1), if for all \varphi\in W^{s, p}_0(\Omega)
\langle \mathcal{L}_K u, \varphi\rangle = \int_\Omega f(x, u)\varphi\, dx. |
There is a wide literature on problem (2.1), especially for the model case \mathcal{L}_K = (-\Delta)_p^s , see for instance [9,14,16,17,24]. We will only need to recall the following properties, which can be proved adapting [16, Proposition 2.3] and [9, Theorem 1.5], respectively:
Lemma 2.2. Let f satisfy (2.2), u\in W^{s, p}_0(\Omega) be a weak solution of (2.1). Then, u\in L^\infty(\Omega) .
Lemma 2.3. Let f satisfy (2.2), and \delta > 0 , c\in C(\overline\Omega)_+ be s.t. for a.e. x\in\Omega and all t\in[0, \delta]
f(x, t) \geqslant -c(x)t^{p-1}. |
Also, let u\in W^{s, p}_0(\Omega)_+ be a weak solution of (2.2). Then, either u = 0 , or u > 0 in \Omega .
We will not cope with regularity of the weak solutions here. In the model case of the fractional p -Laplacian, under hypothesis (2.2), using Lemma 2.2 above and [15, Theorems 1.1, 2.7], it can be seen that whenever u\in W^{s, p}_0(\Omega) solves (2.1), we have u\in C^s({\mathbb R}^N) and there exist \alpha\in(0, s) depending only on the data of the problem, s.t. the function
\frac{u}{{\rm dist}(\cdot, \Omega^c)^s} |
admits a \alpha -Hölder continuous extension to \overline\Omega . The same result is not known for the general operator \mathcal{L}_K , except the linear case p = 2 with a special anisotropic kernel, see [28].
For future use, we prove here a technical lemma:
Lemma 2.4. Let (g_n) be a sequence in \overline{\mathcal{G}} s.t. g_n \overset{*}{\rightharpoonup} g , (u_n) be a bounded sequence in W^{s, p}_0(\Omega) , r\in[1, p^*_s) . Then, there exists u\in W^{s, p}_0(\Omega) s.t. up to a subsequence
\lim\limits_n\, \int_\Omega g_n|u_n|^r\, dx = \int_\Omega g|u|^r. |
Proof. By the compact embedding W^{s, p}_0(\Omega)\hookrightarrow L^r(\Omega) , passing if necessary to a subsequence we have u_n\to u in L^r(\Omega) and u_n(x)\to u(x) for a.e. x\in\Omega , as n\to\infty . In particular |u|^r\in L^1(\Omega) , so
\lim\limits_n\int_\Omega (g_n-g)|u|^r\, dx = 0. |
Besides, recalling that 0 \leqslant g_n \leqslant M in \Omega for all n\in {\mathbb N} , we have by Hölder's inequality
\begin{align*} \int_\Omega\big[g_n|u_n|^r-g|u|^r\big]\, dx & \leqslant \int_\Omega g_n||u_n|^r-|u|^r|\, dx+\int_\Omega(g_n-g)|u|^r\, dx \\ & \leqslant C\int_\Omega\big[|u_n|^{r-1}+|u|^{r-1}\big]|u_n-u|\, dx+\int_\Omega(g_n-g)|u|^r\, dx \\ & \leqslant C\big[\|u_n\|_r^{r-1}+\|u\|_r^{r-1}\big]\|u_n-u\|_r+\int_\Omega(g_n-g)|u|^r\, dx, \end{align*} |
and the latter tends to 0 as n\to\infty .
In this section we consider the eigenvalue problem (1.1) and prove Theorem 1.1. Let \Omega , p , s , K , g_0 be as in Section 1. For any g\in \overline{\mathcal{G}} , as in Subsection 2.2 we say that u\in W^{s, p}_0(\Omega) is a (weak) solution of (1.1) if for all \varphi\in W^{s, p}_0(\Omega)
\langle \mathcal{L}_K u, \varphi\rangle = \lambda\int_\Omega g|u|^{p-2}u\varphi\, dx. |
We say that \lambda\in {\mathbb R} is an eigenvalue if (1.1) admits a solution u\neq 0 , which is then called a \lambda -eigenfunction. Though a full description of the eigenvalues of (1.1) is missing, from [11,12,22] we know that for all g\in L^\infty(\Omega)_+ there exists a principal eigenvalue \lambda(g) > 0 , namely the smallest positive eigenvalue, which admits the following variational characterization:
\begin{equation} \lambda(g) = \inf\limits_{u\neq 0}\, \frac{[u]_K^p}{\int_\Omega g|u|^p\, dx}. \end{equation} | (3.1) |
In addition, from [11] we know that \lambda(g) is an isolated eigenvalue, simple, with constant sign eigenfunctions, while for any eigenvalue \lambda > \lambda(g) the associated \lambda -eigenfunctions change sign in \Omega . So, recalling Lemma 2.3, there exists a unique normalized positive \lambda(g) -eigenfunction u_g\in W^{s, p}_0(\Omega) s.t.
\int_\Omega gu_g^p\, dx = 1, \ [u_g]_K^p = \lambda(g). |
In particular g\mapsto\lambda(g) defines a real-valued functional defined in the rearrangement class of weights \mathcal{G} (or in \overline{\mathcal{G}} ), and we are interested in the minimizers of such functional. Equivalently, we may set for all g\in \overline{\mathcal{G}}
\Phi(g) = \frac{1}{\lambda(g)^2} = \sup\limits_{u\neq 0}\, \frac{\Big[\int_\Omega g|u|^p\, dx\Big]^2}{[u]_K^{2p}}, |
and consider the maximization problem
\max\limits_{g\in \mathcal{G}}\, \Phi(g). |
First, we want to maximize \Phi(g) over \overline{\mathcal{G}} , which is possible due to the following lemma:
Lemma 3.1. The functional \Phi(g) is sequentially weakly* continuous in \overline{\mathcal{G}} .
Proof. Let (g_n) be a sequence in \overline{\mathcal{G}} s.t. g_n \overset{*}{\rightharpoonup} g , and for simplicity denote u_n = u_{g_n} for all n\in {\mathbb N} , and u = u_g . We need to prove that \Phi(g_n)\to\Phi(g) . Since u^p\in L^1(\Omega) , we have
\lim\limits_n\, \int_\Omega g_nu^p\, dx = \int_\Omega gu^p\, dx = 1. |
Also, by definition of \Phi we have for all n\in {\mathbb N}
\Phi(g_n) \geqslant \frac{\Big[\int_\Omega g_nu^p\, dx\Big]^2}{[u]_K^{2p}}, |
and the latter tends to \Phi(g) as n\to\infty . Therefore
\begin{equation} \liminf\limits_n\, \Phi(g_n) \geqslant \Phi(g). \end{equation} | (3.2) |
In particular, for all n\in {\mathbb N} we have
[u_n]_K = \Phi(g_n)^{-\frac{1}{2p}} \leqslant C, |
so (u_n) is bounded in W^{s, p}_0(\Omega) . By reflexivity and the compact embedding W^{s, p}_0(\Omega)\hookrightarrow L^p(\Omega) , passing to a subsequence we have u_n\rightharpoonup v in W^{s, p}_0(\Omega) , u_n\to v in L^p(\Omega) , and u_n(x)\to v(x) for a.e. x\in\Omega , as n\to\infty . In particular, v \geqslant 0 in \Omega . By convexity we have
\liminf\limits_n\, [u_n]_K^{2p} \geqslant [v]_K^{2p}. |
By Lemma 2.4 (with r = p ) we also have
\lim\limits_n\, \int_\Omega g_nu_n^p\, dx = \int_\Omega gv^p\, dx. |
So we get
\begin{align*} \limsup\limits_n\, \Phi(g_n) = \limsup\limits_n\, \frac{\Big[\int_\Omega g_nu_n^p\, dx\Big]^2}{[u_n]_K^{2p}} & \leqslant \frac{\Big[\int_\Omega gv^p\, dx\Big]^2}{[v]_K^{2p}} \leqslant \Phi(g). \end{align*} |
This, besides (3.2), concludes the proof.
Lemma 3.1, along with the compactness of \overline{\mathcal{G}} , proves that \Phi(g) admits a minimizer and a maximizer in \overline{\mathcal{G}} . We next need to ensure that at least one maximizer lies in the smaller set \mathcal{G} . In the next lemmas we will investigate further properties of \Phi .
Lemma 3.2. The functional \Phi is strictly convex in \overline{\mathcal{G}} .
Proof. We introduce an alternative expression for \Phi . For all g\in \overline{\mathcal{G}} , u\in W^{s, p}_0(\Omega)_+ set
F(g, u) = 2\int_\Omega gu^p\, dx-[u]_K^{2p}. |
We fix g\in \overline{\mathcal{G}} and maximize F(g, \cdot) over positive functions. For all u\in W^{s, p}_0(\Omega)_+\setminus\{0\} and \tau > 0 , the function
F(g, \tau u) = 2\tau^p\int_\Omega gu^p\, dx-\tau^{2p}[u]_K^p |
is differentiable in \tau with derivative
\frac{\partial}{\partial\tau}F(g, \tau u) = 2p\tau^{p-1}\int_\Omega gu^p\, dx-2p\tau^{2p-1}[u]_K^{2p}. |
So the maximum of \tau\mapsto F(g, \tau u) is attained at
\tau_0(u) = \frac{\Big[\int_\Omega gu^p\, dx\Big]^\frac{1}{p}}{[u]_K^2} > 0, |
and amounts at
F(g, \tau_0(u)u) = \frac{\Big[\int_\Omega gu^p\, dx\Big]^2}{[u]_K^{2p}}. |
Maximizing further over u , we obtain
\sup\limits_{u > 0}\, F(g, u) = \sup\limits_{u\in W^{s, p}_0(\Omega)_+\setminus\{0\}}\, \frac{\Big[\int_\Omega gu^p\, dx\Big]^2}{[u]_K^{2p}}. |
Noting that [|u|]_K \leqslant[u]_K for all u\in W^{s, p}_0(\Omega) , and recalling (3.1), we have for all g\in \overline{\mathcal{G}}
\begin{equation} \Phi(g) = \sup\limits_{u\in W^{s, p}_0(\Omega)_+\setminus\{0\}}\, F(g, u) = \frac{1}{\lambda(g)^2}. \end{equation} | (3.3) |
We claim that the supremum in (3.3) is attained at the unique function
\begin{equation} \tilde u_g = \frac{u_g}{\lambda(g)^\frac{2}{p}} = \tau_0(u_g)u_g. \end{equation} | (3.4) |
Indeed, by normalization of u_g we have
F(g, \tilde u_g) = \frac{2}{\lambda(g)^2}\, \int_\Omega gu_g^p\, dx-\frac{[u_g]_K^{2p}}{\lambda(g)^4} = \frac{1}{\lambda(g)^2}. |
For uniqueness, first consider a function u = \tau u_g with \tau\neq\tau_0(u_g) . By unique maximization in \tau we have
F(g, \tau u_g) < F(g, \tau_0(u_g) u_g) = F(g, \tilde u_g) = \frac{1}{\lambda(g)^2}. |
Besides, for all v\in W^{s, p}_0(\Omega)_+\setminus\{0\} which is not a \lambda(g) -eigenfunction, arguing as above with v replacing u , and recalling that the infimum in (3.1) is attained only at principal eigenfunctions, we have
F(g, v) \leqslant F(g, \tau_0(v) v) = \frac{\Big[\int_\Omega gv^p\, dx\Big]^2}{[v]_K^{2p}} < \frac{1}{\lambda(g)^2}. |
So, \tilde u_g is the unique maximizer of (3.3).
We now prove that \Phi is convex. Let g_1, g_2\in \overline{\mathcal{G}} , \tau\in(0, 1) and set
g_\tau = (1-\tau)g_1+\tau g_2, |
so g_\tau\in \overline{\mathcal{G}} (a convex set, as seen in Subsection 2.1). For all u\in W^{s, p}_0(\Omega)_+\setminus\{0\} , we have by (3.3)
\begin{align*} F(g_\tau, u) & = 2(1-\tau)\int_\Omega g_1u^p\, dx+2\tau\int_\Omega g_2u^p\, dx-[u]_K^{2p} \\ & = (1-\tau)F(g_1, u)+\tau F(g_2, u) \leqslant (1-\tau)\Phi(g_1)+\tau\Phi(g_2). \end{align*} |
Taking the supremum over u and using (3.3) again,
\Phi(g_\tau) \leqslant (1-\tau)\Phi(g_1)+\tau\Phi(g_2). |
To prove that \Phi is strictly convex, we argue by contradiction, assuming that for some g_1\neq g_2 as above and \tau\in(0, 1)
\Phi(g_\tau) = (1-\tau)\Phi(g_1)+\tau\Phi(g_2). |
Set \tilde u_i = \tilde u_{g_i} ( i = 1, 2 ) and \tilde u_\tau = \tilde u_{g_\tau} for brevity. Then, by (3.3) and the equality above
(1-\tau)F(g_1, \tilde u_\tau)+\tau F(g_2, \tilde u_\tau) = (1-\tau)F(g_1, \tilde u_1)+\tau F(g_2, \tilde u_2). |
Recalling that \tilde u_i is the only maximizer of F(g_i, \cdot) , the last inequality implies \tilde u_1 = \tilde u_2 = \tilde u_\tau , as well as
\Phi(g_1) = F(g_1, \tilde u_\tau) = F(g_2, \tilde u_\tau) = \Phi(g_2). |
Therefore we have \lambda(g_1) = \lambda(g_2) = \lambda . Moreover, \tilde u_\tau > 0 is a \lambda -eigenfunction with both weights g_1 , g_2 , i.e., for all \varphi\in W^{s, p}_0(\Omega)
\lambda\int_\Omega g_1\tilde u_\tau^{p-1}\varphi\, dx = \langle \mathcal{L}_K\tilde u_\tau, \varphi\rangle = \lambda\int_\Omega g_2\tilde u_\tau^{p-1}\varphi\, dx. |
So g_1\tilde u_\tau^{p-1} = g_2\tilde u_\tau^{p-1} in \Omega , which in turn, since \tilde u_\tau > 0 , implies g_1 = g_2 a.e. in \Omega , a contradiction.
The next lemma establishes differentiability of \Phi .
Lemma 3.3. The functional \Phi is Gâteaux differentiable in \overline{\mathcal{G}} , and for all g, h\in \overline{\mathcal{G}}
\langle\Phi'(g), h-g\rangle = 2\int_\Omega(h-g)\tilde u_g^p\, dx, |
where \tilde u_g is the principal eigenfunction normalized as in (3.4).
Proof. First, let (g_n) be a sequence in \overline{\mathcal{G}} s.t. g_n \overset{*}{\rightharpoonup} g , and set for brevity \tilde u_n = \tilde u_{g_n} , \tilde u = \tilde u_g . We claim that
\begin{equation} \lim\limits_n\, \int_\Omega|\tilde u_n-\tilde u|^p\, dx = 0. \end{equation} | (3.5) |
Indeed, by normalization we have for all n\in {\mathbb N}
[\tilde u_n]_K^{2p} = \Phi(g_n), |
and the latter is bounded from above, since \Phi(g) has a maximizer in \overline{\mathcal{G}} . So, (\tilde u_n) is bounded in W^{s, p}_0(\Omega) . By uniform convexity and the compact embedding W^{s, p}_0(\Omega)\hookrightarrow L^p(\Omega) , passing to a subsequence we have \tilde u_n\rightharpoonup v in W^{s, p}_0(\Omega) , \tilde u_n\to v in L^p(\Omega) , and \tilde u_n(x)\to v(x) for a.e. x\in\Omega , as n\to\infty (in particular v \geqslant 0 in \Omega ). By convexity, we see that
\liminf\limits_n\, [\tilde u_n]_K^{2p} \geqslant [v]_K^{2p}. |
By Lemma 2.4, we have
\lim\limits_n\, \int_\Omega g_n\tilde u_n^p\, dx = \int_\Omega gv^p\, dx. |
By Lemma 3.1, we have \Phi(g_n)\to\Phi(g) , so by (3.3) we get
\begin{align*} \Phi(g) & = \lim\limits_n\, F(g_n, \tilde u_n) \\ & \leqslant 2\lim\limits_n\, \int_\Omega g_n\tilde u_n^p\, dx-\liminf\limits_n\, [\tilde u_n]_K^{2p} \\ & \leqslant 2\int_\Omega gv^p\, dx-[v]_K^{2p} \\ & = F(g, v) \leqslant \Phi(g). \end{align*} |
Therefore v is a maximizer of F(g, \cdot) over W^{s, p}_0(\Omega)_+ , hence by uniqueness v = \tilde u . Then we have \tilde u_n\to\tilde u in L^p(\Omega) , which is equivalent to (3.5).
We claim that for all n\in {\mathbb N}
\begin{equation} \Phi(g)+2\int_\Omega(g_n-g)\tilde u^p\, dx \leqslant \Phi(g_n) \leqslant \Phi(g)+2\int_\Omega(g_n-g)\tilde u_n^p\, dx. \end{equation} | (3.6) |
Indeed, by (3.3) we have
\begin{align*} \Phi(g)+2\int_\Omega(g_n-g)\tilde u^p\, dx & \leqslant \Phi(g_n) \\ & = F(g, \tilde u_n)+2\int_\Omega(g_n-g)\tilde u_n^p\, dx \\ & \leqslant \Phi(g)+2\int_\Omega(g_n-g)\tilde u_n^p\, dx. \end{align*} |
Now fix g, h\in \overline{\mathcal{G}} , g\neq h , and a sequence (\tau_n) in (0, 1) s.t. \tau_n\to 0 . By convexity of \overline{\mathcal{G}} , we have for all n\in {\mathbb N}
g_n = g+\tau_n(h-g) \in \overline{\mathcal{G}}. |
Also, clearly g_n \overset{*}{\rightharpoonup} g . By (3.6), setting as usual \tilde u_n = \tilde u_{g_n} and \tilde u = \tilde u_g , we have for all n\in {\mathbb N}
2\tau_n\int_\Omega(h-g)\tilde u^p\, dx \leqslant \Phi(g_n)-\Phi(g) \leqslant 2\tau_n\int_\Omega(h-g)\tilde u_n^p\, dx. |
Dividing by \tau_n > 0 and recalling (3.5), we get
\lim\limits_n\, \frac{\Phi(g+\tau_n(h-g))-\Phi(g)}{\tau_n} = 2\int_\Omega(h-g)\tilde u^p\, dx. |
Note that 2\tilde u^p\in L^1(\Omega)\subset L^\infty(\Omega)^* , and by the arbitrariness of the sequence (\tau_n) we deduce that \Phi is Gâteaux differentiable at g with
\langle\Phi'(g), h-g\rangle = 2\int_\Omega(h-g)\tilde u^p\, dx, |
which concludes the proof.
We can now prove the main result of this section.
Proof of Theorem 1.1. We already know that \Phi has a maximizer \bar g over \overline{\mathcal{G}} . Set \bar w = 2\tilde u_{\bar g}^p\in L^1(\Omega) , then by Lemma 3.3 we have \Phi'(\bar g) = \bar w . Now we maximize on \overline{\mathcal{G}} the linear functional
g \mapsto \int_\Omega g\bar w\, dx. |
By Lemma 2.1 (i), there exists \hat g\in \mathcal{G} s.t. for all g\in \overline{\mathcal{G}}
\int_\Omega\hat g\bar w\, dx \geqslant \int_\Omega g\bar w\, dx. |
In particular we have
\begin{equation} \int_\Omega\hat g\bar w\, dx \geqslant \int_\Omega \bar g\bar w\, dx. \end{equation} | (3.7) |
By Lemma 3.2, the functional \Phi is convex. Therefore, using also Lemma 3.3 and (3.7), we have
\Phi(\hat g) \geqslant \Phi(\bar g)+\int_\Omega(\hat g-\bar g)\bar w\, dx \geqslant \Phi(\bar g). |
Thus, \hat g\in \mathcal{G} is as well a maximizer of \Phi over \overline{\mathcal{G}} , which proves (i) since maximizers of \Phi and minimizers of \lambda(g) coincide. In addition, by the relation above we have
\int_\Omega(\hat g-\bar g)\bar w\, dx = 0. |
We will now prove that \hat g = \bar g , arguing by contradiction. Assume \hat g\neq\bar g , then by the strict convexity of \Phi (Lemma 3.2 again) we have
\Phi(\hat g) > \Phi(\bar g)+\int_\Omega(\hat g-\bar g)\bar w\, dx = \Phi(\bar g), |
against the maximality of \bar g . So, any maximizer of \Phi over \overline{\mathcal{G}} actually lies in \mathcal{G} , which proves (ii). Finally, let \hat g\in \mathcal{G} be any maximizer of \Phi and set \hat w = 2\tilde u_{\hat g}^p\in L^1(\Omega) . By Lemmas 3.2 and 3.3, for all g\in \overline{\mathcal{G}}\setminus\{\hat g\} we have
\Phi(\hat g) \geqslant \Phi(g) > \Phi(\hat g)+\int_\Omega(g-\hat g)\hat w\, dx, |
hence
\int_\Omega\hat g\hat w\, dx > \int_\Omega g\hat w\, dx. |
Equivalently, \hat g is the only maximizer over \overline{\mathcal{G}} of the linear functional above, induced by the function \hat w . By Lemma 2.1 (ii), there exists a nondecreasing map \tilde\eta: {\mathbb R}\to {\mathbb R} s.t. in \Omega
\hat g = \tilde\eta\circ\hat w. |
Now we recall (3.4) and the definition of \hat w , and by setting for all t \geqslant 0
\eta(t) = \tilde\eta\Big(\frac{2t^p}{\lambda(\hat g)^2}\Big), |
while \eta(t) = \eta(0) for all t < 0 , we immediately see that \eta: {\mathbb R}\to {\mathbb R} is a nondecreasing map s.t. \hat g = \eta\circ u_{\hat g} in \Omega , thus proving (iii).
In this section we consider problem (1.2) and prove Theorem 1.2. Let \Omega , p , s , K , g_0 be as in Section 1, and h satisfy (h_1) , (h_2) . For any g\in \overline{\mathcal{G}} , we say that u\in W^{s, p}_0(\Omega) is a weak solution of (1.2) if for all \varphi\in W^{s, p}_0(\Omega)
\langle \mathcal{L}_K(u), \varphi\rangle+\int_\Omega h(x, u)\varphi\, dx = \int_\Omega g\varphi\, dx. |
By classical results (see for instance [18] for the fractional p -Laplacian), for all g\in \overline{\mathcal{G}} problem (1.2) has a unique solution u_g\in W^{s, p}_0(\Omega) . In addition, by Lemma 2.2 we have u_g\in L^\infty(\Omega) . Such solution is the unique minimizer in W^{s, p}_0(\Omega) of the energy functional associated to (1.2). The corresponding energy, depending on g\in \overline{\mathcal{G}} , is given by
\Psi(g) = \frac{[u_g]_K^p}{p}+\int_\Omega\big[H(x, u_g)-gu_g\big]\, dx, |
where for all (x, t)\in\Omega\times {\mathbb R} we have set
H(x, t) = \int_0^t h(x, \tau)\, d\tau. |
We are interested in the minimizers of \Psi(g) over \mathcal{G} . Equivalently, we may set for all g\in \overline{\mathcal{G}} , u\in W^{s, p}_0(\Omega)
E(g, u) = \int_\Omega\big[gu-H(x, u)\big]\, dx-\frac{[u]_K^p}{p}, |
and maximize E(g, \cdot) with respect to u , thus defining
\begin{equation} \Phi(g) = \sup\limits_{u\in W^{s, p}_0(\Omega)}\, E(g, u) = E(g, u_g). \end{equation} | (4.1) |
So, as in Section 3, we are led to the maximization problem
\max\limits_{g\in \mathcal{G}}\, \Phi(g). |
First, we prove the continuity of \Phi .
Lemma 4.1. The functional \Phi is sequentially weakly* continuous in \overline{\mathcal{G}} .
Proof. Let (g_n) be a sequence in \overline{\mathcal{G}} s.t. g_n \overset{*}{\rightharpoonup} g , and denote u_n = u_{g_n} , u = u_g . By (4.1), for all n\in {\mathbb N} we have
\begin{align*} \Phi(g_n) & = E(g_n, u_n) \\ & \geqslant E(g_n, u) \\ & = E(g, u)+\int_\Omega(g_n-g)u\, dx \\ & = \Phi(g)+\int_\Omega(g_n-g)u\, dx. \end{align*} |
Passing to the limit as n\to\infty and using weak* convergence, we get
\begin{equation} \liminf\limits_n\Phi(g_n) \geqslant \Phi(g)+\lim\limits_n\int_\Omega(g_n-g)u\, dx = \Phi(g). \end{equation} | (4.2) |
From (1.2) with datum g_n and solution u_n , multiplying by u_n again, we get for all n\in {\mathbb N}
\begin{equation} \int_\Omega\big[g_n-h(x, u_n)\big]u_n\, dx = [u_n]_K^p. \end{equation} | (4.3) |
Since \|g_n\|_\infty \leqslant M and by the continuous embedding W^{s, p}_0(\Omega)\hookrightarrow L^1(\Omega) , we have
\Big|\int_\Omega g_nu_n\, dx\Big| \leqslant C[u_n]_K, |
with C > 0 independent of n . Also, by (h_2) and the continuous embedding W^{s, p}_0(\Omega)\hookrightarrow L^q(\Omega) , we have
\Big|\int_\Omega h(x, u_n)u_n\, dx\Big| \leqslant C\int_\Omega\big[|u_n|+|u_n|^q\big]\, dx \leqslant C[u_n]_K+C[u_n]_K^q. |
So (4.3) implies for all n\in {\mathbb N}
[u_n]_K^p \leqslant C[u_n]_K+C[u_n]_K^q. |
Recalling that q < p , we deduce that (u_n) is bounded in W^{s, p}_0(\Omega) . Passing to a subsequence, we have u_n\rightharpoonup v in W^{s, p}_0(\Omega) , u_n\to v in L^p(\Omega) , and u_n(x)\to v(x) for a.e. x\in\Omega , as n\to\infty . By convexity we have
\liminf\limits_n\, [u_n]_K^p \geqslant [v]_K^p. |
By Lemma 2.4 (with r = 1 ) we find
\begin{equation} \lim\limits_n\, \int_\Omega g_nu_n\, dx = \int_\Omega gv\, dx. \end{equation} | (4.4) |
Finally, we have
\begin{equation} \lim\limits_n\, \int_\Omega H(x, u_n)\, dx = \int_\Omega H(x, v)\, dx. \end{equation} | (4.5) |
Indeed, applying (h_2) , Lagrange's rule, and Hölder's inequality, we get for all n\in {\mathbb N}
\begin{align*} \int_\Omega\big|H(x, u_n)-H(x, v)\big|\, dx & \leqslant C\int_\Omega\big[1+|u_n|^{q-1}+|v|^{q-1}\big]|u_n-v|\, dx \\ & \leqslant C\|u_n-v\|_1+C\big[\|u_n\|_q^{q-1}+\|v\|_q^{q-1}\big]\|u_n-v\|_q, \end{align*} |
and the latter tends to 0 as n\to\infty , by the continuous embeddings of L^p(\Omega) into L^1(\Omega) , L^q(\Omega) , respectively, thus proving (4.5).
Next, we start from (4.1) and we apply (4.4) and (4.5):
\begin{align*} \limsup\limits_n\Phi(g_n) & = \limsup\limits_n E(g_n, u_n) \\ & \leqslant \lim\limits_n\, \int_\Omega[g_nu_n-H(x, u_n)\big]\, dx-\liminf\limits_n\frac{[u_n]_K^p}{p} \\ & \leqslant \int_\Omega\big[gv-H(x, v)\big]\, dx-\frac{[v]_K^p}{p} = E(g, v), \end{align*} |
and the latter does not exceed \Phi(g) , so
\begin{equation} \limsup\limits_n\Phi(g_n) \leqslant \Phi(g). \end{equation} | (4.6) |
Comparing (4.2) and (4.6), we have \Phi(g_n)\to\Phi(g) , which concludes the proof.
By Lemma 4.1, \Phi has both a minimizer and a maximizer over \overline{\mathcal{G}} . Next we prove strict convexity:
Lemma 4.2. The functional \Phi is strictly convex in \overline{\mathcal{G}} .
Proof. The convexity of \Phi follows as in Lemma 3.2, since \Phi(g) is the supremum of linear functionals (in g ). To prove strict convexity, we argue by contradiction. Let g_1, g_2\in \overline{\mathcal{G}} be s.t. g_1\neq g_2 , set for all \tau\in(0, 1)
g_\tau = (1-\tau)g_1+\tau g_2 \in \overline{\mathcal{G}}, |
and assume that for some \tau\in(0, 1)
\Phi(g_\tau) = (1-\tau)\Phi(g_1)+\tau\Phi(g_2). |
As usual, set u_i = u_{g_i} ( i = 1, 2 ) and u_\tau = u_{g_\tau} . By linearity of E(g, u_\tau) in g and (4.1), the relation above rephrases as
(1-\tau)E(g_1, u_\tau)+\tau E(g_2, u_\tau) = (1-\tau)E(g_1, u_1)+\tau E(g_2, u_2). |
Recalling that E(g_i, u_\tau) \leqslant E(g_i, u_i) ( i = 1, 2 ) and the uniqueness of the maximizer in (4.1), we deduce u_1 = u_2 = u_\tau . Now test (1.2) with an arbitrary \varphi\in W^{s, p}_0(\Omega) :
\int_\Omega g_1\varphi\, dx = \langle \mathcal{L}_K u_\tau, \varphi\rangle+\int_\Omega h(x, u_\tau)\varphi\, dx = \int_\Omega g_2\varphi\, dx. |
So we have g_1 = g_2 a.e. in \Omega , a contradiction. Thus, \Phi is strictly convex.
The last property we need is differentiability.
Lemma 4.3. The functional \Phi is Gâteaux differentiable in \overline{\mathcal{G}} , and for all g, k\in \overline{\mathcal{G}}
\langle\Phi'(g), k-g\rangle = \int_\Omega (k-g)u_g\, dx. |
Proof. First, let (g_n) be a sequence in \overline{\mathcal{G}} s.t. g_n \overset{*}{\rightharpoonup} g , and let u_n = u_{g_n} , u = u_g . From Lemma 4.1 we know that \Phi(g_n) tends to \Phi(g) , i.e.,
\begin{equation} \lim\limits_n E(g_n, u_n) = E(g, u). \end{equation} | (4.7) |
We further claim that
\begin{equation} \lim\limits_n\int_\Omega|u_n-u|^p\, dx = 0. \end{equation} | (4.8) |
Indeed, we recall that for all n\in {\mathbb N}
E(g_n, u_n) = \int_\Omega\big[g_nu_n-H(x, u_n)\big]\, dx-\frac{[u_n]_K^p}{p}. |
Therefore, by (4.7), uniform boundedness of (g_n) , (h_2) , and the compact embeddings of W^{s, p}_0(\Omega) into L^1(\Omega) , L^q(\Omega) , respectively, we have for all n\in {\mathbb N}
\begin{align*} \frac{[u_n]_K^p}{p} & \leqslant C+\int_\Omega\big[g_nu_n-H(x, u_n)\big]\, dx \leqslant C+C(\|u_n\|_1+\|u_n\|_q^q) \leqslant C+C([u_n]_K+[u_n]_K^q). \end{align*} |
Since 1 < q < p , the sequence (u_n) is bounded in W^{s, p}_0(\Omega) . Passing to a subsequence, we have u_n\rightharpoonup v in W^{s, p}_0(\Omega) , u_n\to v in L^p(\Omega) , and u_n(x)\to v(x) for a.e. x\in\Omega , as n\to\infty . Therefore, by convexity
\liminf\limits_n[u_n]_K^p \geqslant [v]_K^p. |
Also, by Lemma 2.4 and continuous embeddings we have
\lim\limits_n\, \int_\Omega\big[g_nu_n-H(x, u_n)\big]\, dx = \int_\Omega\big[gv-H(x, v)\big]\, dx. |
So, recalling (4.7), we get
E(g, u) = \lim\limits_n E(g_n, u_n) \leqslant E(g, v), |
which implies u = v by uniqueness of the maximizer in (4.1). So u_n\to u in L^p(\Omega) , which yields (4.8). In addition, for all n\in {\mathbb N} we have
\begin{equation} \Phi(g)+\int_\Omega (g_n-g)u\, dx \leqslant \Phi(g_n) \leqslant \Phi(g)+\int_\Omega(g_n-g)u_n\, dx. \end{equation} | (4.9) |
Indeed, by definition of \Phi(g) we have
\begin{align*} \Phi(g)+\int_\Omega(g_n-g)u\, dx & = E(g_n, u) \leqslant \Phi(g_n) \\ & = E(g, u_n)+\int_\Omega(g_n-g)u_n\, dx \leqslant \Phi(g)+\int_\Omega(g_n-g)u_n\, dx. \end{align*} |
Now fix k\in \overline{\mathcal{G}}\setminus\{g\} and a sequence (\tau_n) in (0, 1) s.t. \tau_n\to 0 as n\to\infty . Set
g_n = g+\tau_n(k-g) \in \overline{\mathcal{G}}, |
so that g_n \overset{*}{\rightharpoonup} g . By (4.9) with such choice of g_n , we have for all n\in {\mathbb N}
\int_\Omega(k-g)u\, dx \leqslant \frac{\Phi(g+\tau_n(k-g))-\Phi(g)}{\tau_n} \leqslant \int_\Omega(k-g)u_n\, dx. |
Passing to the limit for n\to\infty , and noting that by (4.8) we have in particular u_n\to u in L^1(\Omega) , we get
\lim\limits_n \frac{\Phi(g+\tau_n(k-g))-\Phi(g)}{\tau_n} = \int_\Omega(k-g)u\, dx. |
By arbitrariness of (\tau_n) , and noting that u\in L^1(\Omega)\subset L^\infty(\Omega)^* , we see that \Phi is Gâteaux differentiable at g with
\langle\Phi'(g), k-g\rangle = \int_\Omega(k-g)u\, dx, |
which concludes the proof.
We can now prove our optimization result, with a similar argument as in Section 3.
Proof of Theorem 1.2. By Lemma 4.1 and sequential weak* compactness of \overline{\mathcal{G}} , there exists \bar g\in \overline{\mathcal{G}} s.t. for all g\in \overline{\mathcal{G}}
\Phi(\bar g) \geqslant \Phi(g). |
Set \bar u = u_{\bar g}\in W^{s, p}_0(\Omega) , then by Lemma 4.3 we have for all k\in \overline{\mathcal{G}}\setminus\{\bar g\}
\langle\Phi'(\bar g), k-\bar g\rangle = \int_\Omega(k-\bar g)\bar u\, dx. |
Since \bar u\in L^1(\Omega) , by Lemma 2.1 (i) there exists \hat g\in \mathcal{G} s.t. for all g\in \overline{\mathcal{G}}
\int_\Omega\hat g\bar u\, dx \geqslant \int_\Omega g\bar u\, dx, |
in particular
\begin{equation} \int_\Omega\hat g\bar u\, dx \geqslant \int_\Omega\bar g\bar u\, dx. \end{equation} | (4.10) |
By convexity of \Phi (Lemma 4.2) and (4.10), we have
\Phi(\hat g) \geqslant \Phi(\bar g)+\int_\Omega(\hat g-\bar g)\bar u\, dx \geqslant \Phi(\bar g). |
Therefore, \hat g\in \mathcal{G} is a maximizer of \Phi over \overline{\mathcal{G}} , which proves (i). In fact we have \hat g = \bar g , otherwise by strict convexity (Lemma 4.2 again) and (4.10) we would have
\Phi(\hat g) > \Phi(\bar g)+\int_\Omega(\hat g-\bar g)\bar u\, dx \geqslant \Phi(\bar g), |
against maximality of \bar g . Thus, any maximizer of \Phi over \overline{\mathcal{G}} actually lies in \mathcal{G} , which proves (ii). Finally, let \hat g\in \mathcal{G} be a maximizer of \Phi and set \hat u = u_{\hat g}\in W^{s, p}_0(\Omega) . As we have seen before, \hat g is the only maximizer in \overline{\mathcal{G}} for the linear functional
g \mapsto \int_\Omega g\hat u\, dx, |
hence by Lemma 2.1 (ii) there exists a nondecreasing map \eta: {\mathbb R}\to {\mathbb R} s.t. \hat g = \eta\circ\hat u in \Omega , thus proving (iii).
Remark 4.4. Theorem 1.2 is analogous to Theorem 1.1 above, while in fact the problem is easier since we do not need to consider normalization to ensure uniqueness, unlike in problem (1.1). On the other hand, in this case we have no information on the sign of the solution u_g .
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The first author is a member of GNAMPA (Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni) of INdAM (Istituto Nazionale di Alta Matematica 'Francesco Severi'), and is partially supported by the research project Problemi non locali di tipo stazionario ed evolutivo (GNAMPA, CUP E53C23001670001) and the research project Studio di modelli nelle scienze della vita (UniSS DM 737/2021 risorse 2022-2023). We are grateful to the anonymous Referee for their careful reading of our manuscript and useful suggestions.
The authors declare no conflicts of interest.
[1] |
C. Anedda, F. Cuccu, S. Frassu, Steiner symmetry in the minimization of the first eigenvalue of a fractional eigenvalue problem with indefinite weight, Can. J. Math., 73 (2021), 970–992. https://doi.org/10.4153/S0008414X20000267 doi: 10.4153/S0008414X20000267
![]() |
[2] |
C. Bjorland, L. Caffarelli, A. Figalli, Nonlocal tug-of-war and the infinity fractional Laplacian, Commun. Pure Appl. Math., 65 (2012), 337–380. https://doi.org/10.1002/cpa.21379 doi: 10.1002/cpa.21379
![]() |
[3] |
G. R. Burton, Rearrangements of functions, maximization of convex functionals, and vortex rings, Math. Ann., 276 (1987), 225–253. https://doi.org/10.1007/BF01450739 doi: 10.1007/BF01450739
![]() |
[4] |
G. R. Burton, Variational problems on classes of rearrangements and multiple configurations for steady vortices, Ann. Inst. Henri Poincaré, 6 (1989), 295–319. https://doi.org/10.1016/S0294-1449(16)30320-1 doi: 10.1016/S0294-1449(16)30320-1
![]() |
[5] |
G. R. Burton, J. B. McLeod, Maximisation and minimisation on classes of rearrangements, Proc. R. Soc. Edinburgh: Sec. A Math., 119 (1991), 287–300. https://doi.org/10.1017/S0308210500014840 doi: 10.1017/S0308210500014840
![]() |
[6] | L. Caffarelli, Non-local diffusions, drifts and games, In: H. Holden, K. Karlsen, Nonlinear partial differential equations, Abel Symposia, Springer, Berlin, 7 (2012), 37–52. https://doi.org/10.1007/978-3-642-25361-4_3 |
[7] | F. Cuccu, B. Emamizadeh, G. Porru, Optimization of the first eigenvalue in problems involving the p-Laplacian, Proc. Amer. Math. Soc., 137 (2009), 1677–1687. |
[8] |
F. Cuccu, G. Porru, S. Sakaguchi, Optimization problems on general classes of rearrangements, Nonlinear Anal., 74 (2011), 5554–5565. https://doi.org/10.1016/j.na.2011.05.039 doi: 10.1016/j.na.2011.05.039
![]() |
[9] |
L. M. Del Pezzo, A. Quaas, A Hopf's lemma and a strong minimum principle for the fractional p-Laplacian, J. Differ. Equ., 263 (2017), 765–778. https://doi.org/10.1016/j.jde.2017.02.051 doi: 10.1016/j.jde.2017.02.051
![]() |
[10] |
E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker's guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. https://doi.org/10.1016/j.bulsci.2011.12.004 doi: 10.1016/j.bulsci.2011.12.004
![]() |
[11] | G. Franzina, G. Palatucci, Fractional p-eigenvalues, Riv. Mat. Univ. Parma, 5 (2014), 373–386. |
[12] |
A. Iannizzotto, Monotonicity of eigenvalues of the fractional p-Laplacian with singular weights, Topol. Methods Nonlinear Anal., 61 (2023), 423–443. https://doi.org/10.12775/TMNA.2022.024 doi: 10.12775/TMNA.2022.024
![]() |
[13] | A. Iannizzotto, A survey on boundary regularity for the fractional p-Laplacian and its applications, Bruno Pini Math. Anal. Seminar, 15 (2024), 164–186. |
[14] |
A. Iannizzotto, S. Liu, K. Perera, M. Squassina, Existence results for fractional p-Laplacian problems via Morse theory, Adv. Calc. Var., 9 (2016), 101–125. https://doi.org/10.1515/acv-2014-0024 doi: 10.1515/acv-2014-0024
![]() |
[15] |
A. Iannizzotto, S. Mosconi, Fine boundary regularity for the singular fractional p-Laplacian, J. Differ. Equations, 412 (2024), 322–379. https://doi.org/10.1016/j.jde.2024.08.026 doi: 10.1016/j.jde.2024.08.026
![]() |
[16] | A. Iannizzotto, S. Mosconi, On a doubly sublinear fractional p-Laplacian equation, arXiv, 2024. https://doi.org/10.48550/arXiv.2409.03616 |
[17] |
A. Iannizzotto, S. Mosconi, N. S. Papageorgiou, On the logistic equation for the fractional p-Laplacian, Math. Nachr., 296 (2023), 1451–1468. https://doi.org/10.1002/mana.202100025 doi: 10.1002/mana.202100025
![]() |
[18] |
A. Iannizzotto, D. Mugnai, Optimal solvability for the fractional p-Laplacian with Dirichlet conditions, Fract. Calc. Appl. Anal., 27 (2024), 3291–3317. https://doi.org/10.1007/s13540-024-00341-w doi: 10.1007/s13540-024-00341-w
![]() |
[19] |
H. Ishii, G. Nakamura, A class of integral equations and approximation of p-Laplace equations, Calc. Var. Partial Differential Equations, 37 (2010), 485–522. https://doi.org/10.1007/s00526-009-0274-x doi: 10.1007/s00526-009-0274-x
![]() |
[20] |
B. Kawohl, M. Lucia, S. Prashanth, Simplicity of the first eigenvalue for indefinite quasilinear problems, Adv. Differ. Equations, 12 (2007), 407–434. https://doi.org/10.57262/ade/1355867457 doi: 10.57262/ade/1355867457
![]() |
[21] | G. Leoni, A first course in fractional Sobolev spaces, Vol. 229, American Mathematical Society, 2023. |
[22] |
E. Lindgren, P. Lindqvist, Fractional eigenvalues, Calc. Var. Partial Differential Equations, 49 (2014), 795–826. https://doi.org/10.1007/s00526-013-0600-1 doi: 10.1007/s00526-013-0600-1
![]() |
[23] | G. Molica Bisci, V. D. Rădulescu, R. Servadei, Variational methods for nonlocal fractional problems, Cambridge: Cambridge University Press, 2016. https://doi.org/10.1017/CBO9781316282397 |
[24] |
G. Palatucci, The Dirichlet problem for the p-fractional Laplace equation, Nonlinear Anal., 177 (2018), 699–732. https://doi.org/10.1016/j.na.2018.05.004 doi: 10.1016/j.na.2018.05.004
![]() |
[25] |
B. Pellacci, G. Verzini, Best dispersal strategies in spatially heterogeneous environments: optimization of the principal eigenvalue for indefinite fractional Neumann problems, J. Math. Biol., 76 (2018), 1357–1386. https://doi.org/10.1007/s00285-017-1180-z doi: 10.1007/s00285-017-1180-z
![]() |
[26] | C. Qiu, Y. Huang, Y. Zhou, Optimization problems involving the fractional Laplacian, Electron. J. Differ. Eq., 2016 (2016), 1–15. |
[27] | T. R. Rockafellar, Convex analysis, Princeton University Press, 1970. |
[28] | X. Ros-Oton, Nonlocal elliptic equations in bounded domains: a survey, Publ. Mat., 60 (2016), 3–26. |