Accès libre

Symmetrization for Mixed Operators

  
27 avr. 2024
À propos de cet article

Citez
Télécharger la couverture

Introduction
Comparison results: an overview

In [19], Talenti states that if, for given f ≥ 0, fL2(Ω), uH01Ω u \in H_0^1\left(\Omega \right) solves Δu=finΩ,   u=0onΩ, \left\{{\begin{array}{*{20}{l}}{- \Delta u = f}&{{\rm{in}}\;\Omega,}\\{\;\;\;\;\;\;u = 0}&{{\rm{on}}\;\partial \Omega,}\end{array}} \right. and if vH01Ω# v \in H_0^1\left({{\Omega^\#}} \right) solves Δv=f#inΩ#,    v=0onΩ#, \left\{{\begin{array}{*{20}{l}}{- \Delta v = {f^\#}}&{{\rm{in}}\;{\Omega^\#},}\\{\;\;\;\;\;\;v = 0}&{{\rm{on}}\;\partial {\Omega^\#},}\end{array}} \right. then u#va.e.inΩ#. {u^\#} \le v\;\;\;{\rm{a}}{\rm{.e}}{\rm{.}}\;{\rm{in}}\;{\Omega^\#}. Here Ω# is the ball centered at the origin such that |Ω#| = |Ω| and u# denotes the Schwarz symmetrization of u. There are two available proofs of this outcome. The first one, presented by Talenti [19], employs an isoperimetric inequality that concerns the De Giorgi perimeter of Ω. The second proof, on the other hand, was formulated by Lions [17] and does not rely on this particular inequality. Instead, Lions' approach hinges on a differential inequality that relates the distribution functions of u and v.

In [11], it is shown that, for fLp(Ω) where p satisfies some suitable conditions, if u solves the nonlinear and nonlocal problem (Δ)su=finΩ,  u=0on N\Ω, \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}u = f}&{{\rm{in}}\;\Omega,}\\{\;\;\;\;\;\;\;\;\;\;u = 0}&{{\rm{on}}\;{\mathbb{R}^N}\backslash \Omega,}\end{array}} \right. and if v solves the symmetrized problem (Δ)sv=f#inΩ#,  v=0onN\Ω#, \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}v = {f^\#}}&{{\rm{in}}\;{\Omega^\#},}\\{\;\;\;\;\;\;\;\;\;\;v = 0}&{{\rm{on}}\;{\mathbb{R}^N}\backslash {\Omega^\#},}\end{array}} \right. then u#v {u^\#} \prec v where ≺ is the order relation in the form of mass concentration comparison (see Section 2 for precise definitions).

Can the comparison of mass concentration be refined to provide a pointwise estimate? To determine whether a local case result, such as the one proven by Talenti, could also apply to the non-local case, Section 4 in [11] examines certain special cases. The findings reveal that a pointwise estimate cannot be upheld, indicating that the result in [11] is optimal.

Main results

Mixed local and nonlocal problems have gained recent attention and are currently under intensive investigation. The main focus is on an elliptic operator that combines two different orders of differentiation, with the simplest model case being ℒ := −Δ + (−Δ)s for s ∈ (0, 1).

Initial progress in this direction was achieved through probabilistic methods in [9, 10]. More recently, Biagi, Dipierro, Valdinoci, and Vecchi [2, 3, 4, 5] have undertaken a systematic investigation of problems involving mixed operators, with the publication of a number of results concerning regularity and qualitative behavior for solutions, maximum principles, and related variational principles.

In this note, we establish the comparison principle of Talenti in the context of a mixed local/nonlocal elliptic operator: Lu=finΩ,  u=0inn\Ω, \left\{{\begin{array}{*{20}{l}}{\mathcal{L}u = f}&{{\rm{in}}\;\;\Omega,}\\{\;\;\,u = 0}&{{\rm{in}}\;\;{\mathbb{R}^n}\backslash \Omega,}\end{array}} \right. where Ω is a bounded domain in ℝn with smooth boundary, and fL2(Ω) in order to guarantee the existence and the uniqueness of a weak solution.

Theorem 1.1

Let u ∈ 𝕏(Ω) be the weak solution of problem (1.1). Let vH0sΩ# v \in H_0^s\left({{\Omega^\#}} \right) be the weak solution of the symmetrized problem (Δ)sv=f#inΩ#,  v=0onN\Ω#. \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}v = {f^\#}}&{in\;{\Omega^\#},}\\{\;\;\;\;\;\;\;\;\;\;v = 0}&{on\;{\mathbb{R}^N}\backslash {\Omega^\#}.}\end{array}} \right. Then u#v. {u^\#} \prec v.

As applications of Theorem 1.1 we will give the alternative proof of that Faber-Krahn inequality which was proved recently in [5, Theorem 1.1], and in [14, Theorem 4.1], [8, Corollary 1.2] for the local case.

Corollary 1.2 (A Faber-Krahn inequality)

Let Ω be a bounded open subset ofn satisfying#| = |Ω| and let λ1(Ω) denotes the first eigenvalue of the Dirichlet problem (Δ)su=λ1ΩuinΩ,  u=0inn\Ω. \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}u = {\lambda_1}\left(\Omega \right)u}&{in\;\Omega,}\\{\;\;\;\;\;\;\;\;\;\;u = 0}&{in\;{\mathbb{R}^n}\backslash \Omega.}\end{array}} \right. Then we have λ1Ω#λ1Ω. {\lambda_1}\left({{\Omega^\#}} \right) \le {\lambda_1}\left(\Omega \right). Moreover, if the equality holds in (1.3), then Ω is a ball.

Theorem 1.1 enables us to establish a priori estimates for solutions to problem (1.1), expressed in relation to the data f.

Corollary 1.3 (Some regularity)

Let u be the weak solution to problem (1.1). Then uLr(Ω) with r=2NN+4s r = \frac{{2N}}{{N + 4s}} and there exists a constant C such that uLrΩCfL2Ω. {\left\| u \right\|_{{L^r}\left(\Omega \right)}} \le C{\left\| f \right\|_{{L^2}\left(\Omega \right)}}.

This paper is structured as follows. In Section 2, we present preliminaries and useful results about the functional setting, rearrangements, and symmetrization. In Section 3, we provide the proofs for Theorem 1.1, Corollary 1.2, and Corollary 1.3. Additionally, in Section 4, we extend our findings to the fractional p&q–Laplacian operator.

Preliminaries
Functional setting

Let Ω ⊆ ℝn be a bounded open set with continuous boundary. We then consider the space 𝕏(Ω) defined as follows: 𝕏Ω:=uH1n:u0a.e.onn\Ω. {{\mathbb X}}\left(\Omega \right): = \left\{{u \in {H^1}\left({{\mathbb{R}^n}} \right):u \equiv 0\;{\rm{a}}{\rm{.e}}{\rm{.}}\;\;{\rm{on}}\;{\mathbb{R}^n}\backslash \Omega} \right\}.

We can establish that 𝕏(Ω) is a real Hilbert space by using the scalar product defined as: u,v𝕏Ω:=Ωu,vdx. {\left\langle {u,\;v} \right\rangle_{{{\mathbb X}}\left(\Omega \right)}}: = \int_\Omega {\left\langle {\nabla u,\;\nabla v} \right\rangle dx}.

The corresponding norm for this scalar product is given by: u𝕏Ω:=Ωu2dx1/2. {\left\| u \right\|_{{{\mathbb X}}\left(\Omega \right)}}: = {\left({\int_\Omega {{{\left| {\nabla u} \right|}^2}dx}} \right)^{1/2}}.

Additionally, the linear map E0:H01Ω𝕏Ω {\mathcal{E}_0}:H_0^1\left(\Omega \right) \to {{\mathbb X}}\left(\Omega \right) defined by E0u:=u1Ω {\mathcal{E}_0}\left(u \right): = u \cdot {1_\Omega} is a bijective isometry connecting H01Ω H_0^1\left(\Omega \right) and 𝕏(Ω).

On the space 𝕏(Ω), we consider the bilinear form u,v:=Ωu,vdx+2nuxuyvxvy|xy|n+2sdxdy; {\cal B}\left(u,~v \right):=\int_{\Omega}{\left\langle \nabla u,\nabla v \right\rangle dx}+\iint_{{{\mathbb{R}}^{2n}}}{\frac{\left(u\left(x \right)-u\left(y \right) \right)\left(v\left(x \right)-v\left(y \right) \right)}{|x-y{{|}^{n+2s}}}dxdy}; moreover, for every u ∈ 𝕏(Ω) we define Du:=u,u. \mathcal{D}\left(u \right): = {\cal B}\left({u,\;u} \right).

Definition 2.1

Let fL2(Ω). We say that a function u : ℝn → ℝ is a weak solution of problem (1.1), if it satisfies the following properties:

u ∈ 𝕏(Ω);

for every test function φ ∈ 𝕏(Ω), one has u,φ=Ωfφdx. {\cal B}\left({u,\;\varphi} \right) = \int_\Omega {f\varphi dx}.

Applying the Lax–Milgram Theorem to the bilinear form ℬ yields the following existence result, see, [4, Theorem 1.1].

Theorem 2.2

For every fL2(Ω), there exists a unique weak solution u ∈ 𝕏(Ω) of (1.1), further satisfying the ‘a-priori’ estimate u𝕏Ωc0fL2Ω. {\left\| u \right\|_{{{\mathbb X}}\left(\Omega \right)}} \le {c_0}{\left\| f \right\|_{{L^2}\left(\Omega \right)}}. Here, c0 > 0 is a constant independent of f.

We also recall that the solution v to the symmetrized problem (1.2) is radial and radially decreasing, see for instance [2, Theorem 1.1].

Rearrangements and symmetrization
Definition 2.3

Let h : Ω → [0, +∞[ be a measurable function, then the decreasing rearrangement h of h is defined as follows: h*s=inf{t0:xΩ : hx>t<s},s0,Ω. {h^*}\left(s \right) = \;\inf \{t \ge 0:\left| {\left\{{x \in \Omega \;:\;\left| {h\left(x \right)} \right| > t} \right\}} \right| < s\},\;\;\;\;s \in \left[ {0,\;\Omega} \right]. While the Schwartz rearrangement of h is defined as follows h#x=h*(ωn|x|n) ,xΩ#. {h^\#}\left(x \right) = {h^*}({\omega_n}|x{|^n})\;,\;\;\;\;x \in {\Omega^\#}. We denote by ωn the measure of the unit ball in ℝn, and Ω# the ball, centered at the origin, with the same measure as Ω.

It is easily checked that h, h and h# are equi-distributed, i.e. xΩ :hx>t=s0,Ω : h*s>t={xΩ#:h#x>t},t0, \begin{array}{*{20}{l}}{\left| {\left\{{x \in \Omega \;:\;\;\left| {h\left(x \right)} \right| > t} \right\}} \right|}&{= \left| {\left\{{s \in \left({0,\;\;\left| \Omega \right|} \right)\;:\;{h^*}\left(s \right) > t} \right\}} \right|}\\{}&{= \;\left| {\{x \in {\Omega^\#}:{h^\#}\left(x \right) > t\}} \right|,\;\;\;\;t \ge 0,}\end{array} and then if hLP (Ω), 1 ≤ p ≤ ∞, then hLP (0, |Ω|), h#Lp#), and hLpΩ=h*Lp0,Ω=h#LpΩ#. {\left\| h \right\|_{{L^p}\left(\Omega \right)}} = {\left\| {{h^*}} \right\|_{{L^p}\left({0,\left| \Omega \right|} \right)}} = {\left\| {{h^\#}} \right\|_{{L^p}\left({{\Omega^\#}} \right)}}. Moreover, the following inequality, known as Hardy–Littlewood inequality, holds true Ωhxgxdx0Ωh*sg*sds. \int_\Omega {\left| {h\left(x \right)g\left(x \right)} \right|dx} \le \int_0^{\left| \Omega \right|} {{h^*}\left(s \right){g^*}\left(s \right)ds}.

Mass concentration
Definition 2.4

Let f,gLloc1n f,g \in L_{loc}^1\left({{\mathbb{R}^n}} \right) . We say that f is less concentrated than g, and we write fg if for all r > 0 we get Br0f#xdxBr0g#xdx. \int_{{B_r}\left(0 \right)} {{f^\#}\left(x \right){\rm{d}}x} \le \int_{{B_r}\left(0 \right)} {{g^\#}\left(x \right){\rm{d}}x}.

The partial order relationship ≺ is called the comparison of mass concentrations.

Lemma 2.5 ([1, Corollary 2.1])

Let f,gL+1Ω f,g \in L_ +^1\left(\Omega \right) . Then the following are equivalent:

fg;

for all ϕL+Ω \phi \in L_ +^\infty \left(\Omega \right) , ΩfxϕxdxΩ#f#xϕ#xdx, \int_\Omega {f\left(x \right)\phi \left(x \right)dx} \le \int_{{\Omega^\#}} {{f^\#}\left(x \right){\phi^\#}\left(x \right)dx},

for all convex, nonnegative functions Φ: [0, ∞) → [0, ∞) with Φ(0) = 0 it holds that ΩΦfxdxΩΦgxdx. \int_\Omega {\Phi \left({f\left(x \right)} \right)dx} \le \int_\Omega {\Phi \left({g\left(x \right)} \right)dx.}

Proofs
Proof of Theorem 1.1

Let 𝒢t,h, t, h > 0 be the truncation function Gt,hθ=hif θ>t+h,θtif t<θt+h,0if θt. {\mathcal{G}_{t,h}}\left(\theta \right) = \left\{{\begin{array}{*{20}{l}}h&{{\rm{if}}\;\theta > t + h,}\\{\theta - t}&{{\rm{if}}\;t < \theta \le t + h,}\\0&{{\rm{if}}\;\theta \le t.}\end{array}} \right. Let us take φ(x) = 𝒢t,h(u(x)) as a test function in (2.1), we obtain Ωu,Gt,huxdx+2nuxuyGt,huxGt,huy|xy|n+2sdxdy=ΩfxGt,huxdx. \int_{\Omega}{\left\langle \nabla u,\nabla {{\mathcal{G}}_{t,h}}\left(u\left(x \right) \right) \right\rangle dx}+\iint_{{{\mathbb{R}}^{2n}}}{\frac{\left(u\left(x \right)-u\left(y \right) \right)\left({{\mathcal{G}}_{t,h}}\left(u\left(x \right) \right)-{{\mathcal{G}}_{t,h}}\left(u\left(y \right) \right) \right)}{|x-y{{|}^{n+2s}}}dxdy}=\int_{\Omega}{f\left(x \right){{\mathcal{G}}_{t,h}}\left(u\left(x \right) \right)dx}.

It is proven in [11] that 2nuxuyGt,huxGt,huy|xy|n+2sdxdy2nu#xu#yGt,hu#xGt,hu#y|xy|n+2sdxdy. \begin{array}{*{35}{l}}\iint_{{{\mathbb{R}}^{2n}}}{\frac{\left(u\left(x \right)-u\left(y \right) \right)\left({{\mathcal{G}}_{t,h}}\left(u\left(x \right) \right)-{{\mathcal{G}}_{t,h}}\left(u\left(y \right) \right) \right)}{|x-y{{|}^{n+2s}}}dxdy} \\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ge \iint_{{{\mathbb{R}}^{2n}}}{\frac{\left({{u}^{\#}}\left(x \right)-{{u}^{\#}}\left(y \right) \right)\left({{\mathcal{G}}_{t,h}}\left({{u}^{\#}}\left(x \right) \right)-{{\mathcal{G}}_{t,h}}\left({{u}^{\#}}\left(y \right) \right) \right)}{|x-y{{|}^{n+2s}}}dxdy}. \\\end{array} As a consequence we have Ωu,Gt,huxdx+2nu#xu#yGt,hu#xGt,hu#y|xy|n+2sdxdyΩfxGt,huxdx. \begin{array}{*{35}{l}}\int_{\Omega}{\left\langle \nabla u,\nabla {{\mathcal{G}}_{t,h}}\left(u\left(x \right) \right) \right\rangle dx} \\ \ \ \ \ \ \ \ \ \ \ +\iint_{{{\mathbb{R}}^{2n}}}{\frac{\left({{u}^{\#}}\left(x \right)-{{u}^{\#}}\left(y \right) \right)\left({{\mathcal{G}}_{t,h}}\left({{u}^{\#}}\left(x \right) \right)-{{\mathcal{G}}_{t,h}}\left({{u}^{\#}}\left(y \right) \right) \right)}{|x-y{{|}^{n+2s}}}dxdy}\le \int_{\Omega}{f\left(x \right){{\mathcal{G}}_{t,h}}\left(u\left(x \right) \right)dx}. \\\end{array} Letting h go to 0 yields ddtΩt|u|2dx+0rr+μτμρΘN,sτ,ρρN1dρ)τN1dτ0rf*ωNρNρN1dρ, \begin{array}{*{20}{l}}{- \frac{{\rm{d}}}{{{\rm{d}}t}}\left({\int_{{\Omega_t}} {|\nabla u{|^2}{\rm{d}}x}} \right)}\\{\;\;\;\;\; + \int_0^r {\left({\mathop \smallint \nolimits_r^{+ \infty} \left({\mu \left(\tau \right) - \mu \left(\rho \right)} \right){\Theta_{N,s}}\left({\tau,\;\rho} \right){\rho^{N - 1}}{\rm{d}}\rho)} \right){\tau^{N - 1}}\;{\rm{d}}\tau \le \mathop \smallint \nolimits_0^r {f^*}\left({{\omega_N}{\rho^N}} \right){\rho^{N - 1}}{\rm{d}}\rho},}\end{array} where Ωt={xΩ:ux>t}, {\Omega_t} = \{x \in \Omega :\;u\left(x \right) > t\}, and ΘN,sτ,ρ=1NωNx=1y=11|τxρy|N+2sdHN1ydHN1x. {\Theta_{N,s}}\left({\tau,\;\rho} \right) = \frac{1}{{N{\omega_N}}}\int_{\left| {x'} \right| = 1} {\left({\int_{\left| {y'} \right| = 1} {\frac{1}{{|\tau x' - \rho y'{|^{N + 2s}}}}\;{\rm{d}}{H^{N - 1}}\left({y'} \right)}} \right)d{H^{N - 1}}\left({x'} \right)}. From (3.1) we get 0rr+μτμρΘN,sτ,ρρN1dρ)τN1dτ0rf*ωNρNρN1dρ. \int_0^r {\left({\int_r^{+ \infty} {\left({\mu \left(\tau \right) - \mu \left(\rho \right)} \right){\Theta_{N,s}}\left({\tau,\;\rho} \right){\rho^{N - 1}}{\rm{d}}\rho)}} \right){\tau^{N - 1}}\;{\rm{d}}\tau} \le \int_0^r {{f^*}\left({{\omega_N}{\rho^N}} \right){\rho^{N - 1}}\;{\rm{d}}\rho}. The rest of the proof is the same as the proof of [11, Theorem 3.1] (Step 3 and Step 4).

Proof of Corollary 1.2

Let u ∈ 𝕏(Ω) be such that Lu=λ1ΩuinΩ,u=0inn\Ω. \left\{{\begin{array}{*{20}{l}}{\mathcal{L}u = {\lambda_1}\left(\Omega \right)u}&{{\rm{in}}\;\Omega,}\\{\;\;\,u = 0}&{{\rm{in}}\;{\mathbb{R}^n}\backslash \Omega.}\end{array}} \right. Let v ∈ 𝕏(Ω#) be such that (Δ)sv=λ1Ωu#inΩ#,u=0inn\Ω#. \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}v = {\lambda_1}\left(\Omega \right){u^\#}}&{{\rm{in}}\;{\Omega^\#},}\\{\;\;\;\;\;\;\;\;\;\;u = 0}&{{\rm{in}}\;{\mathbb{R}^n}\backslash {\Omega^\#}.}\end{array}} \right. Then by Theorem 1.1, u#v. If we take Φ(t) = t|v| in Lemma 2.5, we get Ω#u#vdxΩ#v2dx. \int_{{\Omega^\#}} {{u^\#}\left| v \right|{\rm{d}}x} \le \int_{{\Omega^\#}} {{{\left| v \right|}^2}{\rm{d}}x}. But by the Rayleigh-quotient characterization of the first eigenvalue, λ1Ω=DΩ#vΩ#u#vDΩ#vΩ#v2dxλ1Ω#. {\lambda_1}\left(\Omega \right) = \frac{{{\mathcal{D}_{{\Omega^\#}}}\left(v \right)}}{{\int_{{\Omega^\#}} {{u^\#}v}}} \ge \frac{{{\mathcal{D}_{{\Omega^\#}}}\left(v \right)}}{{\int_{{\Omega^\#}} {{{\left| v \right|}^2}dx}}} \ge {\lambda_1}\left({{\Omega^\#}} \right). This gives the proof of (1.3).

Proof of Corollary 1.3

We will use [7, Theorem 3.2] the integral form for the solution v to the symmetrized problem (1.2), namely vx=Ω#GΩ#x,yf#ydy, v\left(x \right) = \int_{{\Omega^\#}} {{{\rm{G}}_{{\Omega^\#}}}\left({x,\;y} \right){f^\#}\left(y \right){\rm{d}}y}, where GΩ# is the Green function of the fractional Laplacian on the ball. From [15, Theorem 3.2], we have GΩ#x,yCxyN2s {{\rm{G}}_{{\Omega^\#}}}\left({x,\;y} \right) \le \frac{C}{{{{\left| {x - y} \right|}^{N - 2s}}}} for any xy in Ω#, then, extending f to 0 out of Ω, Hardy-Littlewood-Sobolev inequality [16, Theorem 4.3] implies, uLrΩ=u#LrΩ#vLqΩ#=vLqΩCfL2Ω, {\left\| u \right\|_{{L^r}\left(\Omega \right)}} = {\left\| {{u^\#}} \right\|_{{L^r}\left({{\Omega^\#}} \right)}} \le {\left\| v \right\|_{{L^q}\left({{\Omega^\#}} \right)}} = {\left\| v \right\|_{{L^q}\left(\Omega \right)}} \le C{\left\| f \right\|_{{L^2}\left(\Omega \right)}}, where r=2NN+4s r = \frac{{2N}}{{N + 4s}} .

Further extensions

In [12], it is proven that for p ≥ 2 and fLm(Ω) satisfying certain conditions, if u is a solution to the nonlinear and nonlocal problem Δpsu=finΩ,  u=0onn\Ω, \left\{{\begin{array}{*{20}{l}}{\left({- \Delta} \right)_p^su = f}&{{\rm{in}}\;\Omega,}\\{\;\;\;\;\;\;\;\;\;\;u = 0}&{{\rm{on}}\;{\mathbb{R}^n}\backslash \Omega,}\end{array}} \right. and v is a solution to the symmetrized problem (Δ)sv=g#inΩ#,  v=0onn\Ω#, \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}v = {g^\#}}&{{\rm{in}}\;{\Omega^\#},}\\{\;\;\;\;\;\;\;\;\;\;v = 0}&{{\rm{on}}\;{\mathbb{R}^n}\backslash {\Omega^\#},}\end{array}} \right. then u#v {u^\#} \prec v where g = g(|x|) is the radial function defined by gr=Hn,s,prnsp2p1nsp2p11rnBrf#dx1p1+nωnp1Brf#dx2pp1f#x, g\left(r \right) = {\rm{H}}\left({n,\;s,p} \right)r^{\frac{{\left({n - s} \right)\left({p - 2} \right)}}{{p - 1}}}\left[ {\frac{{\left({n - s} \right)\left({p - 2} \right)}}{{p - 1}}\frac{1}{{{r^n}}}{{\left({\int_{{B_r}} {{f^\#}dx}} \right)}^{\frac{1}{{p - 1}}}} + \frac{{n{\omega_n}}}{{p - 1}}{{\left({\int_{{B_r}} {{f^\#}dx}} \right)}^{\frac{{2 - p}}{{p - 1}}}}{f^\#}\left(x \right)} \right], with Hn,s,p=γn,s,2nωnPsB1p2p1γ(n,s,p)1p1, {\rm{H}}\left({n,\;s,p} \right) = \frac{{\gamma \left({n,s,2} \right)}}{{n{\omega_n}}}\frac{{{{\left({{\mathcal{P}_s}\left({{B_1}} \right)} \right)}^{\frac{{p - 2}}{{p - 1}}}}}}{{\gamma {{(n,s,p)}^{\frac{1}{{p - 1}}}}}}, being PsB1=BrBrc1|xy|n+sdxdy. {\mathcal{P}_s}\left({{B_1}} \right) = \int_{{B_r}} {\int_{B_r^c} {\frac{1}{{|x - y{|^{n + s}}}}dxdy}}.

We establish the comparison principle of Talenti for nonlocal, nonlinear, and nonhomogeneous elliptic problems of the form: Δpsu+Δqsu=finΩ,  u=0onn\Ω, \left\{{\begin{array}{*{20}{l}}{\left({- \Delta} \right)_p^su + \left({- \Delta} \right)_q^su = f}&{{\rm{in}}\;\Omega,}\\{\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\,\;\;\;\;\;u = 0}&{{\rm{on}}\;{\mathbb{R}^n}\backslash \Omega,}\end{array}} \right. where Ω is a bounded domain in ℝn with smooth boundary, 2 ≤ q < p < ∞, and f satisfies suitable conditions to ensure the existence and uniqueness of a weak solution.

In the case of the usual Sobolev spaces, for any 1 ≤ p < q ≤ ∞, it is easy to see that W01,qΩW01,pΩ W_0^{1,q}\left(\Omega \right) \subset W_0^{1,p}\left(\Omega \right) . In the fractional case, this kind of embedding is NOT TRUE. In fact, in [6, Lemma 2.6] it is proved that W0s1,pΩW0s2,qΩforany0<s1<s2<1q<p< W_0^{{s_1},p}\left(\Omega \right) \hookrightarrow W_0^{{s_2},q}\left(\Omega \right)\;\;\;{\rm{for}}\;{\rm{any}}\;0 < {s_1} < {s_2} < 1 \le q < p < \infty this also holds when p = q (see [13, Theorem 2.2]). However, the embedding W0s,pΩW0s,qΩforany0<s<1q<p< W_0^{s,p}\left(\Omega \right) \hookrightarrow W_0^{s,q}\left(\Omega \right)\;\;\;{\rm{for}}\;{\rm{any}}\;0 < s < 1 \le q < p < \infty is not true (see [18, Theorem 1.1]). So, in order to deal with our problem (4.1), we consider the space Ws:=W0s,pΩW0s,qΩ {\mathcal{W}^s}: = W_0^{s,p}\left(\Omega \right) \cap W_0^{s,q}\left(\Omega \right) endowed with the norm [·]s := [·]s,p + [·]s,q.

We say that u ∈ 𝒲s is a weak solution of problem (4.1) if Kn,s,p2nnuxuyp2uxuyφxφy|xy|n+spdxdy+Kn,s,q2nnuxuyq2uxuyφxφy|xy|n+sqdxdy=nfxφxdxforallφWs. \begin{array}{*{20}{l}}{\frac{{K\left({n,s,p} \right)}}{2}\int_{{\mathbb{R}^n}} {\int_{{\mathbb{R}^n}} {\frac{{{{\left| {u\left(x \right) - u\left(y \right)} \right|}^{p - 2}}\left({u\left(x \right) - u\left(y \right)} \right)\left({\varphi \left(x \right) - \varphi \left(y \right)} \right)}}{{|x - y{|^{n + sp}}}}dxdy}}}\\{\;\;\; + \frac{{K\left({n,s,q} \right)}}{2}\int_{{\mathbb{R}^n}} {\int_{{\mathbb{R}^n}} {\frac{{{{\left| {u\left(x \right) - u\left(y \right)} \right|}^{q - 2}}\left({u\left(x \right) - u\left(y \right)} \right)\left({\varphi \left(x \right) - \varphi \left(y \right)} \right)}}{{|x - y{|^{n + sq}}}}dxdy}}}\\{\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; = \int_{{\mathbb{R}^n}} {f\left(x \right)\varphi \left(x \right)dx} \;\;\;{\rm{for}}\;{\rm{all}}\;\varphi \in {\mathcal{W}^s}.}\end{array}

Theorem 4.1

Let u ∈ 𝒲s be the weak solution of problem (4.1). Let v ∈ 𝒲s be the weak solution of the symmetrized problem (Δ)sv=ginΩ#,  v=0onn\Ω#. \left\{{\begin{array}{*{20}{l}}{{{(- \Delta)}^s}v = g}&{in\;{\Omega^\#},}\\{\;\;\;\;\;\;\;\;\;\;v = 0}&{on\;{\mathbb{R}^n}\backslash {\Omega^\#}.}\end{array}} \right. Then u#v. {u^\#} \prec v.

Langue:
Anglais
Périodicité:
2 fois par an
Sujets de la revue:
Mathématiques, Mathématiques générales