Given a bounded open set Ω⊆Rn, we consider the eigenvalue problem for a nonlinear mixed local/nonlocal operator with vanishing conditions in the complement of Ω. We prove that the second eigenvalue λ2(Ω) is always strictly larger than the first eigenvalue λ1(B) of a ball B with volume half of that of Ω. This bound is proven to be sharp, by comparing to the limit case in which Ω consists of two equal balls far from each other. More precisely, differently from the local case, an optimal shape for the second eigenvalue problem does not exist, but a minimizing sequence is given by the union of two disjoint balls of half volume whose mutual distance tends to infinity.
Citation: Stefano Biagi, Serena Dipierro, Enrico Valdinoci, Eugenio Vecchi. A Hong-Krahn-Szegö inequality for mixed local and nonlocal operators[J]. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023014
[1] | Lucrezia Cossetti . Bounds on eigenvalues of perturbed Lamé operators with complex potentials. Mathematics in Engineering, 2022, 4(5): 1-29. doi: 10.3934/mine.2022037 |
[2] | Eleonora Cinti, Roberto Ognibene, Berardo Ruffini . A quantitative stability inequality for fractional capacities. Mathematics in Engineering, 2022, 4(5): 1-28. doi: 10.3934/mine.2022044 |
[3] | Luz Roncal . Hardy type inequalities for the fractional relativistic operator. Mathematics in Engineering, 2022, 4(3): 1-16. doi: 10.3934/mine.2022018 |
[4] | Fernando Farroni, Gioconda Moscariello, Gabriella Zecca . Lewy-Stampacchia inequality for noncoercive parabolic obstacle problems. Mathematics in Engineering, 2023, 5(4): 1-23. doi: 10.3934/mine.2023071 |
[5] | Maurizio Garrione . Beams with an intermediate pier: Spectral properties, asymmetry and stability. Mathematics in Engineering, 2021, 3(2): 1-21. doi: 10.3934/mine.2021016 |
[6] | 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 |
[7] | Pier Domenico Lamberti, Michele Zaccaron . Spectral stability of the curlcurl operator via uniform Gaffney inequalities on perturbed electromagnetic cavities. Mathematics in Engineering, 2023, 5(1): 1-31. doi: 10.3934/mine.2023018 |
[8] | Alessia E. Kogoj, Ermanno Lanconelli, Enrico Priola . Harnack inequality and Liouville-type theorems for Ornstein-Uhlenbeck and Kolmogorov operators. Mathematics in Engineering, 2020, 2(4): 680-697. doi: 10.3934/mine.2020031 |
[9] | Serena Dipierro, Giovanni Giacomin, Enrico Valdinoci . The fractional Malmheden theorem. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023024 |
[10] | Biagio Cassano, Lucrezia Cossetti, Luca Fanelli . Spectral enclosures for the damped elastic wave equation. Mathematics in Engineering, 2022, 4(6): 1-10. doi: 10.3934/mine.2022052 |
Given a bounded open set Ω⊆Rn, we consider the eigenvalue problem for a nonlinear mixed local/nonlocal operator with vanishing conditions in the complement of Ω. We prove that the second eigenvalue λ2(Ω) is always strictly larger than the first eigenvalue λ1(B) of a ball B with volume half of that of Ω. This bound is proven to be sharp, by comparing to the limit case in which Ω consists of two equal balls far from each other. More precisely, differently from the local case, an optimal shape for the second eigenvalue problem does not exist, but a minimizing sequence is given by the union of two disjoint balls of half volume whose mutual distance tends to infinity.
Dedicatoria. Al Ingenioso Hidalgo Don Ireneo.
In this paper we consider a nonlinear operator arising from the superposition of a classical p-Laplace operator and a fractional p-Laplace operator, of the form
Lp,s=−Δp+(−Δ)sp | (1.1) |
with s∈(0,1) and p∈[2,+∞). Here, as usual, Δpu=div(|∇u|p−2∇u), while the fractional p-Laplace operator is defined (up to a multiplicative constant that we neglect) as
(−Δ)spu(x):=2p.v.∫Rn|u(x)−u(y)|p−2(u(x)−u(y))|x−y|n+psdy |
where p.v. stands for the principal value notation.
Given a bounded open set Ω⊆Rn, we consider the eigenvalue problem for the operator Lp,s with homogeneous Dirichlet boundary conditions (i.e., the eigenfunctions are prescribed to vanish in the complement of Ω). In particular, we define λ1(Ω) to be the smallest of such eigenvalues and λ2(Ω) to be the second smallest one (in the sense made precise in [8,29]).
The main result that we present here is a version of the Hong–Krahn–Szegö inequality for the second Dirichlet eigenvalue λ2(Ω), according to the following statement:
Theorem 1.1. Let Ω⊆Rn be a bounded open set. Let B be any Euclidean ball with volume |Ω|/2. Then,
λ2(Ω)>λ1(B). | (1.2) |
Furthermore, equality is never attained in (1.2); however, the estimate is sharp in the following sense: if {xj}j,{yj}j⊆Rn are two sequences such that
limj→+∞|xj−yj|=+∞, |
and if we define Ωj:=Br(xj)∪Br(yj), then
limj→+∞λ2(Ωj)=λ1(Br). | (1.3) |
To the best of our knowledge, Theorem 1.1 is new even in the linear case p=2. Also, an interesting consequence of the fact that equality in (1.2) is never attained is that, for all c>0, the shape optimization problem
inf|Ω|=cλ2(Ω) |
does not admit a solution.
Remark 1.2. We stress that in this paper we deal with the case p≥2. As a matter of fact, as we shall see in Section 4, a key tool for the proof of Theorem 1.1 is the interior regularity of the Lp,s-Dirichlet eigenfunctions (see Section 2 for the relevant definitions); we establish this regularity result by adapting an idea already exploited by Brasco, Lindgren and Schikorra [12] in the purely non-local case, which requires the bound p≥2.
On the other hand, after this paper was completed, the manuscript [28] appeared in the literature, in which the Authors mention a result implying the global Hölder regularity of the Lp,s-Dirichlet eigenfunctions for every p>1; see, precisely, [28,Remark 2.4]. Using this result, one could possibly drop the assumption p≥2 and prove Theorem 1.1 for every p>1.
Before diving into the technicalities of the proof of Theorem 1.1, we devote Section 1.1 to showcase the available results on the shape optimization problems related to the first and the second eigenvalues of several elliptic operators.
One of the classical shape optimization problems is related to the detection of the domain that minimizes the first eigenvalue of the Laplacian with homogeneous boundary conditions. This is the content of the Faber–Krahn inequality [24,32], whose result can be stated by saying that among all domains of fixed volume, the ball has the smallest first eigenvalue.
In particular, as a physical application, one has that that among all drums of equal area, the circular drum possesses the lowest voice, and this somewhat corresponds to our intuition, since a very elongated rectangular drum produces a high pitch related to the oscillations along the short edge.
Another physical consequence of the Faber–Krahn inequality is that among all the regions of a given volume with the boundary maintained at a constant temperature, the one which dissipates heat at the slowest possible rate is the sphere, and this also corresponds to our everyday life experience of spheres minimizing contact with the external environment thus providing the optimal possible insulation.
From the mathematical point of view, the Faber–Krahn inequality also offers a classical stage for rearrangement methods and variational characterizations of eigenvalues.
In view of the discussion in Section A, the subsequent natural question investigates the optimal shape of the second eigenvalue. This problem is addressed by the Hong–Krahn–Szegö inequality [31,33,37], which asserts that among all domains of fixed volume, the disjoint union of two equal balls has the smallest second eigenvalue.
Therefore, for the case of the Laplacian with homogeneous Dirichlet data, the shape optimization problems related to both the first and the second eigenvalues are solvable and the solution has a simple geometry.
It is also interesting to point out a conceptual connection between the Faber–Krahn and the Hong–Krahn–Szegö inequalities, in the sense that the proof of the second typically uses the first one as a basic ingredient. More specifically, the strategy to prove the Hong–Krahn–Szegö inequality is usually:
● Use that in a connected open set all eigenfunctions except the first one must change sign,
● Deduce that λ2(Ω)=max{λ1(Ω+),λ1(Ω−)}, for suitable subdomain Ω+ and Ω− which are either nodal domains for the second eigenfunction, if Ω is connected, or otherwise connected components of Ω,
● Utilize the Faber–Krahn inequality to show that λ1(Ω±) is reduced if we replace Ω± with a ball of volume |Ω±|,
● Employ the homogeneity of the problem to deduce that the volumes of these two balls are equal.
That is, roughly speaking, a cunning use of the Faber–Krahn inequality allows one to reduce to the case of disjoint balls, which can thus be addressed specifically.
A natural extension of the optimal shape results for the Laplacian recalled in Section 1.1.1 is the investigation of the nonlinear operator setting and in particular the case of the p-Laplacian. This line of research was carried out in [10] in which a complete analogue of the results of Section 1.1.1 have been established for the p-Laplacian. In particular, the first Dirichlet eigenvalue of the p-Laplacian is minimized by the ball and the second by any disjoint union of two equal balls.
We stress that, in spite of the similarity of the results obtained, the nonlinear case presents its own specific peculiarities. In particular, in the case of the p-Laplacian one can still define the first eigenvalue by minimization of a Rayleigh quotient, in principle the notion of higher eigenvalues become more tricky, since discreteness of the spectrum is not guaranteed and the eigenvalues theory for nonlinear operators offers plenty of open problems at a fundamental level. For the second eingevalue however one can obtain a variational characterization in terms of a mountain-pass result, still allowing the definition of a spectral gap between the smallest and the second smallest eigenvalue.
We now consider the question posed by the minimization of the first and second eigenvalues in a nonlocal setting.
The optimal shape problems for the first eigenvalue of the fractional Laplacian with homogeneous external datum was addressed in [3,9,11,41], showing that the ball is the optimizer.
As for the nonlinear case, the spectral properties of the fractional p-Laplacian possess their own special features, see [26], and they typically combine the difficulties coming from the nonlocal world with those arising from the theory of nonlinear operators. In [11] the optimal shape problem for the first Dirichlet eigenvalue of the fractional p-Laplacian was addressed, by detecting the optimality of the ball as a consequence of a general Pólya–Szegö principle.
For the second eigenvalue, however, the situation in the nonlocal case is quite different from the classical one, since in general nonlocal energy functionals are deeply influenced by the mutual position of the different connected components of the domain, see [35].
In particular, the counterpart of the Hong–Krahn–Szegö inequality for the fractional Laplacian and the fractional p-Laplacian was established in [13] and it presents significant differences with the classical case: in particular, the shape optimizer for the second eigenvalue of the fractional p-Laplacian with homogeneous external datum does not exist and one can bound such an eigenvalue from below by the first eigenvalue of a ball with half of the volume of the given domain (and this is the best lower bound possible, since the case of a domain consisting of two equal balls drifting away from each other would attain such a bound in the limit).
The study of mixed local/nonlocal operators has been recently received an increasing level of attention, both in view of their intriguing mathematical structure, which combines the classical setting and the features typical of nonlocal operators in a framework that is not scale-invariant [1,4,5,6,8,16,17,20,21,23,27,39], and of their importance in practical applications such as the animal foraging hypothesis [22,36].
In regard to the shape optimization problem, a Faber–Krahn inequality for mixed local and nonlocal linear operators when p=2 has been established in [7], showing the optimality of the ball in the minimization of the first eigenvalue. The corresponding inequality for the nonlinear setting presented in (1.1) will be given here in the forthcoming Theorem 4.1.
The inequality of Hong–Krahn–Szegö type for mixed local and nonlocal linear operators presented in (1.1) would thus complete the study of the optimal shape problems for the first and second eigenvalues of the operator in (1.1).
The rest of this paper is organized as follows. Section 2 sets up the notation and collects some auxiliary results from the existing literature.
In Section 3 we discuss a regularity theory which, in our setting, plays an important role in the proof of Theorem 1.1 in allowing us to speak about nodal regions for the corresponding eigenfunction (recall the bullet point strategy presented on page 3). In any case, this regularity theory holds in a more general setting and can well come in handy in other situations as well.
Section 4 introduces the corresponding Faber–Krahn inequality for the operator in (1.1) and completes the proof of Theorem 1.1.
In Appendix A we also discuss the importance of first and second eigenvalues in general problems of applied mathematics (not necessarily related to partial differential equations, nor to integro-differential equations).
To deal with the nonlinear and mixed local/nonlocal operator in (1.1), given an open and bounded set Ω⊆Rn, it is convenient to introduce the space
X1,p0(Ω)⊆W1,p(Rn), |
defined as the closure of C∞0(Ω) with respect to the global norm
u↦(∫Rn|∇u|pdx)1/p. |
We highlight that, since Ω is bounded, X1,p0(Ω) can be equivalently defined by taking the closure of C∞0(Ω) with respect to the full norm
u↦(∫Rn|u|pdx)1/p+(∫Rn|∇u|pdx)1/p; |
however, we stress that X1,p0(Ω) is different from the usual space W1,p0(Ω), which is defined as the closure of C∞0(Ω) with respect to the norm
u↦(∫Ω|∇u|pdx)1/p. |
As a matter of fact, while the belonging of a function u to W1,p0(Ω) only depends on its behavior inside of Ω (actually, u does not even need to be defined outside of Ω), the belonging of u to X1,p0(Ω) is a global condition, and it depends on the behavior of u on the whole space Rn (in particular, u has to be defined on Rn). Just to give an example of the difference between these spaces, let u∈C∞0(Rn)∖{0} be such that
supp(u)∩¯Ω=∅. |
Since u≡0 inside of Ω, we clearly have that u∈W1,p0(Ω); on the other hand, since u≢0 in Rn∖Ω, one has u∉X1,p0(Ω) (even if u∈W1,p(Rn)).
Although they do not coincide, the spaces X1,p0(Ω) and W1,p0(Ω) are related: to be more precise, using [14,Proposition 9.18] and taking into account the definition of X1,p0(Ω), one can see that
(i) if u∈W1,p0(Ω), then u⋅1Ω∈X1,p0(Ω);
(ii) if u∈X1,p0(Ω), then u|Ω∈W1,p0(Ω).
Moreover, we can actually characterize X1,p0(Ω) as follows:
X1,p0(Ω)={u∈W1,p(Rn):u|Ω∈W1,p0(Ω)andu=0a.e.inRn∖Ω}. |
The main issue in trying to use (ⅰ)–(ⅱ) to identify W1,p0(Ω) with X1,p0(Ω) is that, if u is globally defined and u∈W1,p(Rn), then
u|Ω∈W1,p0(Ω)⇒u⋅1Ω∈X1,p0(Ω); |
however, we cannot say in general that u=u⋅1Ω. Even if they cannot allow to identify X1,p0(Ω) with W1,p0(Ω), assertions (ⅰ)–(ⅱ) can be used to deduce several properties of the space X1,p0(Ω) starting from their analog in W1,p0(Ω); for example, we have the following fact, which shall be used in the what follows:
u∈X1,p0(Ω)⇒|u|,u+,u−∈X1,p0(Ω). |
Remark 2.1. In the particular case when the open set Ω is of class C1, it follows from [14,Proposition 9.18] that, if u∈W1,p(Rn) and u=0 a.e. in Rn∖Ω, then
u|Ω∈W1,p0(Ω). |
As a consequence, we have
X1,p0(Ω)={u∈W1,p(Ω):u=0a.e.inRn∖Ω}. |
This fact shows that, when Ω is sufficiently regular, X1,p0(Ω) coincides with the space Xp(Ω) introduced in [5] (for p=2) and in [8] (for a general p>1).
For future reference, we introduce the following set
M(Ω):={u∈X1,p0(Ω):∫Rn|u|pdx=1}. | (2.1) |
After these preliminaries, we can turn our attention to the Dirichlet problem for the operator Lp,s. Throughout the rest of this paper, to simplify the notation we set
Jp(t):=|t|p−2t for all t∈R. | (2.2) |
Moreover, we define
p∗:={npn−pifp<n,+∞ifp≥n,and(p∗)′:={p∗p∗−1ifp<n,1ifp≥n. |
Definition 2.2. Let q≥(p∗)′, and let f∈Lq(Ω). We say that a function u∈W1,p(Rn) is a weak solution to the equation
Lp,su=finΩ | (2.3) |
if, for every ϕ∈X1,p0(Ω), the following identity is satisfied
∫Ω|∇u|p−2⟨∇u,∇ϕ⟩dx+∬R2nJp(u(x)−u(y))(ϕ(x)−ϕ(y))|x−y|n+psdxdy=∫Ωfϕdx, | (2.4) |
Moreover, given any g∈W1,p(Rn), we say that a function u∈W1,p(Rn) is a weak solution to the (Lp,s)-Dirichlet problem
{Lp,su=finΩ,u=ginRn∖Ω, | (2.5) |
if u is a weak solution to (2.3) and, in addition,
u−g∈X1,p0(Ω). |
Remark 2.3. (1) We point out that the above definition is well-posed: indeed, if u,v∈W1,p(Ω), by Hölder's inequality and [19,Proposition 2.2] we get
∬R2n|u(x)−u(y)|p−1|v(x)−v(y)||x−y|n+psdxdy ≤(∬R2n|u(x)−u(y)|p|x−y|n+psdxdy)1/p(∬R2n|v(x)−v(y)|p|x−y|n+psdxdy)1/p≤c‖u‖W1,p(Rn)‖v‖W1,p(Rn)<+∞. |
Moreover, since f∈Lq(Ω) and q≥(p∗)′, again by Hölder's inequality and by the Sobolev Embedding Theorem (applied here to v∈W1,p(Rn)), we have
∫Ω|f||v|dx≤‖f‖L(p∗)′(Ω)‖v‖Lp∗(Ω)<+∞. |
(2) If W1,p(Rn) is is a weak solution to the (Lp,s)-Dirichlet problem (2.5), it follows from the definition of X1,p0(Ω) that
(u−g)|Ω∈W1,p0(Ω)andu=ga.e.inRn∖Ω. |
Thus, X1,p0(Ω) is the 'right space' for the weak formulation of (2.5).
With Definition 2.2 at hand, we now introduce the notion of Dirichlet eigenvalue/eigenfunction for the operator Lp,s.
Definition 2.4. We say that λ∈R is a Dirichlet eigenvalue for Lp,s if there exists a solution u∈W1,p(Ω)∖{0} of the (Lp,s)-Dirichlet problem
{Lp,su=λ|u|p−2uinΩ,u=0inRn∖Ω. | (2.6) |
In this case, we say that u is an eigenfunction associated with λ.
Remark 2.5. We note that Definition 2.4 is {well-posed}. Indeed, if u is any function in W1,p(Rn), by the Sobolev Embedding Theorem we have
f:=|u|p−2u∈Lp∗p−1(Ω); |
then, a direct computation shows that q:=p∗/(p−1)≥(p∗)′. As a consequence, the notion of weak solution for (2.6) agrees with the one contained in Definition 2.2. In particular, if u is an eigenfunction associated with some eigenvalue λ, then
u∈X1,p0(Ω), |
and thus u|Ω∈W1,p0(Ω) and u=0a.e.inRn∖Ω.
After these definitions, we close the section by reviewing some results about eigenvalues/eigenfucntions for Lp,s which shall be used here below.
To begin with, we recall the following result proved in [8,Proposition 5.1] which establishes the existence of the smallest eigenvalue and detects its basic properties.
Proposition 2.6. The smallest eigenvalue λ1(Ω) for the operator Lp,s is strictly positive and satisfies the following properties:
1) λ1(Ω) is simple;
2) the eigenfunctions associated with λ1(Ω) do not change sign in Rn;
3) every eigenfunction associated to an eigenvalue
λ>λ1(Ω) |
is nodal, i.e., sign changing.
Moreover, λ1(Ω) admits the following variational characterization
λ1(Ω)=minu∈M(Ω)(∫Ω|∇u|pdx+∬R2n|u(x)−u(y)|p|x−y|n+psdxdy), | (2.7) |
where M(Ω) is as in (2.1). The minimum is always attained, and the eigenfunctions for Lp,s associated with λ1(Ω) are precisely the minimizers in (2.7).
We observe that, on account of Proposition 2.6, there exists a unique non-negative eigenfunction u0∈M(Ω)⊆X1,p0(Ω) associated with λ1(Ω); in particular, u0 is a minimizer in (2.7), so that
λ1(Ω)=∫Ω|∇u0|pdx+∬R2n|u0(x)−u0(y)|p|x−y|n+psdxdy. | (2.8) |
We shall refer to u0 as the principal eigenfunction of Lp,s.
The next result was proved in [29,Section 5] and concerns the second eigenvalue for Lp,s.
Theorem 2.7. We define:
λ2(Ω):=inff∈Kmaxu∈Im(f){∫Ω|∇u|pdx+∬R2n|u(x)−u(y)|p|x−y|n+psdxdy}, | (2.9) |
where K:={f:S1→M(Ω):fiscontinuousandodd}, with M(Ω) as in (2.1).
Then:
1) λ2(Ω) is an eigenvalue for Lp,s;
2) λ2(Ω)>λ1(Ω);
3) If λ>λ1(Ω) is an eigenvalue for Lp,s, then λ≥λ2(Ω).
In the rest of this paper, we shall refer to λ1(Ω) and λ2(Ω) as, respectively, the first and second eigenvalue of Lp,s (in Ω). We notice that, as a consequence of (2.7)–(2.9), both λ1(⋅) and λ2(⋅) are translation-invariant, that is,
λ1(x0+Ω)=λ1(Ω) and λ2(x0+Ω)=λ2(Ω). |
To proceed further, we now recall the following global boundedness result for the eigenfunctions of Lp,s (associated with any eigenvalue λ) established in [8,Theorem 4.4].
Theorem 2.8. Let u∈X1,p0(Ω)∖{0} be an eigenfunction for Lp,s, associated with an eigenfunction λ≥λ1(Ω). Then, u∈L∞(Rn).
Remark 2.9. Actually, in [8,Theorem 4.4] it is proved the global boundedness of any non-negative weak solution to the more general Dirichlet problem
{Lp,s=f(x,u)inΩ,u≡0a.e.inRn∖Ω, |
where f:Ω×R→R is a Carathéodory function satisfying the properties
(a) f(⋅,t)∈L∞(Ω) for every t≥0;
(b) There exists a constant cp>0 such that
|f(x,t)|≤cp(1+tp−1)fora.e.x∈Ωandeveryt≥0. |
However, by scrutinizing the proof of the theorem, it is easy to check that the same argument can be applied to our context, where we have
f(x,t)=λ|t|p−2tforallx∈Ωandt∈R, |
but we do not make any assumption on the sign of u (see also [40,Proposition 4]).
Finally, we state here an algebraic lemma which shall be useful in the forthcoming computations.
Lemma 2.10. Let 1<p<+∞ be fixed. Then, the following facts hold.
1) For every a,b∈R such that ab≤0, it holds that
Jp(a−b)a≥{|a|p−(p−1)|a−b|p−2ab,if1<p≤2,|a|p−(p−1)|a|p−2ab,ifp>2. |
2) There exists a constant cp>0 such that
|a−b|p≤|a|p+|b|p+cp(|a|2+|b|2)p−22|ab|,∀a,b∈R. |
In this section we prove the interior Hölder regularity of the eigenfunctions for Lp,s, which is a fundamental ingredient for the proof of Theorem 1.1. As a matter of fact, on account of Theorem 2.8, we establish the interior Hölder regularity for any bounded weak solution of the non-homogeneous equation (2.3), when
f∈L∞(Ω). |
In what follows, we tacitly understand that
2≤p≤nands∈(0,1); |
moreover, Ω⊆Rn is a bounded open set and f∈L∞(Ω).
Remark 3.1. The reason why we restrict ourselves to consider 2≤p≤n follows from the definition of weak solution to (2.3).
Indeed, if u is a weak solution to (2.3), then by definition we have u∈W1,p(Rn); as a consequence, if p>n, by the classical Sobolev Embedding Theorem we can immediately conclude that u∈C0,γ(Rn), where γ=1−n/p.
In order to state (and prove) the main result of this section, we need to fix a notation: for every z∈Rn,ρ>0 and u∈Lp(Rn), we define
Tail(u,z,ρ):=(ρp∫Rn∖Bρ(z)|u|p|x−z|n+psdx)1/p. |
The quantity Tail(u,z,ρ) is referred to as the (Lp,s)-tail of u, see e.g., [18,34].
Theorem 3.2. Let f∈L∞(Ω), and let u∈W1,p(Rn)∩L∞(Rn) be a weak solution to (2.3). Then, there exists some β=β(n,s,p)∈(0,1) such that u∈C0,βloc(Ω).
More precisely, for every ball BR0(z)⋐Ω we have the estimate
[u]pC0,β(BR0(z))≤C(‖f‖L∞(Ω)+‖u‖pL∞(Ω)+Tail(u,z,R1)p+1), | (3.1) |
where
R1:=R0+dist(BR0(z),∂Ω)2 |
and C>0 is a constant independent of u and R1.
In order to prove Theorem 3.2, we follow the approach in [12]; broadly put, the main idea behind this approach is to transfer to the solution u the oscillation estimates proved in [27] for the Lp,s-harmonic functions.
To begin with, we establish the following basic existence/uniqueness result for the weak solutions to the (Lp,s)-Dirichlet problem (2.5).
Proposition 3.3. Let f∈L∞(Ω) and g∈W1,p(Rn) be fixed. Then, there exists a unique solution u=uf,g∈W1,p(Rn) to the Dirichlet problem (2.5).
Proof. We consider the space
W(g):={u∈W1,p(Rn):u−g∈X1,p0(Ω)}, |
and the functional J:W(g)→R defined as follows:
J(u):=1p∫Ω|∇u|pdx+1p∬Ω×Ω|u(x)−u(y)|p|x−y|n+ps+2p∬Ω×(Rn∖Ω)|u(x)−g(y)|p|x−y|n+ps−∫Ωfudx. |
On account of [12,Remark 2.13], we have that J is strictly convex; hence, by using the Direct Methods in the Calculus of Variations, we derive that J has a unique minimizer u=uf,g on W(g), which is the unique weak solution to (2.5).
Thanks to Proposition 3.3, we can prove the following result. Throughout the rest of this paper, if u∈L1loc(Ω) and if A⊆Ω is a measurable set with positive measure, we adopt the classical notation
−∫Au(x)dx:=1|A|∫Au(x)dx. |
In particular, if A=B(x0,r), we set
¯ux0,r:=−∫B(x0,r)u(x)dx. |
Lemma 3.4. Let f∈L∞(Ω) and let u∈W1,p(Rn) be a weak solution to (2.3). Moreover, let B be a given Euclidean ball such that B⋐Ω, and let v∈W1,p(Rn) be the unique weak solution to the Dirichlet problem
{Lp,sv=0inΩ,v=uinRn∖Ω. | (3.2) |
Then, there exists a constant C=C(n,s,p)>0 such that
[u−v]pWs,p(Rn)≤C|B|p′−p′(n−sp)np‖f‖p′L∞(Ω). | (3.3) |
In particular, we have
−∫B|u−v|pdx≤C|B|p′−p′(n−sp)np+spn−1‖f‖p′L∞(Ω). | (3.4) |
Proof. We observe that the existence of v is ensured by Proposition 3.3. Then, taking into account that u is a weak solution to (2.3) and v is the weak solution to (3.2), for every ϕ∈X1,p0(B) we get
∫B(|∇u|p−2⟨∇u,∇ϕ⟩−|∇v|p−2⟨∇v,∇ϕ⟩)dx+∬R2n(Jp(u(x)−u(y))−Jp(v(x)−v(y)))(ϕ(x)−ϕ(y))|x−y|n+psdxdy=∫Bfϕ. |
Choosing, in particular, ϕ:=u−v (notice that, since v is a weak solution of (3.2), by definition we have v−u∈X1,p0(Ω)), we obtain
∫ΩB(∇u,∇v)dx+∬R2n(Jp(t1)−Jp(t2))(t1−t2)|x−y|n+psdxdy=∫Bf(u−v)dx, | (3.5) |
where t1:=u(x)−u(y),t2:=v(x)−v(y) and
B(a,b):=|a|p+|b|p−(|a|p−2+|b|p−2)⟨a,b⟩ for all a,b∈R. |
Now, an elementary computation based on Cauchy-Schwarz's inequality gives
B(a,b)≥0 for all a,b∈R. | (3.6) |
Moreover, since p≥2, by exploiting [12,Remark A.4] we have
(Jp(t1)−Jp(t2))(t1−t2)≥1C|t1−t2|p, | (3.7) |
where C=C(p)>0 is a constant only depending on p. Thus, by combining (3.5), (3.6) and (3.7), we obtain the following estimate:
[u−v]pWs,p(Rn)=∬R2n|t1−t2|p|x−y|n+psdxdy≤C(∫ΩB(∇u,∇v)dx+∬R2n(Jp(t1)−Jp(t2))(t1−t2)|x−y|n+psdxdy)≤C∫Bf(u−v)dx≤C‖f‖L∞(Ω)∫B|u−v|dx≤C|B|1−1p∗s‖f‖L∞(Ω)‖u−v‖Lp∗s(B), |
where we have also used the Hölder's inequality and p∗s>1 is the so-called fractional critical exponent, that is,
p∗s:=npn−sp. |
Finally, by applying the fractional Sobolev inequality to ϕ=u−v (notice that ϕ is compactly supported in B), we get
[u−v]pWs,p(Rn)≤C|B|1−1p∗s‖f‖L∞(Ω)[u−v]Ws,p(Rn), |
and this readily yields the desired (3.3). To prove (3.4) we observe that, by using the Hölder inequality and again the fractional Sobolev inequality, we have
−∫B|u−v|pdx≤(−∫B|u−v|p∗sdx)pp∗s≤C|B|−p∗sp[u−v]pWs,p(Rn); |
thus, estimate (3.4) follows directly from (3.3).
Using Lemma 3.4, we can prove the following excess decay estimate.
Lemma 3.5. Let f∈L∞(Ω) and let u∈W1,p(Rn) be a weak solution to (2.3). Moreover, let x0∈Ω and let R∈(0,1) be such that B4R(x0)⋐Ω.
Then, for every 0<r≤R we have the estimate
−∫Br(x0)|u−¯ux0,r|pdx≤C(Rr)nRγ‖f‖p′L∞(Ω)+C(rR)αp(Rγ‖f‖p′L∞(Ω)+−∫B4R(x0)|u|pdx+Tail(u,x0,4R)p), | (3.8) |
where C,γ and α are positive constants only depending on n, s and p.
Proof. Let v∈W1,p(Rn) be the unique weak solution to the problem
{Lp,sv=0inB3R(x0),v=uonRn∖B3R(x0). | (3.9) |
We stress that the existence of v is guaranteed by Proposition 3.3. We also observe that, for every r∈(0,R], we have that
|¯ux0,r−¯vx0,r|p=|−∫Br(x0)(u−v)dx|p≤−∫Br(x0)|u−v|pdx. |
As a consequence, we obtain
−∫Br(x0)|u−¯ux0,r|pdx≤κ−∫Br(x0)|u−v|pdx+κ−∫Br(x0)|v−¯vx0,r|pdx+κ−∫Br(x0)|¯ux0,r−¯vx0,r|pdx≤κ(−∫Br(x0)|u−v|pdx+−∫Br(x0)|v−¯vx0,r|pdx), | (3.10) |
where κ=κp>0 is a constant only depending on p.
Now, since B3R(x0)⋐Ω and v is the weak solution to (3.9), by Lemma 3.4 we have
−∫Br(x0)|u−v|pdx≤Crnp′−p′(n−sp)p+sp−n‖f‖p′L∞(Ω)≤C(Rr)nRnp′−p′(n−sp)p+sp−n‖f‖p′L∞(Ω). | (3.11) |
On the other hand, since v∈W1,p(Rn) and v is Lp,s-harmonic in B3R(x0) (that is, Lp,sv=0 in the weak sense), we can apply [27,Theorem 5.1], obtaining
−∫Br(x0)|v−¯vx0,r|pdx=−∫Br(x0)|−∫Br(x0)(v(x)−v(y))dy|pdx≤−∫Br(x0)(−∫Br(x0)|v(x)−v(y)|pdy)dx≤(oscBr(x0)v)p≤C(rR)αp(Tail(v,x0,R)p+−∫B2R(x0)|v|pdx), | (3.12) |
where C and α are positive constants only depending on n, s and p. By combining estimates (3.11)-(3.12) with (3.10), we then get
−∫Br(x0)|u−¯ux0,r|pdx≤C(Rr)nRγ‖f‖p′L∞(Ω)+C(rR)αp(Tail(v,x0,R)p+−∫B2R(x0)|v|pdx), | (3.13) |
where we have set
γ:=np′−p′(n−sp)p+sp−n>0. | (3.14) |
To complete the proof of (3.8) we observe that, since u≡v a.e. on Rn∖B3R(x0) (and 0<R≤1), by definition of Tail(v,x0,R) we have
Tail(v,x0,R)p=Rp∫Rn∖BR(x0)|v|p|x−x0|n+psdx=Rp∫Rn∖B4R(x0)|v|p|x−x0|n+psdx+Rp∫B4R(x0)∖BR(x0)|v|p|x−x0|n+psdx≤C(Tail(u,x0,4R)p+−∫B4R(x0)|v|pdx). | (3.15) |
Moreover, by using again Lemma 3.4, we get
−∫B4R(x0)|v|pdx≤C−∫B4R(x0)|u−v|pdx+C−∫B4R(x0)|u|pdx≤C(Rγ‖f‖p′L∞(Ω)+−∫B4R(x0)|u|pdx). | (3.16) |
Thus, by inserting (3.15)-(3.16) into (3.13), we obtain the desired (3.8).
By combining Lemmata 3.4 and 3.5, we can provide the
Proof of Theorem 3.2. The proof follows the lines of [12,Theorem 3.6]. First, we consider a ball BR0(z)⊂⊂Ω and we define the quantities
d:=dist(BR0(z),∂Ω)>0andR1:=d2+R0. | (3.17) |
Thus, we can choose a point x0∈BR0(z) and the ball B4R(x0), where R<min{1,d8}. In particular, this implies that B4R(x0)⊂BR1(z). Since R<1, we can then apply Lemma 3.5: this gives, for every 0<r≤R,
−∫Br(x0)|u−¯ux0,r|pdx≤C(Rr)nRγ‖f‖p′L∞(Ω)+C(rR)αp(Rγ‖f‖p′L∞(Ω)+−∫B4R(x0)|u|pdx+Tail(u,x0,4R)p)≤C(Rr)nRγ‖f‖p′L∞(Ω)+C(rR)αp(dγ‖f‖p′L∞(Ω)+‖u‖pL∞(Ω)dx+Tail(u,x0,4R)p), | (3.18) |
where γ>0 is as in (3.14). Now, we notice that for every x∉BR1(z) it holds that
|x−x0|≥|x−z|−|z−x0|≥R1−|z−x0|R1|x−z|. |
Therefore, we have
Tail(u,x0,4R)p=(4R)p∫Rn∖BR1(z)|u|p|x−x0|n+psdx+(4R)p∫BR1(z)∖B4R(x0)|u|p|x−x0|n+psdx≤(4RR1)p(R1R1−|z−x0|)n+psTail(u,z,R1)p+C‖u‖pL∞(Ω)≤Tail(u,z,R1)p+C‖u‖pL∞(Ω) |
for a constant C depending on n, s and p. We recall that in the last estimate we exploited that
4RR1<d2R0+d2<1and4RR1−|x0−z|≤4RR1−R0<1. |
Consequently, continuing the estimate started with (3.18), we find that
−∫Br(x0)|u−¯ux0,r|pdx≤C(Rr)nRγ‖f‖p′L∞(Ω)+C(rR)αp(dγ‖f‖p′L∞(Ω)+‖u‖pL∞(Ω)dx+Tail(u,z,R1)p). | (3.19) |
We can now define the positive number
θ:=1+γn+αp, |
and take r:=Rθ in (3.19), which yields
r−βp−∫Br(x0)∩BR0(z)|u−¯ux0,r|pdx≤C((dγ+1)‖f‖p′L∞(Ω+‖u‖pL∞(Ω)+Tail(u,z,R1)p), |
where we have set
β:=γαn+αp+γ>0. |
This shows that u∈Lp,n+βγ(BR0(z)), the Campanato space isomorphic to the Hölder space C0,β(¯BR0(z)). This completes the proof of Theorem 3.2.
By gathering together Theorems 2.8 and 3.2, we can easily prove the needed interior Hölder regularity of the eigenfunctions of Lp,s.
Theorem 3.6. Let λ≥λ1(Ω) be an eigenvalue of Lp,s, and let ϕλ∈X1,p0(Ω)∖{0} be an eigenfunction associated with λ. Then, ϕλ∈C(Ω).
Proof. On account of Theorem 2.8, we know that ϕλ∈L∞(Rn). As a consequence, ϕλ is a globally bounded weak solution to (2.3), with
f:=λ|ϕλ|p−2ϕλ∈L∞(Ω). |
We are then entitled to apply Theorem 3.2, which ensures that ϕλ∈C0,βloc(Ω) for some β=β(n,s,p)∈(0,1). This ends the proof of Theorem 3.6.
In this last section of the paper we provide the proof of Theorem 1.1. Before doing this, we establish two preliminary results.
First of all, we prove the following Faber-Krahn type inequality for Lp,s.
Theorem 4.1. Let Ω⊆Rn be a bounded open set, and let m:=|Ω|∈(0,∞). Then, if B(m) is any Euclidean ball with volume m, one has
λ1(Ω)≥λ1(B(m)). | (4.1) |
Moreover, if the equality holds in (4.1), then Ω is a ball.
Proof. The proof is similar to that in the linear case, see [7,Theorem 1.1]; however, we present it here in all the details for the sake of completeness.
To begin with, let ˆB(m) be the Euclidean ball with centre 0 and volume m. Moreover, let u0∈M(Ω) be the principal eigenfunction for Lp,s. We recall that, by definition, u0 is the unique non-negative eigenfunction associated with the first eigenvalue λ1(Ω); in particular, we have (see (2.8))
λ1(Ω)=∫Ω|∇u0|pdx+∬R2n|u0(x)−u0(y)|p|x−y|n+psdxdy. | (4.2) |
Then, we define u∗0:Rn→R as the (decreasing) Schwarz symmetrization of u0. Now, since u0∈M(Ω), from the well-known inequality by Pólya and Szegö (see e.g., [38]) we deduce that
u∗0∈M(ˆB(m))and∫ˆB(m)|∇u∗0|pdx≤∫Ω|∇u|pdx. | (4.3) |
Furthermore, by [2,Theorem 9.2] (see also [25,Theorem A.1]), we also have
∬R2n|u∗0(x)−u∗0(y)|p|x−y|n+psdxdy≤∬R2n|u0(x)−u0(y)|p|x−y|n+psdxdy. | (4.4) |
Gathering all these facts and using (4.2), we get
λ1(Ω)=∫Ω|∇u0|2dx+∬R2n|u0(x)−u0(y)|2|x−y|n+2sdxdy≥∫ˆB(m)|∇u∗0|2dx+∬R2n|u∗0(x)−u∗0(y)|2|x−y|n+2sdxdy≥λ1(ˆB(m)). | (4.5) |
From this, since λ1(⋅) is translation-invariant, we derive the validity of (4.1) for every Euclidean ball B(m) with volume m.
To complete the proof of Theorem 4.1, let us suppose that
λ1(Ω)=λ1(B(m)) |
for some (and hence, for every) ball B(m) with |B(m)|=m. By (4.5) we have
∫Ω|∇u0|pdx+∬R2n|u0(x)−u0(y)|p|x−y|n+psdxdy=λ1(Ω)=λ1(ˆB(m))=∫ˆB(m)|∇(u0)∗|pdx+∬R2n|u∗0(x)−u∗0(y)|p|x−y|n+psdxdy. |
In particular, by (4.3) and (4.4) we get
∬R2n|u0(x)−u0(y)|p|x−y|n+psdxdy=∬R2n|u∗0(x)−u∗0(y)|p|x−y|n+psdxdy. |
We are then in the position to apply once again [25,Theorem A.1], which ensures that u0 must be proportional to a translation of a symmetric decreasing function. As a consequence of this fact, we immediately deduce that
Ω={x∈Rn:u0(x)>0} |
must be a ball (up to a set of zero Lebesgue measure). This completes the proof of Theorem 4.1.
Then, we establish the following lemma on nodal domains.
Lemma 4.2. Let λ>λ1(Ω) be an eigenvalue of Lp,s, and let ϕλ∈X1,p0(Ω)∖{0} be an eigenfunction associated with λ. We define the sets
Ω+:={x∈Ω:ϕλ(x)>0}andΩ−:={x∈Ω:ϕλ(x)<0}. |
Then λ>max{λ1(Ω+),λ1(Ω−)}.
The proof of Lemma 4.2 takes inspiration from [13,Lemma 6.1] (see also [29,Lemma 4.2]).
Proof of Lemma 4.2. First of all, on account of Theorem 3.6 we have that the sets Ω+ and Ω− are open, and therefore the eigenvalues λ1(Ω±) are well–defined.
Moreover, thanks to Proposition 2.6, we know that ϕλ changes sign in Ω, and therefore it is convenient to write ϕλ=ϕ+λ−ϕ−λ, where ϕ+λ and ϕ−λ denote, respectively, the positive and negative parts of ϕλ, with the convention that both the functions ϕ+λ and ϕ−λ are non-negative.
Let us now prove that λ>λ1(Ω+). By using the fact that ϕλ is an eigenfuction of Lp,s corresponding to λ, it follows that
∫Ω|∇ϕλ|p−2⟨∇ϕλ,v⟩dx+∬R2n|ϕλ(x)−ϕλ(y)|p−2(ϕλ(x)−ϕλ(y))(v(x)−v(y)|x−y|n+psdxdy=λ∫Ω|ϕλ|p−2ϕλvdx, for all v∈X1,p0(Ω). |
In consideration of the fact that ϕ+λ∈X1,p0(Ω), we can take v=ϕ+λ as a test function.
Now, since
ϕ+λ(x)ϕ−λ(x)=0for a.e.x∈Ω, |
we easily get that
(ϕ+λ(x)−ϕ+λ(y))(ϕ−λ(x)−ϕ−λ(y))≤0. |
Moreover, since both Ω+ and Ω− are non-void open set (remind that ϕλ is continuous on Ω and it changes sign in Ω), we have
∬R2n|ϕλ(x)−ϕλ(y)|p−2(ϕ+λ(x)−ϕ+λ(y))(ϕ−λ(x)−ϕ−λ(y))|x−y|n+psdxdy≤−∫Ω+∫Ω−|ϕλ(x)−ϕλ(y)|p−2ϕ+λ(x)ϕ−λ(y)|x−y|n+psdxdy<0 |
and
∬R2n|ϕ+λ(x)−ϕ+λ(y)|p−2(ϕ+λ(x)−ϕ+λ(y))(ϕ−λ(x)−ϕ−λ(y))|x−y|n+psdxdy≤−∫Ω+∫Ω−|ϕ+λ(x)|p−2ϕ+λ(x)ϕ−λ(y)|x−y|n+psdxdy<0. |
We can therefore exploit Lemma 2.10-(1) with
a:=ϕ+λ(x)−ϕ+λ(y)andb:=ϕ−λ(x)−ϕ−λ(y), |
obtaining (remind that, by assumption, p≥2)
λ∫Ω+|ϕ+λ|pdx=λ∫Ω|ϕλ|p−2ϕλϕ+λdx=∫Ω|∇ϕλ|p−2⟨∇ϕλ,∇ϕ+λ⟩dx+∬R2n|ϕλ(x)−ϕλ(y)|p−2(ϕλ(x)−ϕλ(y))(ϕ+λ(x)−ϕ+λ(y)|x−y|n+psdxdy=∫Ω+|∇ϕ+λ|pdx+∬R2n|ϕλ(x)−ϕλ(y)|p−2(ϕλ(x)−ϕλ(y))(ϕ+λ(x)−ϕ+λ(y)|x−y|n+psdxdy>∫Ω+|∇ϕ+λ|pdx+∬R2n|ϕ+λ(x)−ϕ+λ(y)|p|x−y|n+psdxdy≥λ1(Ω+)∫Ω+|ϕ+λ|pdx, |
where we used the variational characterization of λ1(Ω+), see (2.7). In particular, this gives that λ>λ1(Ω+). With a similar argument (see e.g., [13,Lemma 6.1]), one can show that λ>λ1(Ω−) as well, and this closes the proof of Lemma 4.2.
By virtue of Theorem 4.1 and Lemma 4.2, we can provide the
Proof of Theorem 1.1. We split the proof into two steps.
Step Ⅰ: In this step, we prove inequality (1.2). To this end, let ϕ∈M(Ω) be a Lp-normalized eigenfunction associated with λ2(Ω) (recall the definition of the space M(Ω) in (2.1)). On account of Theorem 2.7, we know that ϕ∈C(Ω).
Moreover, since ϕ changes sign in Ω (see Proposition 2.6), we can define the non-void open sets
Ω+:={u>0}andΩ−:={u<0}. |
Then, by combining Lemma 4.2 with Theorem 4.1, we get
λ2(Ω)>max{λ1(B+),λ1(B−)}, | (4.6) |
where B+ is a Euclidean ball with volume equal to |Ω+| and B− is a Euclidean ball with volume |Ω−|.
Now, since Ω+∪Ω−=Ω, we have
|B+|+|B−|=|Ω+|+|Ω−|≤|Ω|=m. |
Taking into account this inequality, we claim that
max{λ1(B+),λ1(B−)}≥λ1(B), | (4.7) |
being B a ball of volume |Ω|/2. In order to prove (4.7), we distinguish three cases.
(i) |B+|,|B−|≤m/2. In this case, since λ1(⋅) is translation-invariant, we can assume without loss of generality that B⊆B+,B−; as a consequence, since λ1(⋅) is non-increasing, we obtain
λ1(B+),λ1(B−)≥λ1(B), |
and this proves the claimed (4.7).
(ii) |B−|<m/2<|B+|. In this case, we can assume that B−⊆B⊆B+; from this, since λ1(⋅) is non-increasing, we obtain
λ1(B+)≥λ1(B)≥λ1(B−), |
and this immediately implies the claimed (4.7).
(iii) |B+|<m/2<|B−|. In this last case, it suffices to interchange the roles of the balls B− and B+, and to argue exactly as in case (ii).
Gathering (4.6) and (4.7), we obtain the claim in (1.2).
Step Ⅱ: Now we prove the sharpness of (1.2). To this end, according to the statement of the theorem, we fix r>0 and we define
Ωj:=Br(xj)∪Br(yj), |
where {xj}j,{yj}j⊆Rn are two sequences satisfying
limj→+∞|xj−yj|=+∞. | (4.8) |
On account of (4.8), we can assume that
Br(xj)∩Br(yj)=∅ for all j≥1. | (4.9) |
Let now u0∈M(Br) be a Lp-normalized eigenfunction associated with λ1(Br) (here, Br=Br(0)). For every natural number j≥1, we set
ϕj(x):=u0(x−xj)andψj(x):=u0(x−yj). | (4.10) |
Since λ1(⋅) is translation-invariant, it is immediate to check that ϕj and ψj are normalized eigenfunctions associated with λ1(Br(xj)) and λ1(Br(yj)), respectively.
Moreover, taking into account (4.9), it is easy to see that
ϕj≡0onRn∖Br(xj)⊇Br(yj)andψj≡0onRn∖Br(yj)⊇Br(xj) | (4.11) |
and ϕjψj≡0 on Rn.
We then consider the function f defined as follows:
f(z1,z2):=|z1|2−ppz1ϕj−|z2|2−ppz2ψjwithz=(z1,z2)∈S1. |
Taking into account that Br(xj),Br(yj)⊆Ωj and u0∈M(Br), it is readily seen that f(S1)⊆X1,p0(Ωj).
Furthermore, the function f is clearly odd and continuous. Also, using (4.9) and the fact that ϕ≡0 out of Br, one has
‖f(z1,z2)‖pLp(Ωj)=‖|z1|2−ppz1ϕj−|z2|2−ppz2ψj‖pLp(Ωj)=|z1|2‖ϕj‖pLp(Br(xj))+|z2|2‖ψj‖pLp(Br(yj))=(|z1|2+|z2|2)‖u0‖pLp(Br)=1. |
We are thereby entitled to use f in the definition of λ2(Ω), see (2.9): setting aj:=ϕj(x)−ϕj(y) and bj:=ψj(x)−ψj(y) to simplify the notation, this gives, together with (4.9) and (4.11), that
λ2(Ωj)≤maxv∈Im(f){∫Ωj|∇v|pdx+∬R2n|v(x)−v(y)|p|x−y|n+psdxdy}=max|ω1|p+|ω2|p=1{∫Ωj|∇(ω1ϕj−ω2ψj)|pdx+∬R2n|ω1aj−ω2bj|p|x−y|n+psdxdy}=max|ω1|p+|ω2|p=1{|ω1|p∫Br(xj)|∇ϕj|pdx+|ω2|p∫Br(yj)|∇ψj|pdx+∬R2n|ω1aj−ω2bj|p|x−y|n+psdxdy}=max|ω1|p+|ω2|p=1{∫Br|∇u0|pdx+∬R2n|ω1aj−ω2bj|p|x−y|n+ps}. |
On the other hand, by applying Lemma 2.10-(2), we get
max|ω1|p+|ω2|p=1{∫Br|∇u0|pdx+∬R2n|ω1aj−ω2bj|p|x−y|n+ps}≤max|ω1|p+|ω2|p=1{∫Br|∇u0|pdx+|ω1|p∬R2n|ϕj(x)−ϕj(y)|p|x−y|n+psdxdy+|ω2|p∬R2n|ψj(x)−ψj(y)|p|x−y|n+psdxdy+cp∬R2n|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj||x−y|n+psdxdy}=max|ω1|p+|ω2|p=1{∫Br|∇u0|pdx+∬R2n|u0(x)−u0(y)|p|x−y|n+psdxdy+cp∬R2n|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj||x−y|n+psdxdy}=λ1(Br)+cpmax|ω1|p+|ω2|p∬R2n|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj||x−y|n+psdxdy, |
where we have also used that u0 is a normalized eigenfunction associated with the first eigenvalue λ1(Br).
Summarizing, we have proved that
λ2(Ωj)≤λ1(Br)+cpmax|ω1|p+|ω2|p∬R2n|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj||x−y|n+psdxdy. | (4.12) |
We now set
Rj:=max|ω1|p+|ω2|p=1∬R2n|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj||x−y|n+psdxdy |
and we claim that Rj→0 as j→+∞.
Indeed, since ϕjψj≡0 on Rn, we have that
ajbj=−ψj(x)ψj(y)−ϕj(y)ψj(x). |
As a consequence, recalling (4.11)
0≤Rj≤2max|ω1|p+|ω2|p=1∫Br(xj)∫Br(yj)|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj||x−y|n+psdxdy≤2|xj−yj|−2rmax|ω1|p+|ω2|p=1∫Br(xj)∫Br(yj)|(ω1aj)2+(ω2bj)2|p−22|ω1ω2ajbj|dxdy=2|xj−yj|−2r∫Br∫Br|u0(x)2+u0(y)2|p−22|u0(x)u0(y)|dxdy=:2c0|xj−yj|−2r. |
Taking into account (4.8), we thereby conclude that
limj→+∞Rj=0. | (4.13) |
Gathering together (4.12) and (4.13), we obtain the desired result in (1.3).
The authors are members of INdAM. S. Biagi is partially supported by the INdAM-GNAMPA project Metodi topologici per problemi al contorno associati a certe classi di equazioni alle derivate parziali. S. Dipierro and E. Valdinoci are members of AustMS. S. Dipierro is supported by the Australian Research Council DECRA DE180100957 PDEs, free boundaries and applications. E. Valdinoci is supported by the Australian Laureate Fellowship FL190100081 Minimal surfaces, free boundaries and partial differential equations. E. Vecchi is partially supported by the INdAM-GNAMPA project Convergenze variazionali per funzionali e operatori dipendenti da campi vettoriali.
The authors declare no conflict of interest.
[1] |
N. Abatangelo, M. Cozzi, An elliptic boundary value problem with fractional nonlinearity, SIAM J. Math. Anal., 53 (2021), 3577–3601. http://dx.doi.org/10.1137/20M1342641 doi: 10.1137/20M1342641
![]() |
[2] |
F. J. Almgren, E. H. Lieb, Symmetric decreasing rearrangement is sometimes continuous, J. Amer. Math. Soc., 2 (1989), 683–773. http://dx.doi.org/10.1090/S0894-0347-1989-1002633-4 doi: 10.1090/S0894-0347-1989-1002633-4
![]() |
[3] |
R. Bañuelos, R. Latała, P. J. Méndez-Hernández, A Brascamp-Lieb-Luttinger-type inequality and applications to symmetric stable processes, Proc. Amer. Math. Soc., 129 (2001), 2997–3008. http://dx.doi.org/10.1090/S0002-9939-01-06137-8 doi: 10.1090/S0002-9939-01-06137-8
![]() |
[4] |
G. Barles, C. Imbert, Second-order elliptic integro-differential equations: viscosity solutions' theory revisited, Ann. Inst. H. Poincaré Anal. Non Linéaire, 25 (2008), 567–585. http://dx.doi.org/10.1016/j.anihpc.2007.02.007 doi: 10.1016/j.anihpc.2007.02.007
![]() |
[5] | S. Biagi, S. Dipierro, E. Valdinoci, E. Vecchi, Mixed local and nonlocal elliptic operators: regularity and maximum principles, Commun. Part. Diff. Eq., in press. http://dx.doi.org/10.1080/03605302.2021.1998908 |
[6] |
S. Biagi, E. Vecchi, S. Dipierro, E. Valdinoci, Semilinear elliptic equations involving mixed local and nonlocal operators, Proc. Roy. Soc. Edinb. A, 151 (2021), 1611–1641. http://dx.doi.org/10.1017/prm.2020.75 doi: 10.1017/prm.2020.75
![]() |
[7] | S. Biagi, S. Dipierro, E. Valdinoci, E. Vecchi, A Faber-Krahn inequality for mixed local and nonlocal operators, J. Anal. Math., in press. |
[8] | S. Biagi, D. Mugnai, E. Vecchi, Global boundedness and maximum principle for a Brezis-Oswald approach to mixed local and nonlocal operators, 2021, arXiv: 2103.11382. |
[9] |
L. Brasco, E. Cinti, S. Vita, A quantitative stability estimate for the fractional Faber-Krahn inequality, J. Funct. Anal., 279 (2020), 108560. http://dx.doi.org/10.1016/j.jfa.2020.108560 doi: 10.1016/j.jfa.2020.108560
![]() |
[10] |
L. Brasco, G. Franzina, On the Hong-Krahn-Szego inequality for the p-Laplace operator, Manuscripta Math., 141 (2013), 537–557. http://dx.doi.org/10.1007/s00229-012-0582-x doi: 10.1007/s00229-012-0582-x
![]() |
[11] |
L. Brasco, E. Lindgren, E. Parini, The fractional Cheeger problem, Interfaces Free Bound., 16 (2014), 419–458. http://dx.doi.org/10.4171/IFB/325 doi: 10.4171/IFB/325
![]() |
[12] |
L. Brasco, E. Lindgren, A. Schikorra, Higher Hölder regularity for the fractional p-Laplacian in the superquadratic case, Adv. Math., 338 (2018), 782–846. http://dx.doi.org/10.1016/j.aim.2018.09.009 doi: 10.1016/j.aim.2018.09.009
![]() |
[13] |
L. Brasco, E. Parini, The second eigenvalue of the fractional p-Laplacian, Adv. Calc. Var., 9 (2016), 323–355. http://dx.doi.org/10.1515/acv-2015-0007 doi: 10.1515/acv-2015-0007
![]() |
[14] | H. Brezis, Functional analysis, Sobolev spaces and partial differential equations, New York: Springer, 2011. http://dx.doi.org/10.1007/978-0-387-70914-7 |
[15] | S. Brin, L. Page, The anatomy of a large-scale hypertextual Web search engine, Computer Networks and ISDN Systems, 30 (1998), 107–117. |
[16] |
X. Cabré, S. Dipierro, E. Valdinoci, The Bernstein technique for integro-differential equations, Arch. Rational Mech. Anal., 243 (2022), 1597–1652. http://dx.doi.org/10.1007/s00205-021-01749-x doi: 10.1007/s00205-021-01749-x
![]() |
[17] |
Z.-Q. Chen, P. Kim, R. Song, Z. Vondraček, Boundary Harnack principle for Δ+Δα/2, Trans. Amer. Math. Soc., 364 (2012), 4169–4205. http://dx.doi.org/10.1090/S0002-9947-2012-05542-5 doi: 10.1090/S0002-9947-2012-05542-5
![]() |
[18] |
A. Di Castro, T. Kuusi, G. Palatucci, Nonlocal Harnack inequalities, J. Funct. Anal., 267 (2014), 1807–1836. http://dx.doi.org/10.1016/j.jfa.2014.05.023 doi: 10.1016/j.jfa.2014.05.023
![]() |
[19] |
E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker's guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. http://dx.doi.org/10.1016/j.bulsci.2011.12.004 doi: 10.1016/j.bulsci.2011.12.004
![]() |
[20] | S. Dipierro, E. Proietti Lippi, E. Valdinoci, Linear theory for a mixed operator with Neumann conditions, Asymptotic Anal., in press. http://dx.doi.org/10.3233/ASY-211718 |
[21] | S. Dipierro, E. Proietti Lippi, E. Valdinoci, (Non)local logistic equations with Neumann conditions, Ann. Inst. H. Poincaré Anal. Non Linéaire, in press. |
[22] |
S. Dipierro, E. Valdinoci, Description of an ecological niche for a mixed local/nonlocal dispersal: an evolution equation and a new Neumann condition arising from the superposition of Brownian and Lévy processes, Physica A, 575 (2021), 126052. http://dx.doi.org/10.1016/j.physa.2021.126052 doi: 10.1016/j.physa.2021.126052
![]() |
[23] | B. C. dos Santos, S. M. Oliva, J. D. Rossi, A local/nonlocal diffusion model, Appl. Anal., in press. http://dx.doi.org/10.1080/00036811.2021.1884227 |
[24] | G. Faber, Beweis, dass unter allen homogenen Membranen von gleicher Fläche und gleicher Spannung die kreisförmige den tiefsten Grundton gibt, München: Sitzungsberichte, 1923,169–172. |
[25] |
R. L. Frank, R. Seiringer, Non-linear ground state representations and sharp Hardy inequalities, J. Funct. Anal., 255 (2008), 3407–3430. http://dx.doi.org/10.1016/j.jfa.2008.05.015 doi: 10.1016/j.jfa.2008.05.015
![]() |
[26] | G. Franzina, G. Palatucci, Fractional p-eigenvalues, Riv. Math. Univ. Parma, 5 (2014), 373–386. |
[27] | P. Garain, J. Kinnunen, On the regularity theory for mixed local and nonlocal quasilinear elliptic equations, 2021, arXiv: 2108.02986. |
[28] |
J. Giacomoni, D. Kumar, K. Sreenadh, Global regularity results for non-homogeneous growth fractional problems, J. Geom. Anal., 32 (2022), 36. http://dx.doi.org/10.1007/s12220-021-00837-4 doi: 10.1007/s12220-021-00837-4
![]() |
[29] |
D. Goel, K. Sreenadh, On the second eigenvalue of combination between local and nonlocal p-Laplacian, Proc. Amer. Math. Soc., 147 (2019), 4315–4327. http://dx.doi.org/10.1090/proc/14542 doi: 10.1090/proc/14542
![]() |
[30] | D. Hilbert, Grundzüge einer allgeminen Theorie der linaren Integralrechnungen. (Erste Mitteilung), Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch–Physika lische Klasse, 1904, 49–91. |
[31] |
I. Hong, On an inequality concerning the eigenvalue problem of membrane, Kōdai Math. Sem. Rep., 6 (1954), 113–114. http://dx.doi.org/10.2996/kmj/1138843535 doi: 10.2996/kmj/1138843535
![]() |
[32] |
E. Krahn, Über eine von Rayleigh formulierte Minimaleigenschaft des Kreises, Math. Ann., 94 (1925), 97–100. http://dx.doi.org/10.1007/BF01208645 doi: 10.1007/BF01208645
![]() |
[33] | E. Krahn, Über Minimaleigenschaften der Kugel in drei und mehr Dimensionen, Acta Univ. Dorpat A, 9 (1926), 1–44. |
[34] |
T. Kuusi, G. Mingione, Y. Sire, Nonlocal equations with measure data, Commun. Math. Phys., 337 (2015), 1317–1368. http://dx.doi.org/10.1007/s00220-015-2356-2 doi: 10.1007/s00220-015-2356-2
![]() |
[35] |
E. Lindgren, P. Lindqvist, Fractional eigenvalues, Calc. Var., 49 (2014), 795–826. http://dx.doi.org/10.1007/s00526-013-0600-1 doi: 10.1007/s00526-013-0600-1
![]() |
[36] |
G. Pagnini, S. Vitali, Should I stay or should I go? Zero-size jumps in random walks for Lévy flights, Fract. Calc. Appl. Anal., 24 (2021), 137–167. http://dx.doi.org/10.1515/fca-2021-0007 doi: 10.1515/fca-2021-0007
![]() |
[37] |
G. Pólya, On the characteristic frequencies of a symmetric membrane, Math. Z., 63 (1955), 331–337. http://dx.doi.org/10.1007/BF01187944 doi: 10.1007/BF01187944
![]() |
[38] | G. Pólya, G. Szegö, Isoperimetric inequalities in mathematical physics, Princeton, N.J.: Princeton University Press, 1951. |
[39] | A. M. Salort, E. Vecchi, On the mixed local–nonlocal Hénon equation, 2021, arXiv: 2107.09520. |
[40] |
R. Servadei, E. Valdinoci, A Brezis-Nirenberg result for non-local critical equations in low dimension, Commun. Pure Appl. Anal., 12 (2013), 2445–2464. http://dx.doi.org/10.3934/cpaa.2013.12.2445 doi: 10.3934/cpaa.2013.12.2445
![]() |
[41] |
Y. Sire, J. L. Vázquez, B. Volzone, Symmetrization for fractional elliptic and parabolic equations and an isoperimetric application, Chin. Ann. Math. Ser. B, 38 (2017), 661–686. http://dx.doi.org/10.1007/s11401-017-1089-2 doi: 10.1007/s11401-017-1089-2
![]() |
1. | Stefano Biagi, Dimitri Mugnai, Eugenio Vecchi, Necessary condition in a Brezis–Oswald-type problem for mixed local and nonlocal operators, 2022, 132, 08939659, 108177, 10.1016/j.aml.2022.108177 | |
2. | Cristiana De Filippis, Giuseppe Mingione, Gradient regularity in mixed local and nonlocal problems, 2022, 0025-5831, 10.1007/s00208-022-02512-7 | |
3. | Xiangrui Li, Shuibo Huang, Meirong Wu, Canyun Huang, Existence of solutions to elliptic equation with mixed local and nonlocal operators, 2022, 7, 2473-6988, 13313, 10.3934/math.2022735 | |
4. | Prashanta Garain, Erik Lindgren, Higher Hölder regularity for mixed local and nonlocal degenerate elliptic equations, 2023, 62, 0944-2669, 10.1007/s00526-022-02401-6 | |
5. | G.C. Anthal, J. Giacomoni, K. Sreenadh, A Choquard type equation involving mixed local and nonlocal operators, 2023, 527, 0022247X, 127440, 10.1016/j.jmaa.2023.127440 | |
6. | Serena Dipierro, Kanishka Perera, Caterina Sportelli, Enrico Valdinoci, An existence theory for superposition operators of mixed order subject to jumping nonlinearities, 2024, 37, 0951-7715, 055018, 10.1088/1361-6544/ad3793 | |
7. | Alberto Maione, Dimitri Mugnai, Eugenio Vecchi, Variational methods for nonpositive mixed local–nonlocal operators, 2023, 26, 1311-0454, 943, 10.1007/s13540-023-00147-2 | |
8. | Sabri Bahrouni, Symmetrization for Mixed Operators, 2024, 2391-4238, 10.2478/amsil-2024-0013 | |
9. | João Vitor da Silva, Alessio Fiscella, Victor A. Blanco Viloria, Mixed local-nonlocal quasilinear problems with critical nonlinearities, 2024, 408, 00220396, 494, 10.1016/j.jde.2024.07.028 | |
10. | Gianmarco Giovannardi, Dimitri Mugnai, Eugenio Vecchi, An Ahmad-Lazer-Paul-type result for indefinite mixed local-nonlocal problems, 2023, 527, 0022247X, 127442, 10.1016/j.jmaa.2023.127442 | |
11. | Serena Dipierro, Edoardo Proietti Lippi, Enrico Valdinoci, Some maximum principles for parabolic mixed local/nonlocal operators, 2024, 0002-9939, 10.1090/proc/16899 | |
12. | Sun-Sig Byun, Deepak Kumar, Ho-Sik Lee, Global gradient estimates for the mixed local and nonlocal problems with measurable nonlinearities, 2024, 63, 0944-2669, 10.1007/s00526-023-02631-2 | |
13. | Xifeng Su, Enrico Valdinoci, Yuanhong Wei, Jiwen Zhang, On some regularity properties of mixed local and nonlocal elliptic equations, 2025, 416, 00220396, 576, 10.1016/j.jde.2024.10.003 | |
14. | Nicolò Cangiotti, Maicol Caponi, Alberto Maione, Enzo Vitillaro, Klein–Gordon–Maxwell Equations Driven by Mixed Local–Nonlocal Operators, 2023, 91, 1424-9286, 375, 10.1007/s00032-023-00387-0 | |
15. | Prashanta Garain, Wontae Kim, Juha Kinnunen, On the regularity theory for mixed anisotropic and nonlocal p-Laplace equations and its applications to singular problems, 2024, 36, 0933-7741, 697, 10.1515/forum-2023-0151 | |
16. | Shuibo Huang, Hichem Hajaiej, Lazer-mckenna type problem involving mixed local and nonlocal elliptic operators, 2025, 32, 1021-9722, 10.1007/s00030-024-01007-5 | |
17. | Lovelesh Sharma, Brezis Nirenberg type results for local non-local problems under mixed boundary conditions, 2024, 16, 2836-3310, 872, 10.3934/cam.2024038 | |
18. | Dimitri Mugnai, Edoardo Proietti Proietti Lippi, Quasilinear Fractional Neumann Problems, 2024, 13, 2227-7390, 85, 10.3390/math13010085 | |
19. | Carlo Alberto Antonini, Matteo Cozzi, Global gradient regularity and a Hopf lemma for quasilinear operators of mixed local-nonlocal type, 2025, 425, 00220396, 342, 10.1016/j.jde.2025.01.030 |