The aim of this work is to show a non-sharp quantitative stability version of the fractional isocapacitary inequality. In particular, we provide a lower bound for the isocapacitary deficit in terms of the Fraenkel asymmetry. In addition, we provide the asymptotic behaviour of the s-fractional capacity when s goes to 1 and the stability of our estimate with respect to the parameter s.
Citation: Eleonora Cinti, Roberto Ognibene, Berardo Ruffini. A quantitative stability inequality for fractional capacities[J]. Mathematics in Engineering, 2022, 4(5): 1-28. doi: 10.3934/mine.2022044
[1] | 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 |
[2] | 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 |
[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] | Virginia Agostiniani, Lorenzo Mazzieri, Francesca Oronzio . A geometric capacitary inequality for sub-static manifolds with harmonic potentials. Mathematics in Engineering, 2022, 4(2): 1-40. doi: 10.3934/mine.2022013 |
[5] | Serena Dipierro, Giovanni Giacomin, Enrico Valdinoci . The fractional Malmheden theorem. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023024 |
[6] | 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 |
[7] | 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 |
[8] | Juan Pablo Borthagaray, Wenbo Li, Ricardo H. Nochetto . Finite element algorithms for nonlocal minimal graphs. Mathematics in Engineering, 2022, 4(2): 1-29. doi: 10.3934/mine.2022016 |
[9] | 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 |
[10] | Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional $ p $-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002 |
The aim of this work is to show a non-sharp quantitative stability version of the fractional isocapacitary inequality. In particular, we provide a lower bound for the isocapacitary deficit in terms of the Fraenkel asymmetry. In addition, we provide the asymptotic behaviour of the s-fractional capacity when s goes to 1 and the stability of our estimate with respect to the parameter s.
The classical isocapacitary inequality states that among sets which share the same amount of Lebesgue measure, balls minimize the electrostatic (Newtonian) capacity, that is, for any measurable set Ω with finite measure, the following scale invariant inequality holds true
|Ω|(2−n)/ncap(Ω)≥|B|(2−n)/ncap(B). | (1.1) |
Here |⋅| stands for the n−dimensional Lebesgue measure, n≥3, B is any ball in Rn and cap(⋅) is the standard electrostatic capacity in Rn, defined for compact sets as
cap(Ω):=inf{∫Rn|∇u|2dx:u∈C∞c(Rn),u≥1in Ω}. | (1.2) |
We observe that (1.1) can be rephrased in terms of the isocapacitary deficit, by saying that
dcap(Ω):=|Ω|(2−n)/ncap(Ω)|B|(2−n)/ncap(B)−1≥0. | (1.3) |
It is well known that the isocapacitary inequality is rigid, in the sense that dcap(Ω) vanishes if and only if Ω is equivalent to a ball up to a set of null Lebesgue measure. Thus, it appears as a natural quest the attempt of obtaining a quantitative stability version of (1.3). There are several possible geometric quantities that can properly measure the difference between a generic set and a ball with the same volume. The most natural one is the so-called Fraenkel asymmetry, first proposed by L. E. Fraenkel given by
A(Ω)=inf{|ΩΔB||Ω|:B is a ball with |B|=|Ω|}. |
The first attempts in this direction were made in the '90s. In particular, in [23] stability inequalities of the form
dcap(Ω)≥CnA(Ω)n+1 |
were proved, restricting to the class of convex sets when n≥3*. Nevertheless, in [23] the optimal exponent was conjectured to be 2, that is
dcap(Ω)≥C′nA(Ω)2, | (1.4) |
*In the planar case the suitable capacity is the logarithmic capacity.
which is asymptotically sharp for small asymmetries. Inequality (1.4) was proved in the planar case in [25,Corollary 2] (see also [2] and [24] for related results with other notions of deficiencies). As far as higher dimensions are concerned, (1.4) was proved by Fraenkel in [18] for starshaped sets, while in [21] the authors provided the inequality (1.4) with a suboptimal exponent but for general sets, i.e.,
dcap(Ω)≥C″nA(Ω)4. | (1.5) |
The conjecture in its full generality was finally established in [29]. It is worth stressing that to get this result the authors need to exploit the suboptimal inequality (1.5). We finally mention [28], where the author treated the case of the p-capacity and proved the corresponding sharp inequality. We point out that the approach followed in [28,29], while leading to the sharp exponent 2, does not allow to work out the explicit constant which multiplies the asymmetry. On the contrary, inequality (1.5) is not sharp for small values of A(Ω) but comes with an explicit constant C″n>0.
In this work we tackle the problem of quantification of the isocapacitary inequality in the fractional framework.
Let s∈(0,1) and let n>2s. We consider the fractional generalization of the capacity, defined for compact sets as follows
caps(Ω)=inf{[u]2s:u∈C∞c(Rn),u≥1on Ω}, | (1.6) |
where
[u]s:=(∫R2n|u(x)−u(y)|2|x−y|n+2sdxdy)12 |
denotes the fractional Gagliardo seminorm of order s. The definition of fractional capacity of a general closed set Ω⊆Rn is given in Definition 2.1, which can be easily proved to be equivalent to (1.6) when Ω is compact. As a straightforward consequence of the fractional analogue of the Pólya-Szegö inequality (proved in [1,Theorem 9.2], see also Proposition 2.10 below for an "extended" version) one can easily derive the fractional isocapacitary inequality, stating that
|Ω|(2s−n)/ncaps(Ω)≥|B|(2s−n)/ncaps(B), | (1.7) |
for any closed Ω⊆Rn with finite measure and for any closed ball B. The aim of this work is to quantify the fractional isocapacitary deficit
dcaps(Ω):=|Ω|(2s−n)/ncaps(Ω)|B|(2s−n)/ncaps(B)−1 |
in terms of the asymmetry of Ω. We point out that, in view of the scaling properties of the fractional capacity, the term
|B|(2s−n)/ncaps(B) |
is a universal constant, not depending on the choice of the ball B. It is also worth remarking that caps(⋅) can be defined, through (1.6), on open sets O and its value coincide with caps(¯O).
The fractional Pölya-Szegö inequality entails the rigidity of inequality (1.7), in the sense that the equality holds if and only if Ω is a ball in Rn, see [19,Theorem A.1]. We now present our main result, which amounts to a quantitative stability inequality for the fractional capacity.
Theorem 1.1. Let s∈(0,1) and n>2s. There exists a constant Cn,s>0, depending only on n and s, such that for any closed set Ω⊂Rn with finite measure, there holds
dcaps(Ω)=|Ω|(2s−n)/ncaps(Ω)|B|(2s−n)/ncaps(B)−1≥Cn,sA3s(Ω). | (1.8) |
Moreover the constant Cn,s can be explicitly computed, see Remark 3.5.
Our second result investigates the asymptotic behaviour of the function s↦caps(Ω) when s→1−, for a compact set Ω⊆Rn. In particular, we obtain that a suitable normalization of caps behaves like the standard capacity as s→1− (see (4.1) for the precise definition of the classical notion of capacity).
Proposition 1.2. Let n≥3, then for every Ω⊂Rn compact set, we have
lim sups↗1(1−s)caps(Ω)≤ωn2cap(Ω), | (1.9) |
where ωn:=|B1| and B1 denotes the unitary ball in Rn. If in addition Ω is the closure of an open bounded set with Lipschitz boundary then
lims↗1(1−s)caps(Ω)=ωn2cap(Ω). | (1.10) |
We observe that the exponent 3/s appearing in (1.8) is likely not sharp, †. Nevertheless, since the constant Cn,s in (1.8) can be explicitly computed (see Remark 3.5), together with its limit as s→1− (see Remark 4.1), our result entails an improvement of (1.5), asymptotically as s→1−. In particular, thanks also to Proposition 1.2 and Lemma 3.3, we are able to state the following.
†The optimal exponent was conjectured to be 2 in the fractional case as well.
Corollary 1.3. Let n≥3 and Ω⊆Rn be a closed set with finite measure. Then (1.8) is stable as s→1− and there holds
dcap(Ω)≥CnA(Ω)3, |
for some Cn>0 depending only on n, whose explicit value can be found in Remark 4.1.
Our proof is inspired by that in [3] where the authors deal with the (non-sharp) quantitative stability of the first eigenvalue of the fractional Laplacian with homogeneous Dirichlet exterior conditions. Such a result, in turn, relies on ideas established in [2,21]. Here, we provide a sketch of these arguments, starting with the classical case and then trying to emphasize the differences occurring in the fractional framework.
It is well known that, for any closed Ω⊆Rn with finite measure, there exists a unique function 0≤uΩ≤1, belonging to a suitable functional space, that achieves cap(Ω). Such a function is called the capacitary potential of Ω. First, by means of the coarea formula one gets
cap(Ω)∼∫Rn|∇uΩ|2dx∼∫10(∫{uΩ=t}|∇uΩ|dHn−1)dt. |
The right-hand side of the latter equality, after some manipulation can be written in terms of the perimeter of the superlevel sets {uΩ≥t}, i.e.,
cap(Ω)∼∫10P({uΩ≥t})2f1(t)dt. |
being f1 a suitably chosen real function depending only on the measure of {uΩ≥t}. The idea is then to exploit the sharp quantitative isoperimetric inequality [13,17,20]
dPer(E)∼|E|(n−1)/nP(E)−|B|(n−1)/nP(B)≳A(E)2, |
holding for any E⊆Rn in a suitable class and for any ball B⊆Rn. Plugging this into the previous estimate, after some further manipulation, one gets
dcap(Ω)≳∫10A({u≥t})2f2(t)dt, |
where f2 is an explicit positive integrable function depending only on the size of the superlevels of uΩ. Then we reason in the spirit of [2]:
heuristically, as long as t is close to ‖uΩ‖L∞(Rn)=1 we expect the set {uΩ≥t} is close to Ω in L1 and that A({uΩ≥t})∼A(Ω), and then, the idea is to seek for a threshold T such that at once
● A({u≥t})∼A(Ω) as long as t∈(T,1) and
● the quantity ∫1Tf2(t)dt results proportional to a power of A(Ω).
The previous two properties lead directly to the sought inequality. A bit more precisely, if the threshold T is such that 1−T≳A(Ω), then the above strategy works, while if 1−T≲A(Ω), then the fact that the asymmetry is large (with respect to 1−T) allows, by a simple comparison argument, to get an asymptotically stronger inequality of the form
dcap(Ω)≳A(Ω). |
In the fractional case the existence of a capacitary potential uΩ is guaranteed as well, see Remark 2.2. However, the arguments described above cannot be directly implemented in the fractional scenario, due to nonlocal effects. Indeed the very first step fails, since a suitable coarea formula for non-integer Sobolev spaces is missing. A way to overcome these difficulties is provided by the so called Caffarelli-Silvestre extension for functions in fractional Sobolev spaces. Loosely speaking, this tool allows us to interpret nonlocal energies of functions defined on Rn as local energies of functions depending on one more variable. Namely, one can prove a characterization of the s-capacity in the fashion of
caps(Ω)∼inf{∫Rn+1+z1−2s|∇U(x,z)|2dxdz:U(x,0)=uΩ(x)}, | (1.11) |
where
Rn+1+:={(x,z):x∈Rn,z>0} |
and U varies in a suitable functional space on Rn+1+. Moreover, one can prove that the infimum in (1.11) is uniquely achieved by a function 0≤UΩ≤1. We refer to Section 2.2 for the precise setting and definitions. At this point, the above strategy may be applied on every horizontal slice {(x,z):x∈Rn} and with UΩ(⋅,z) in place of uΩ. This way, we end up with
dcaps(Ω)≳∫∞0z1−2s∫10A({U(⋅,z)≥t})2fz(t)dtdz |
where, again, fz is an explicit real-valued function depending on the measure of the superlevels of U(⋅,z). Here it appears evident the extra inconvenience due to the presence of the integral in the z−variable. To get rid of this latter problem, we adapt ideas in [3] to show the existence of a good interval (0,z0), for which the (asymmetries of the) superlevels of UΩ(⋅,z) are close to (those of) the superlevels of uΩ, leading to
dcaps(Ω)≳∫z00z1−2s∫10A({UΩ(⋅,z)≥t})2fz(t)dtdz∼∫10A({uΩ≥t})2fz(t)dt |
and hence conclude similarly as in the classical local case.
In this section we introduce some prerequisites that are necessary in order to prove our main result.
First of all, we precisely define the functional setting we work in. For any open set O⊆Rn, we consider the homogeneous fractional Sobolev space Ds,2(O), defined as the completion of C∞c(O) with respect to the Gagliardo seminorm of order s
[u]s=(∫R2n|u(x)−u(y)|2|x−y|n+2sdxdy)12. |
The space Ds,2(O) is an Hilbert space, naturally endowed with the following scalar product
(u,v)Ds,2(Rn):=∫R2n(u(x)−u(y))(v(x)−v(y))|x−y|n+2sdxdy. |
Moreover, by trivial extension we have that Ds,2(O) is continuously embedded in Ds,2(Rn). We refer to [8] and [5] for more details concerning fractional homogeneous spaces and their characterizations. We limit ourselves to recall the fractional Sobolev inequality, which reads as follows:
Sn,s‖u‖2L2∗s(Rn)≤[u]2sfor all u∈Ds,2(Rn), | (2.1) |
where
2∗s:=2nn−2s |
denotes the critical Sobolev exponent in the fractional framework and Sn,s>0 denotes the best constant in the inequality. In particular, this result ensures the continuity of the embedding
Ds,2(Rn)↪L2∗s(Rn) |
and it provides the following characterization
Ds,2(Rn)={u∈L2∗s(Rn):[u]s<∞}. | (2.2) |
We refer to [14,Theorem 1.1] (see also [30,Theorem 7] in the Appendix) and to [5,Theorem 3.1] for the proofs of (2.1) and (2.2), respectively.
We now introduce the definition of fractional capacity of a closed subset of Rn.
Definition 2.1. Let Ω⊆Rn be closed and let ηΩ∈C∞c(Rn) be such that ηΩ=1 in an open neighbourhood of Ω. We define the fractional capacity of order s (or s-capacity) of the set Ω as follows:
caps(Ω):=inf{[u]2s:u∈Ds,2(Rn),u−ηΩ∈Ds,2(Rn∖Ω)}. |
First of all, we point out that the above definition does not depend on the choice of the cut-off function ηΩ. Indeed, if ˜ηΩ∈C∞c(Rn) satisfies ˜ηΩ=1 in a neighbourhood of Ω, then trivially
ηΩ+Ds,2(Rn∖Ω)=˜ηΩ+Ds,2(Rn∖Ω). |
We also observe that, if Ω⊆Rn is a compact set, then, by a simple regularization argument, one can easily prove that
caps(Ω)=inf{[u]2s:u∈C∞c(Rn),u≥1in Ω}. |
We also point out that it is not restrictive to assume that the admissible competitors u in the definition of caps(Ω) satisfy
0≤u≤1,a.e. in Rn, |
since
[u+∧1]s≤[u]sfor all u∈Ds,2(Rn), |
where u+ denotes the positive part of u and a∧b=min{a,b}. We refer to lemmas 2.6 and 2.7 in [32] for the proofs.
Remark 2.2. By direct methods of the calculus of variations, it is easy to check that caps(Ω) is uniquely achieved (when caps(Ω)<∞) by a function u∈Ds,2(Rn) such that u−ηΩ∈Ds,2(Rn∖Ω). Hereafter, we denote such function by uΩ and we call it the s-capacitary potential (or simply the capacitary potential) associated to Ω. Moreover, it is easy to observe that 0≤uΩ≤1 a.e. in Rn and that uΩ satisfies a variational equation, that is
(uΩ,φ)Ds,2(Rn)=0,for all φ∈Ds,2(Rn∖Ω). |
The notion of s-capacity of a set is in relation with the fractional Laplace operator of order s, which is defined, for u∈C∞c(Rn), as follows
(−Δ)su(x):=2P.V.∫Rnu(x)−u(y)|x−y|n+2sdy=2limr→0+∫{|x−y|>r}u(x)−u(y)|x−y|n+2sdy, |
where P.V. means that the integral has to be seen in the principal value sense. It is natural to extend the definition of fractional Laplacian applied to any function in Ds,2(Rn), in a distributional sense. More precisely, given u∈Ds,2(Rn), we have that (−Δ)su∈(Ds,2(Rn))∗ (with (Ds,2(Rn))∗ denoting the dual of Ds,2(Rn)) and it acts as follows
(Ds,2(Rn))∗⟨(−Δ)su,v⟩Ds,2(Rn)=(u,v)Ds,2(Rn),for all v∈Ds,2(Rn). |
Therefore, in view of Remark 2.2, we can say that the capacitary potential uΩ∈Ds,2(Rn) weakly satisfies
{(−Δ)suΩ=0,in Rn∖Ω,uΩ=1,in Ω. |
The proof of our main result strongly relies on an extension procedure for functions in fractional Sobolev spaces, first established in [11], which, in some sense, allows to avoid some nonlocal issues and recover a local framework. Such a procedure is commonly called Caffarelli-Silvestre extension. In this paragraph, we introduce the functional spaces emerging in the extended formulation and we discuss some of their properties, also in relation with the s-capacity of a set. We remark that, being the Caffarelli-Silvestre extension a classical tool nowadays, the results we present here can be regarded as folklore. But still, up to our knowledge, there are no explicit proofs available in the literature, hence we decided to report them here.
For any closed set K⊆∂Rn+1+≃Rn, we define the space D1,2(¯Rn+1+∖K;z1−2s) as the completion of C∞c(¯Rn+1+∖K) with respect to the norm
‖U‖D1,2(¯Rn+1+∖K;z1−2s):=(∫Rn+1+z1−2s|∇U|2dxdz)12. |
However hereafter we simply write D1,2z(¯Rn+1+∖K) in place of D1,2(¯Rn+1+∖K;z1−2s).
We have that D1,2z(¯Rn+1+∖K) is an Hilbert space with respect to the scalar product
(U,V)D1,2z(¯Rn+1+∖K):=∫Rn+1+z1−2s∇U⋅∇Vdxdz. |
First of all we shot that the space D1,2z(¯Rn+1+∖K) is a well defined functional space, which is not obvious. In order to prove that, it is sufficient to prove that it is the case for D1,2z(¯Rn+1+), in view of the continuous embedding
D1,2z(¯Rn+1+∖K)↪D1,2z(¯Rn+1+). |
To show this, we first recall by [15,Proposition 3.3] the following weighted Sobolev inequality
(∫Rn+1+z1−2s|U|2γdxdz)12γ≤S′n,s(∫Rn+1+z1−2s|∇U|2dxdz)12for all U∈D1,2z(¯Rn+1+), | (2.3) |
where S′n,s is a positive constant and γ:=1+2n−2s.
In particular this inequality yields the following continuous embedding
D1,2z(¯Rn+1+)↪L2γ(Rn+1+;z1−2s), |
where
L2γ(Rn+1+;z1−2s):={U∈L1loc(Rn+1+):∫Rn+1+z1−2s|U|2γdxdz<∞}. |
We now provide a characterization of D1,2z(¯Rn+1+) as a concrete functional space.
Proposition 2.3. The space D1,2z(¯Rn+1+) is a functional space. In particular there holds
D1,2z(¯Rn+1+)={U∈L2γ(Rn+1+;z1−2s):‖U‖D1,2z(¯Rn+1+)<+∞}. |
Proof. The fact that
D1,2z(¯Rn+1+)⊆{U∈L2γ(Rn+1+;z1−2s):‖U‖D1,2z(¯Rn+1+)<+∞} |
immediately follows from (2.3). We now prove the reverse inclusion. Namely, we show that any function
U∈L2γ(Rn+1+;z1−2s) | (2.4) |
such that
‖U‖D1,2z(¯Rn+1+)<+∞ | (2.5) |
can be approximated by functions in C∞c(¯Rn+1+) in the topology induced by the norm ‖⋅‖D1,2z(¯Rn+1+). First, suppose that U is compactly supported in ¯Rn+1+ and let
˜U(x,z):={U(x,z),if z>0,U(x,−z),if z<0. |
Moreover, we let {ρε}ε>0 be a family of mollifiers‡ in Rn+1 and we set
Uε=˜U⋆ρε|¯Rn+1+. |
‡We call mollifier a smooth, symmetric decreasing, positive and compactly supported function, which converges in distribution to a centered Dirac measure δ, as ε→0, and such that ‖ρε‖L1(Rn+1)=1.
Clearly Uε pointwisely converge to U in ¯Rn+1, as ε→0. Moreover it is equibounded in D1,2z(¯Rn+1+). Thus we easily conclude by means of the dominated convergence theorem. We consider now the general case. Fix ε>0 and let UR=UηR where ηR is the restriction to ¯Rn+1+ of a radial, smooth cut-off function defined on Rn+1 such that ηR=1 on BR, ηR=0 on Rn+1∖B2R and supRn+1|∇ηR|≤4R−1. Since, UR∈D1,2z(¯Rn+1+), by the previous step there exists VR∈C∞c(¯Rn+1+) such that ‖UR−VR‖D1,2z(¯Rn+1+)≤ε/2, so that, by triangular inequality
‖U−VR‖D1,2z(¯Rn+1+)≤‖U−UR‖D1,2z(¯Rn+1+)+ε2. |
We are left to show that UR→U in D1,2z(¯Rn+1+), as R→∞. In view of the properties of ηR, we have that
‖U−UR‖2D1,2z(¯Rn+1+)=∫Rn+1+z1−2s|∇U−∇(UηR)|2dxdz≤2∫Rn+1+z1−2s|∇U−ηR∇U|2dxdz+2∫Rn+1+z1−2s|U∇ηR|2dxdz≤2∫Rn+1+∖B+Rz1−2s|∇U|2dxdz+32R2∫B+2R∖B+Rz1−2s|U|2dxdz, |
where B+r:=Br∩Rn+1+. Thanks to (2.5), the first term on the right-hand side in the above inequalities is infinitesimal as R tends to infinity, so it can be chosen smaller than ε/4. For what concerns the second term, by Hölder inequality we obtain that
∫B+2R∖B+Rz1−2s|U|2dxdz≤(∫B+2R∖B+Rz1−2sdxdz)(γ−1)/γ(∫B+2R∖B+Rz1−2s|U|2γdxdz)1/γ. |
By (2.4) we have that
∫B+2R∖B+Rz1−2s|U|2γdxdz→0,as R→∞, |
while, by an explicit computation, also recalling that γ=1+nn−2s one gets that
supR≥11R2(∫B+2R∖B+Rz1−2sdxdz)(γ−1)/γ<+∞. |
Hence we can choose R large enough so that
16R2∫B+2R∖B+Rz1−2sU2dxdz≤ε/4, |
and conclude that
‖U−VR‖D1,2z(¯Rn+1+)≤ε. |
Another fundamental fact that relates the space D1,2z(¯Rn+1+) with Ds,2(Rn) is the existence of a trace map from the former to the latter. Before stating the precise result, we recall the following classical Hardy inequality, whose proof can be found e.g., in [26].
Lemma 2.4. Let p∈(1,∞) and a<1. Then there exists a constant C(p,a)>0 such that
∫∞0ρa|1ρ∫ρ0f(t)dt|pdρ≤C(p,a)∫∞0ρa|f(ρ)|pdρ, |
for all f∈C∞c([0,∞)).
Proposition 2.5. There exists a linear and continuous trace operator
Tr:D1,2z(¯Rn+1+)→Ds,2(Rn) |
such that Tr(U)(x)=U(x,0) for every U∈C∞c(¯Rn+1+).
Proof. Throughout the proof, we assume the space Ds,2(Rn) to be endowed with the following norm
[u]s,#:=(n∑i=1[u]2s,i)12. |
where
[u]2s,i:=∫∞0∫Rn|u(x+ρei)−u(x)|2ρ1+2sdxdρ, |
with ei denoting the unit vector in the positive xi variable. The norm [⋅]s,# is equivalent to [⋅]s, as proved in [3,Proposition B.1]. By density, it is sufficient to prove that there exists C>0 such that
[U(⋅,0)]s,#≤C‖U‖D1,2z(¯Rn+1+),for all U∈C∞c(¯Rn+1+). | (2.6) |
For U∈C∞c(¯Rn+1+), x∈Rn and ρ>0, we rewrite
U(x,0)=U(x,ρ)−∫ρ0∂U∂z(x,t)dt,U(x+ρei,0)=U(x+ρei,ρ)−∫ρ0∂U∂z(x+ρei,t)dt. |
Therefore
|U(x+ρei,0)−U(x,0)|2ρ1+2s≤2ρ1−2s|U(x+ρei,ρ)−U(x,ρ)|2ρ2+2ρ1−2s|1ρ∫ρ0(∂U∂z(x+ρei,t)−∂U∂z(x,t))dt|2. |
If we integrate in the x variable we obtain
∫Rn|U(x+ρei,0)−U(x,0)|2ρ1+2sdx≤2ρ1−2s(∫Rn|∂U∂xi(x,ρ)|2dx+2∫Rn|1ρ∫ρ0(∂U∂z(x,t)dt)|2dx), |
where we used the fact that
∫Rn|U(x+ρei,ρ)−U(x,ρ)|2ρ2dx≤∫Rn|∂U∂xi(x,ρ)|2dx |
for the first term and a change of variable for the second. By integration with respect to ρ in (0,∞) and thanks to Lemma 2.4 (choosing p=2 and a=1−2s) we infer
[U(⋅,0)]2s,i≤2∫Rn+1+ρ1−2s|∂U∂xi(x,ρ)|2dxdρ+C∫Rn+1+ρ1−2s|∂U∂z(x,ρ)|2dxdρ, |
for some constant C>0 depending only on s. If we now sum for i=1,…,n and take the square root, we obtain (2.6), thus concluding the proof.
The next step is to prove that the trace map introduced in the proposition above is onto. We first introduce the Poisson kernel of the upper half-space Rn+1+, defined as
Pz(x):=cn,sz2s(|x|2+z2)n+2s2,for (x,z)∈Rn+1+ |
where
cn,s:=(∫Rn1(1+|x|2)n+2s2dx)−1=π−n2Γ(n+2s2)Γ(s), | (2.7) |
is given in such a way that
∫RnPz(x)dx=1,for all z>0, |
see [22,Remark 2.2]. Essentially, a convolution with this kernel allows to extend to the upper half-space Rn+1+ functions that are defined on Rn. This is done first for functions in C∞c(Rn) and then extended by density to the whole Ds,2(Rn). Namely, we have the following.
Proposition 2.6. Let φ∈Ds,2(Rn). Then the function
Uφ(x,z):=(Pz⋆φ)(x) | (2.8) |
belongs to D1,2z(¯Rn+1+) and
‖Uφ‖2D1,2z(¯Rn+1+)=αn,s[φ]2s, | (2.9) |
where
αn,s:=s(1−s)πn2Γ(1−s)Γ(n+2s2)Γ(s)Γ(2−s)>0. | (2.10) |
In particular the trace operator established in Proposition 2.5 is surjective.
Proof. For any φ∈C∞c(Rn), we let
(Eφ)(x,z):=(Pz⋆φ)(x). |
It is easy to check that
‖Eφ‖2D1,2z(¯Rn+1+)=∫Rn+1+z1−2s|∇(Eφ)|2dxdz=αn,s[φ]2s. | (2.11) |
For the computation of the explicit constant see e.g., [10,Remark 3.11]. Moreover, by the weighted Sobolev inequality (2.3)
∫Rn+1+z1−2s|Eφ|2γdxdz<∞. |
Hence, by Proposition 2.3, we have that Eφ∈D1,2z(¯Rn+1+). Therefore the map
E:C∞c(Rn)→D1,2z(¯Rn+1+) |
is linear and continuous, thus it can be uniquely extended in the whole Ds,2(Rn) and (2.11) still holds. We now prove that Eφ=Uφ for any φ∈Ds,2(Rn). Let φ∈Ds,2(Rn) and (φi)i⊆C∞c(Rn) be such that φi→φ in Ds,2(Rn) as i→∞. Thanks to (2.1), we have that
φi→φin L2∗s(Rn),as i→∞, |
which, by definition, implies that
Eφi=Uφi→Uφpointwise in Rn+1+,as i→∞. |
On the other hand Eφi→Eφ in D1,2z(¯Rn+1+) as i→∞; hence, up to a subsequence
Eφi→Eφa.e. in Rn+1+,as i→∞. |
Therefore Eφ=Uφ and the proof is complete.
The following corollary, in view of the previous results, tells us that there is an isometry between the space Ds,2(Rn) and the subspace of D1,2z(¯Rn+1+) containing the (unique) minimizers of a certain functional.
Corollary 2.7. Let φ∈Ds,2(Rn). Then the minimization problem
minU∈D1,2z(¯Rn+1+){∫Rn+1+z1−2s|∇U|2dxdz:TrU=φ} |
admits a unique solution, which coincides with Uφ as in (2.8). Moreover
{−div(z1−2s∇Uφ)=0,inRn+1+,Uφ=φ,onRn,−limz→0+z1−2s∂Uφ∂z=αn,s(−Δ)sφ,onRn, |
in a weak sense, that is
∫Rn+1+z1−2s∇Uφ⋅∇Vdxdz=αn,s(φ,TrV)Ds,2(Rn),for allV∈D1,2z(¯Rn+1+), |
where αn,s>0 is as in (2.10).
Proof. The proof is standard. For instance see [3,Proposition 2.6] for the first part and [11] for the second.
Remark 2.8. In view of the extension procedure described above, we can relate the notion of s-capacity with another notion of (weighted) capacity of sets in Rn+1. More precisely, with a slight abuse of notation, we call D1,2z(Rn+1) the completion of C∞c(Rn+1) with respect to the norm
‖U‖D1,2z(Rn+1):=(∫Rn+1|z|1−2s|∇U|2dxdz)12 |
and, for any closed K⊆Rn+1, we let
capRn+1(K;|z|1−2s):=inf{‖U‖2D1,2z(Rn+1):U∈D1,2z(Rn+1),U−ζK∈D1,2(Rn+1∖K)}, |
where ζK∈C∞c(Rn+1) is equal to 1 in a neighborhood of K. Thanks to (2.9), after an even reflection in the z variable, one can see that
12capRn+1(Ω;|z|1−2s)=αn,scaps(Ω). |
for any closed Ω⊆Rn. Moreover the unique function achieving capRn+1(Ω;|z|1−2s) coincides with the even-in-z reflection of Pz⋆uΩ, with uΩ∈Ds,2(Rn) denoting the capacitary potential of caps(Ω), as in Remark 2.2.
Hereafter we denote by
UΩ(x,z):=(Pz⋆uΩ)(x),for (x,z)∈Rn+1+ |
the restriction to the upper half-space of the potential associated to cap(Ω;|z|1−2s). Hence, UΩ∈D1,2z(¯Rn+1+) satisfies 0≤UΩ≤1 a.e. in Rn+1+ and weakly solves
{−div(z1−2s∇UΩ)=0,in Rn+1+,UΩ=1,on Ω,−limz→0+z1−2s∂UΩ∂z=0,on Rn∖Ω, |
in the sense that UΩ−ζΩ∈D1,2z(¯Rn+1+∖Ω) and
∫Rn+1+z1−2s∇UΩ⋅∇Vdxdz=0,for all V∈D1,2z(¯Rn+1+∖Ω). |
Thanks to [31,Theorem 1.1], we can observe that UΩ∈C∞(¯Rn+1+∖Ω). Hence, in particular uΩ∈C∞(Rn∖Ω).
Remark 2.9. We emphasize a technical difference with respect to [3]. Namely that the functional setting we adopt here is tailored for the problem under investigation. Indeed, being the s-capacity obtained by minimization of the (nonlocal) energy, it is natural to expect the minimizer to have only finite seminorm [⋅]s, together with the possibility of approximating it by means of smooth and compactly supported functions, at least for compact Ω. Therefore, any other integrability assumption on the capacitary potential appears as artificial. Observe that, differently from our Proposition 2.6 and Corollary 2.7, in which the trace function φ needs just to belong to Ds,2(Rn), in the analogue extension result Proposition 2.6 of [3] an additional integrability assumption on φ is required.
This paragraph is devoted to the isocapacitary inequality (1.7). The classical and simplest proof of the (standard) isocapacitary inequality,
|Ω|(2−n)/ncap(Ω)≥|B|(2−n)/ncap(B) | (2.12) |
is by rearrangement: given any function u:Rn→R, its symmetric decreasing rearrangement is the radial decreasing function u∗:Rn→R such that |{u>t}|=|{u∗>t}|. As a consequence of the Pólya-Szegö inequality
∫Rn|∇u|2dx≥∫Rn|∇u∗|2dx |
applied to the capacitary potential of a closed Ω⊆Rn one can easily see that cap(Ω)≥cap(B) as long as |B|=|Ω|<∞, from which, by scaling (2.12) holds as well. Indeed, the symmetric rearrangement of the capacitary potential of Ω coincides with the potential of a ball with the same volume as Ω. Following this path, one can prove the fractional isocapacitary inequality using symmetric rearrangements for the extended problem in Rn+1+.
As in [3,22], we define in Rn+1+ the partial Schwartz symmetrization U∗ of a nonnegative function U∈D1,2z(Rn+1+). By construction, the function U∗ is obtained by taking for almost every z>0, the n−dimensional Schwartz symmetrization of the map
x↦U(x,z). |
More precisely: for almost every fixed z>0, the function U∗(⋅,z) is defined to be the unique radially symmetric decreasing function on Rn such that for all t>0
|{U∗(⋅,z)>t}|=|{U(⋅,z)>t}|. |
Proposition 2.10. Let φ∈Ds,2(Rn) be a nonnegative function and let Uφ∈D1,2z(¯Rn+1+) as in (2.8). Then
U∗φ∈D1,2z(¯Rn+1+) |
and the following Pólya-Szegö type inequalities hold true
∫Rn+1+z1−2s|∇U∗φ|2dxdz≤∫Rn+1+z1−2s|∇Uφ|2dxdz,∫Rn+1+z1−2s|∂zU∗φ|2dxdz≤∫Rn+1+z1−2s|∂zUφ|2dxdz. | (2.13) |
Moreover, we have Tr(U∗φ)=φ∗. In particular, we get
∫Rn+1+z1−2s|∇U∗φ|2dxdz≥αn,s[φ∗]2Ds,2(Rn). |
Proof. By density, it is enough to show the result for φ∈C∞c(Rn). In that case, the proof follows directly by [3,Proposition 3.2].
Now, a proof of the fractional isocapacitary inequality (1.7) can be obtained as a direct consequence of the previous result, applied to φ=uΩ, and Remark 2.8.
In this section we give the proof of our main result. The idea consists in introducing quantitative elements in the proof of the isocapacitary inequality established in the previous section. As already explained in the Introduction, the major inconvenience when working with the extended problem in Rn+1+ consists in the fact that we need to transfer information on the superlevel sets of the extension of the capacitary potential {UΩ(⋅,z)≥t} for fixed z>0, to information on the superlevel sets of its trace uΩ in Rn. This was done in Section 4 of [3] for a problem concerning the stability of the first eigenvalue of the Dirichlet fractional Laplacian.
We start by recalling the following technical result, whose proof can be found in [3,Lemma 4.1].
Lemma 3.1. Let Ω,E be two measurable subsets of Rn of finite measure and such that
|ΩΔE||Ω|≤δ3A(Ω), |
for some δ∈(0,1). Then
A(E)≥CδA(Ω), |
where
Cδ:=3−2δ3+2δ. | (3.1) |
The following lemma corresponds to Proposition 2.6 of [3]. Observe that, here, the extension Uφ(⋅,z) of a function φ∈Ds,2(Rn) does not belong to L2(Rn) for fixed z≥0; nevertheless, with the same computations as in [3], we can show that it is close enough (depending on z) in L2 to its trace u. We report a proof of the lemma for the sake of completeness.
Lemma 3.2. For any φ∈Ds,2(Rn), denoting by Uφ∈D1,2z(¯Rn+1+) its extension, there holds
‖Uφ(⋅,z)−φ‖L2(Rn)≤√cn,s[φ]szs,forz>0, |
where cn,s is given in (2.7).
Proof. We first prove the following preliminary fact
∫RnPz(y)‖τyφ−φ‖L2(Rn)dy≤√cn,szs[φ]s<∞for all φ∈Ds,2(Rn)and all z≥0, | (3.2) |
where τyφ(⋅)=φ(⋅−y). Indeed, multiplying and dividing by |y|(n+2s)/2 and applying Cauchy-Schwartz inequality yields
∫RnPz(y)‖τyφ−φ‖L2(Rn)dy≤(∫RnPz(y)2|y|n+2sdy)12(∫Rn‖τyφ−φ|y|s‖2L2(Rn)dy|y|n)12=(∫RnPz(y)2|y|n+2sdy)12[φ]s, |
where, in the last step, we used the fact that
∫Rn‖τyφ−φ|y|s‖2L2(Rn)dy|y|n=[φ]2s. |
The proof of (3.2) ends by observing that
(∫RnPz(y)2|y|n+2sdy)12=√cn,szs. |
Let us now consider φ∈Ds,2(Rn). By definition of Uφ and Minkowski's inequality we have that
‖(Uφ(⋅,z)−φ)χBR‖L2(Rn)≤∫RnPz(y)‖τyφ−φ‖L2(BR)dy, |
for any R>0, where χBR denotes the characteristic function of the ball BR. We conclude the proof applying Fatou's Lemma for R→∞ and combining the resulting inequality with (3.2).
The following result allows us to focus on compact sets without loss of generality. Therefore hereafter in this section we always assume Ω⊆Rn to be a compact set.
Lemma 3.3 (Reduction to compact sets). Let Ω⊆Rn be a closed set with finite measure. Then
limr→∞caps(Ω∩¯Br)=caps(Ω)andlimr→∞A(Ω∩¯Br)=A(Ω), |
where Br:={x∈Rn:|x|<r}.
Proof. The convergence of the capacity follows from the fact that, for any sequence of closed sets Ωi⊆Rn such that Ωi⊆Ωi+1, there holds
caps(∪∞i=1Ωi)=limi→∞caps(Ωi) |
which, in turn, can be proved by following step by step the proof of [16,Theorem 4.15,Point (viii)]. The rest of the proof is easy and can be omitted.
Hereafter in this section, for 0≤t≤1 and z≥0, we let
Ωt,z:={x∈Rn:UΩ(x,z)≥t}andΩt:=Ωt,0={x∈Rn:uΩ(x)≥t}. |
We notice that, by the continuity of uΩ, it follows that |ΩtΔΩ|→0, as t→1−. Moreover, we set, respectively
μz(t):=|Ωt,z|andμ(t):=μ0(t)=|Ωt|. |
We immediately observe that μ is left-continuous and non-increasing in (0,1) and that
limt→0+μ(t)=+∞andlimt→1−μ(t)=|Ω|. |
We now let
T=T(Ω,γ)=inf{0≤t≤1:|{uΩ≥t}|≤|Ω|(1+γA(Ω))},=inf{0≤t≤1:μ(t)≤|Ω|(1+γA(Ω))}. | (3.3) |
The constant γ is chosen in (0,1/9) and will be settled later on. Notice that if A(Ω)>0 then T<1. In addition, in view of the left-continuity of μ, we know that
μ(T)≥|Ω|(1+γA(Ω)). | (3.4) |
The following proposition, which will be crucial in the proof of our main result, allows to bound from below the asymmetry of the superlevel sets of UΩ(⋅,z) with the asymmetry of Ω (for certain levels t and for z small enough).
Proposition 3.4. Let Ω be such that A(Ω)>0, γ∈(0,1/9), and let T∈(0,1) as in (3.3). Set also ˆT=1−T. Then, letting cn,s be as in (2.7), if
T+18ˆT≤t≤T+38ˆTand0<z≤z0:=(ˆT16√γA(Ω)|Ω|cn,scaps(Ω))1s, |
we have
||Ωt,z|−|Ω|||Ω|≤3γA(Ω), | (3.5) |
and
A(Ωt,z)≥cγA(Ω), | (3.6) |
where cγ=C3γ and C3γ is as in (3.1).
Proof. Let τ=T+ˆT2. First of all, we observe that, by triangle inequality
||Ωt,z|−|Ω|||Ω|≤|Ωt,zΔΩ||Ω|=|Ωt,z∖Ω||Ω|+|Ω∖Ωt,z||Ω|≤|Ωt,z∖ΩT+ˆT16||Ω|+|ΩT+ˆT16∖Ω||Ω|+|Ω∖Ωt,z||Ω|≤|Ωt,z∖ΩT+ˆT16||Ω|+γA(Ω)+|Ωτ∖Ωt,z||Ω|. | (3.7) |
where, in the last inequality we used the definition of T and the facts that T+ˆT16>T and Ω⊆Ωτ. For x∈Ωτ∖Ωt,z we have that
uΩ(x)−UΩ(x,z)≥τ−t≥T+ˆT2−(T+38ˆT)=18ˆT, |
which shows that Ωτ∖Ωt,z⊂{|uΩ−UΩ(⋅,z)|≥18ˆT}. Hence, using Chebichev's inequality and Lemma 3.2 with φ=uΩ, it holds that
|Ωτ∖Ωt,z||Ω|≤|{|uΩ−UΩ(⋅,z)|≥18ˆT}||Ω|≤64|Ω|ˆT2‖uΩ−UΩ(⋅,z)‖2L2(Rn)≤64cn,s[uΩ]2|Ω|ˆT2z2s≤γA(Ω), |
as long as z≤(ˆT8√γA(Ω)|Ω|cn,scaps(Ω))1s. Similarly, one can check that
|Ωt,z∖ΩT+ˆT16||Ω|≤γA(Ω). |
as long as z≤(ˆT16√γA(Ω)|Ω|cn,scaps(Ω))1s. This, together with (3.7) entails that
||Ωt,z|−|Ω|||Ω|≤3γA(Ω), |
which proves (3.5). Finally, applying Lemma 3.1 (with δ=3γ and γ chosen to be in (0,1/6)), we deduce (3.6).
We can now give the proof of our main result.
Proof of Theorem 1.1. By scaling invariance, we may assume |Ω|=1. We also fix γ=10−1∈(0,1/9) throughout the proof. We have to prove that
caps(Ω)−caps(B)≥Ccaps(B)A(Ω)3s, | (3.8) |
for some C>0, where B is a ball with unit volume.
We start by observing that if caps(Ω)>2caps(B), then, since A(Ω)≤2, we easily deduce that
caps(Ω)−caps(B)>caps(B)=caps(B)23s23s≥caps(B)23sA(Ω)3s, |
which proves (3.8) with C=2−3s.
Thus, we can just consider the case in which
caps(Ω)≤2caps(B). |
We recall that in (3.3), we have defined the level T as
T=inf{0≤t≤1:μ(t)≤|Ω|(1+γA(Ω))}. |
Notice that by Lemma 3.1 as long as 1≥t≥T it holds
A(Ωt)≥cγA(Ω), |
for some cγ independent of t and Ω.
We now distinguish between two cases, in terms of the relation of T with A(Ω). More precisely, we let
λ:=n−2snγandκ:=λ4(1+2λ). | (3.9) |
We consider the ranges
1−T≥κA(Ω)and1−T<κA(Ω). |
Case1−T≥κA(Ω)_. In this case we argue as in the proof of Proposition 4.4 and of Theorem 1.3 (case T>T0) in [3]. The idea consists in introducing quantitative elements in the proof of the Pólya-Szegö inequality by applying the quantitative isoperimetric inequality on each (horizontal) level set Ωt,z of the function UΩ(⋅,z).
First, we recall that
caps(Ω)=[uΩ]2Ds,2(Rn)=α−1n,s∫Rn+1+z1−2s|∇xUΩ|2dxdz+α−1n,s∫Rn+1+z1−2s|∂zUΩ|2dxdz. |
For what concerns the z-derivative, from (2.13) we know that
∫Rn+1+z1−2s|∂zUΩ|2dxdz≥∫Rn+1+z1−2s|∂zU∗Ω|2dxdz. |
For the x-derivative, we argue as in the local case. By the coarea formula, we have
∫Rn+1+z1−2s|∇xUΩ|2dxdz=∫+∞0z1−2s(∫+∞0(∫{x∈Rn:UΩ(x,z)=t}|∇xUΩ|2dHn−1(x)|∇xUΩ|)dt)dz≥∫+∞0z1−2s(∫+∞0P(Ωt,z)2∫{x∈Rn:UΩ(x,z)=t}dHn−1(x)|∇xUΩ|dt)dz | (3.10) |
where P(Ωt,z) denotes the perimeter of the set Ωt,z, and in the last step we have used Jensen's inequality. Using the quantitative isoperimetric inequality, one can prove that
P(Ωt,z)2≥P(Ω∗t,z)2+cnμz(t)2(n−1)nA(Ωt,z)2, | (3.11) |
where
Ω∗t,z:={x∈Rn:U∗Ω(x,z)≥t}, |
and
cn:=2ω2nn(2−2n−1n)3(181)2n12. |
The proof of (3.11) can be easily carried out by following [4,Lemma 2.9], see also the proof of Proposition 4.4 in [3]. By definition of symmetric rearrangement, we have that
μz(t)=|Ω∗t,z|for a.e. t∈(0,1) |
and, from Lemma 3.2 and inequality (3.19) in [12] (see also (2.6) in [20]), we know that
−μ′z(t)=∫{x∈Rn:U∗Ω(x,z)=t}dHn−1(x)|∇xU∗Ω|≥∫{x∈Rn:UΩ(x,z)=t}dHn−1(x)|∇xUΩ|. |
Therefore, combining (3.10) and (3.11) with this last inequality, we can estimate the L2-norm of the x-gradient as follows
∫Rn+1+z1−2s|∇xUΩ|2dxdz≥∫+∞0z1−2s(∫+∞0P(Ω∗t,z)2−μ′z(t)dt)dz+cn∫+∞0z1−2s(∫+∞0(μz(t)n−1n)2A(Ωt,z)2−μ′z(t)dt)dz=∫Rn+1+z1−2s|∇xU∗Ω|2dxdz+cn∫+∞0z1−2s(∫+∞0(μz(t)n−1n)2A(Ωt,z)2−μ′z(t)dt)dz. |
Moreover, we have seen in the proof of Theorem 2.10 that
α−1n,s∫Rn+1+z1−2s|∇U∗Ω|2dxdz≥caps(B). |
Collecting all together, we obtain
caps(Ω)=α−1n,s∫Rn+1+z1−2s|∇xUΩ|2dxdz+α−1n,s∫Rn+1+z1−2s|∂zUΩ|2dxdz≥caps(B)+α−1n,scn∫+∞0z1−2s(∫+∞0(μz(t)n−1n)2A(Ωt,z)2−μ′z(t)dt)dz. |
We use now Proposition 3.4, to pass from A(Ωt,z) to A(Ω). Let z0=(ˆT16√γA(Ω)|Ω|cn,scaps(Ω))1s be as in Proposition 3.4. Set also ˆT=1−T. We have
caps(Ω)−caps(B)≥α−1n,scn∫+∞0z1−2s(∫+∞0(μz(t)n−1n)2A(Ωt,z)2−μ′z(t)dt)dz≥α−1n,scn∫z00z1−2s(∫T+38ˆTT+ˆT8A(Ωt,z)2(μz(t)n−1n)2−μ′z(t)dt)dz≥α−1n,scnc2γA(Ω)2∫z00z1−2s(∫T+38ˆTT+ˆT8(μz(t)n−1n)2−μ′z(t)dt)dz, |
where, in the last inequality, we used (3.6). It remains to estimate the term
∫z00z1−2s∫T+38ˆTT+ˆT8(μz(t)n−1n)2−μ′z(t)dtdz. | (3.12) |
First, we observe that, since γ<1/9, using (3.5) and the fact that A(Ω)≤2, we have
μz(t)≥1−3γA(Ω)≥13. |
Hence, in order to estimate (3.12), it is enough to control from below the quantity
∫z00z1−2s∫T+38ˆTT+ˆT81−μ′z(t)dtdz. |
By Jensen inequality and recalling the definition of μz, we have
∫T+38ˆTT+ˆT81−μ′z(t)dt≥ˆT2161∫T+38ˆTT+ˆT8−μ′z(t)dt≥ˆT2161|ΩT+ˆT8,z|−|ΩT+38ˆT,z|. |
Using again (3.5), we deduce that
|ΩT+ˆT8,z|−|ΩT+38ˆT,z|≤1+3γA(Ω)−(1−3γA(Ω))=6γA(Ω). |
We can now estimate (3.12) as follows
∫z00z1−2s∫T+38ˆTT+ˆT8(μz(t)n−1n)2−μ′z(t)dtdz≥C1ˆT2A(Ω)∫z00z1−2sdz, |
where C1>0 depends only on n and γ. This, in turn, yields the following estimate for the capacity variation
caps(Ω)−caps(B)≥C2αn,sA(Ω)ˆT2∫z00z1−2sdz, |
with C2>0 depending only on n and γ. Finally, recalling the definition of z0 (in Proposition 3.4) and that we are in the ranges caps(Ω)≤2caps(B) and ˆT=1−T≥κA(Ω), we obtain
\begin{align*} {{{\rm{cap}}}_s}(\Omega)-{{{\rm{cap}}}_s}(B)&\geq \frac{C_3}{\alpha_{n, s}}\left(\frac{C_4}{c_{n, s}} \right)^{\frac{1}{s}-1}\frac{1}{(1-s){{{\rm{cap}}}_s}(B)^{\frac{1}{s}}}{{{\rm{cap}}}_s}(B)\mathcal{A}(\Omega)^{\frac{1}{s}}\widehat{T}^{\frac{2}{s}} \\ &\geq \frac{\kappa^{\frac{2}{s}}C_3}{\alpha_{n, s}}\left(\frac{C_4}{c_{n, s}}\right)^{\frac{1}{s}-1}\frac{1}{(1-s){{{\rm{cap}}}_s}(B)^{\frac{1}{s}}}{{{\rm{cap}}}_s}(B)\mathcal{A}(\Omega)^{\frac{3}{s}}, \end{align*} |
with C_3, C_4 > 0 depending on n and \gamma . In particular, given \gamma = 10^{-1} it is possible to see that
\begin{equation} C_3 = \frac{5}{3^5(181)^2}\frac{(3\omega_n)^{\frac{2}{n}}}{n^{12}}(1-2^{-\frac{1}{n}})^3. \end{equation} | (3.13) |
Therefore (3.8) holds with
\begin{equation*} C = \frac{\kappa^{\frac{2}{s}}C_3}{\alpha_{n, s}}\left(\frac{C_4}{c_{n, s}}\right)^{\frac{1}{s}-1}\frac{1}{(1-s){{{\rm{cap}}}_s}(B)^{\frac{1}{s}}}. \end{equation*} |
\underline {{\rm{Case}}\; 1 - T < \kappa {\cal A}(\Omega)} In this case we get the quantitative inequality by means of a suitable test function for the definition of {{{\rm{cap}}}_s} . Let us define
w_T = \min\left\{1, \frac{u_\Omega}{T}\right\}. |
It is possible to see that w_T is an admissible competitor for {{{\rm{cap}}}_s}(\Omega_T) . Indeed, let \xi_k: = \tilde{\xi}_k\star \rho_{\frac{1}{2k}} , where \rho_\epsilon denotes a standard mollifier in \mathbb{R} and \tilde{\xi}_k\colon \mathbb{R}\to \mathbb{R} is defined as follows
\tilde{\xi}_k(\sigma): = \begin{cases} 0, &\text{if }\sigma\leq \frac{1}{k}, \\ \frac{k}{k-2}\left(\sigma-\frac{1}{k}\right), &\text{if }\frac{1}{k}\leq \sigma\leq 1-\frac{1}{k}, \\ 1, &\text{if }\sigma\geq 1-\frac{1}{k}. \end{cases} |
It is clear that \xi_k\in C^\infty(\mathbb{R}) and that
\xi_k(\sigma)\to \min\{ \sigma^+, 1 \}, \quad\text{uniformly as }k\to\infty. |
Moreover, if we let w_k: = \xi_k\circ w_T , one can see that w_k\in\mathcal{D}^{s, 2}(\mathbb{R}^n) and, thanks to the continuity of w_T , that w_k = 1 in an open \Omega_k\supseteq \Omega_T , thus implying that w_k-\eta_{\Omega_T}\in\mathcal{D}^{s, 2}(\mathbb{R}^n\setminus\Omega_T) , being \eta_{\Omega_T}\in C_c^\infty(\mathbb{R}^n) such that \eta_{\Omega_T} = 1 in a neighbourhood of \Omega_T . Finally, it is easy to check that w_k\to w_T in \mathcal{D}^{s, 2}(\mathbb{R}^n) as k\to \infty , which means that w_T-\eta_{\Omega_T}\in\mathcal{D}^{s, 2}(\mathbb{R}^n\setminus\Omega_T) . Therefore, we have
\begin{equation*} {{{\rm{cap}}}_s}(\Omega_T)\le [w_T]^2_s\le \frac{1}{T^2} [u_{\Omega}]^2_s = \frac{1}{T^2}{{{\rm{cap}}}_s}(\Omega). \end{equation*} |
From the previous inequality, the isocapacitary inequality (1.7) and (3.4) we obtain that
\begin{equation} {{{\rm{cap}}}_s}(\Omega)\ge T^2{{{\rm{cap}}}_s}(B)|\Omega_T|^{\frac{n-2s}{n}}\geq T^2{{{\rm{cap}}}_s}(B)\left(1+\gamma\mathcal A(\Omega)\right)^{\frac{n-2s}{n}}. \end{equation} | (3.14) |
By convexity, we know that
\begin{equation} T^2\geq 1-2(1-T)\geq 1-2\kappa\mathcal{A}(\Omega) \end{equation} | (3.15) |
and that
\begin{equation} (1+\gamma{\mathcal A}(\Omega))^{\frac{n-2s}{n}}\geq 1+\lambda {\mathcal A}(\Omega), \end{equation} | (3.16) |
with \lambda as in (3.9). By the definition of \lambda and \kappa , as in (3.9), we derive that
\begin{equation} (1-2\kappa {\mathcal A}(\Omega))(1+\lambda{\mathcal A}(\Omega))\geq 1+\frac{\lambda}{2}{\mathcal A}(\Omega)\geq (1+C_5{\mathcal A}(\Omega)^{\frac{3}{s}}), \end{equation} | (3.17) |
with
C_5: = 2^{-\frac{3}{s}}\lambda |
Putting together (3.14) with (3.15), (3.16) and (3.17), we obtain (3.8) with C = C_5 , thus concluding the proof.
Remark 3.5. By carefully scanning the proof of Theorem 1.1, one can explicitly find the constant C_{n, s} appearing in (1.8), which amounts to
\begin{equation*} C_{n, s} = \max\left\{ 2^{-\frac{3}{s}}, \frac{\kappa^{\frac{2}{s}}C_3}{\alpha_{n, s}}\left(\frac{C_4}{c_{n, s}}\right)^{\frac{1}{s}-1}\frac{1}{(1-s){{{\rm{cap}}}_s}(B)^{\frac{1}{s}}} \right\}, \end{equation*} |
where \kappa is as in (3.9), and C_3, C_4 depend only on n (indeed their dependence on \gamma stated in the proof of Theorem 1.1 is actually pointless, being \gamma universally fixed in (0, 1/9) , see (3.13) for the explicit value of C_3 ). The constants \alpha_{n, s} and c_{n, s} are as in (2.10) and (2.7), respectively, and they are uniformly bounded away from 0 and +\infty as s\to 1^- , see [3,Remark 2.7]. Moreover, it can be easily checked that this fact holds true for the constant \kappa as well, by its definition, when n\geq 3 .
In this section we prove Thereom 1.2. We recall the definition of the standard (Newtonian) capacity of a closed \Omega \subseteq \mathbb R^n for n\geq3 , which is equivalent to (1.2) when \Omega is a compact set
\begin{equation} {{{\rm{cap}}}}(\Omega) = \inf\left\{\int_{ \mathbb R^n} \left| {\nabla u} \right|^2\, dx\colon u\in \mathcal{D}^{1, 2}( \mathbb R^n)\; \text{and }u-\eta_\Omega\in \mathcal{D}^{1, 2}( \mathbb R^n\setminus \Omega) \right\}, \end{equation} | (4.1) |
where \eta_\Omega\in C_c^\infty(\mathbb R^n) is such that \eta_\Omega = 1 in an open neighbourhood of \Omega and, for any open \mathcal{O} \subseteq \mathbb R^n , the space \mathcal{D}^{1, 2}(\mathcal{O}) is defined as the completion of C_c^\infty(\mathcal{O}) with respect to the norm
u\mapsto \left(\int_{\mathcal{O}}\left| {\nabla u} \right|^2\, dx \right)^{1/2}. |
We also recall that, when \mathcal{O} is bounded, then the space \mathcal{D}^{1, 2}(\mathcal{O}) coincides with the usual Sobolev space W_0^{1, 2}(\mathcal{O}) , thanks to the validity of the Poincaré inequality.
Proof of Proposition 1.2. Proof of (1.9). The first part of the statement follows easily by a celebrated result by Bourgain-Brezis-Mironescu stating that
\lim\limits_{s\nearrow 1}(1-s)\, [\varphi]^2_{s} = \frac{\omega_n}{2}\, \int_{ \mathbb R^n} |\nabla \varphi|^2\, dx, \qquad \mbox{ for every } \varphi\in C^\infty_c( \mathbb R^n). |
Indeed, by taking a function \varphi\in C_c^\infty(\mathbb R^n) satisfying \varphi\ge \chi_\Omega , we deduce that
\limsup\limits_{s\nearrow 1}(1-s)\, {\rm cap}_s(\Omega)\le \lim\limits_{s\nearrow 1}(1-s)\, [\varphi]^2_{s} = \frac{\omega_n}{2}\, \int_{ \mathbb R^n} |\nabla \varphi|^2\, dx. |
Finally, (1.9) follows by taking the infimum over all admissible \varphi .
Proof of (1.10). Let us fix s_0\in(0, 1) and let us denote by u_{s, \Omega} the s -capacitary potential of the set \Omega . Since \Omega is compact, there exists a ball B_{R_0} which contains \Omega . Hence, we have that
u_{s, \Omega} \le u_{s, B_{R_0}}\;\; \mbox{a.e. in } \mathbb R^n. |
This can be easily proved by taking the Kelvin transform of the above functions and applying the maximum principle as in [9,Theorem 3.3.2]. We can now take advantage of the following precise decay rate of u_{s, B_{R_0}} , established in [6,Proposition 3.6]:
\begin{equation} u_{s, \Omega}(x) \le u_{s, B_{R_0}}(x)\le \frac{2{R_0^{n-2s}}}{|x|^{n-2s}}, \quad \mbox{for } |x| > R_0. \end{equation} | (4.2) |
Let us define an almost optimal function, given by a suitable truncation of u_{s, \Omega} . For any fixed \varepsilon > 0 , we set
u_{s, \Omega}^{\varepsilon}: = \frac{(u_{s, \Omega}-\varepsilon)^+}{1-\varepsilon}. |
We claim that u_{s, \Omega}^{\varepsilon}\in \mathcal{D}^{s, 2}(\mathbb R^n) and u_{s, \Omega}^{\varepsilon}-\eta_\Omega\in \mathcal{D}^{s, 2}(\mathbb R^n\setminus\Omega) , with \eta_\Omega\in C_c^\infty(\mathbb R^n) being such that \eta_\Omega = 1 in an open neighbourhood of \Omega . Indeed, since u_{s, \Omega}-\eta_\Omega\in \mathcal{D}^{s, 2}(\mathbb R^n\setminus\Omega) , there exists a sequence \{v_k\}_k \subseteq C_c^\infty(\mathbb R^n\setminus\Omega) such that v_k\to u_{s, \Omega}-\eta_\Omega in \mathcal{D}^{s, 2}(\mathbb R^n) as k\to\infty . If we now let u_k: = v_k+\eta_\Omega we have that u_k\in C_c^\infty(\mathbb R^n) and u_k = 1 in an open \Omega_k\supseteq\Omega . Now, if we consider the function
u_k^{\varepsilon}: = \frac{(u_k-{\varepsilon})^+}{1-{\varepsilon}} |
we have that u_k^{\varepsilon}\in \mathcal{D}^{s, 2}(\mathbb R^n) and u_k^{\varepsilon}-\eta_\Omega\in \mathcal{D}^{s, 2}(\mathbb R^n\setminus\Omega) . Moreover, u_k^{\varepsilon}\to u_{s, \Omega}^\varepsilon in \mathcal{D}^{s, 2}(\mathbb R^n) as k\to\infty , thus proving the claim. We also observe that the family \{u_{s, \Omega}^{\varepsilon}\}_{s\in (s_0, 1)} satisfies the following properties:
1). there exists \bar R = \bar R(\varepsilon) > 0 , depending only on \varepsilon , such that
\mbox{supp }u_{s, \Omega}^\varepsilon \subset B_{\bar R}\, \, \mbox{for any } s\in(s_0, 1). |
This follows by the upper bound (4.2): we choose \bar R > \left(\frac{2{R_0^{n-2s}}}{\varepsilon}\right)^{\frac{1}{n-2}} with R_0 being such that \Omega \subseteq B_{R_0} .
In particular, this implies that u_{s, \Omega}^\varepsilon\in \widetilde{W}_0^{s, 2}(B_{\bar{R}}) , where, for any open \mathcal{O} \subseteq \mathbb R^n we denote
\begin{equation*} \widetilde{W}_0^{s, 2}(\mathcal{O}): = \left\{ u\in L^1_{ \rm{loc}}( \mathbb R^n)\colon [u]_s < \infty\; \text{and }u = 0\; \text{in } \mathbb R^n\setminus\mathcal{O} \right\}, \end{equation*} |
which, in case \mathcal{O} is bounded and has Lipschitz boundary, coincides with the space \mathcal{D}^{s, 2}(\mathcal{O}) , see [7,Proposition B.1];
2). there holds
\begin{equation} (1-s)[u_{s, \Omega}^{\varepsilon}]_s^2 \leq (1-s)\frac{[u_{s, \Omega}]_s^2}{(1-\varepsilon)^2}\le C_1, \end{equation} | (4.3) |
for any \varepsilon > 0 and s\in (s_0, 1) , with C_1 > 0 independent of \varepsilon and s . This is a direct consequence of (1.9).
Hence we can apply [7,Proposition 3.6] to the family \{u_{s, \Omega}^{\varepsilon}\}_{s\in(s_0, 1)} to deduce that there exists an increasing sequence s_k \in (s_0, 1) converging to 1 and a function u_{\Omega}^\varepsilon\in W^{1, 2}_0(B_{\bar R}) such that
\lim\limits_{k\rightarrow \infty} \|u_{s_k, \Omega}^{\varepsilon}-u_{\Omega}^\varepsilon\|_{L^2(B_{\bar R})} = 0. |
Analogously, being \Omega a Lipschitz domain, we know that u_{s, \Omega}^{\varepsilon}-\eta_\Omega\in \widetilde{W}_0^{s, 2}(B_{\bar{R}}\setminus\Omega) and that
(1-s)[u_{s, \Omega}^{\varepsilon}-\eta_\Omega]_s^2\leq C_2 |
for all \varepsilon > 0 and s\in (s_0, 1) , with C_2 > 0 independent of \varepsilon and s . Therefore, we can apply [7,Proposition 3.6] to the family \{u_{s, \Omega}^{\varepsilon}-\eta_\Omega\}_{s\in(s_0, 1)} as well, and this entails the existence of a (not relabeled) subsequence s_k\in(s_0, 1) and of a function v_{\Omega}^{\varepsilon}\in W_0^{1, 2}(B_{\bar{R}}\setminus\Omega) such that
\lim\limits_{k\rightarrow \infty} \|u_{s_k, \Omega}^{\varepsilon}-\eta_\Omega-v_{\Omega}^\varepsilon\|_{L^2(B_{\bar R})} = 0, |
where the functions are trivially extended in \Omega . As a consequence, we obtain that
u_\Omega^{\varepsilon}-\eta_\Omega = v_\Omega^{\varepsilon}\in W_0^{1, 2}(B_{\bar{R}}\setminus \Omega), |
which, in turn, implies that the trivial extension of u_\Omega^{\varepsilon} to the whole \mathbb R^n is an admissible competitor for {\mathrm{cap}}(\Omega) . Hence we have
\begin{align*} \frac{\omega_n}{2}{\rm{cap}}(\Omega) \leq \frac{\omega_n}{2}\int_{ \mathbb R^n}|\nabla u^\varepsilon_{\Omega}|^2\, dx &\leq \liminf\limits_{s\nearrow 1}(1-s)[u^\varepsilon_{s, \Omega}]_s^2\\ &\leq \frac{1}{(1-\varepsilon)^2}\liminf\limits_{s\nearrow 1}(1-s)[u_{s, \Omega}]_s^2, \end{align*} |
where in the second inequality we have used the \Gamma -convergence result by Brasco, Parini, Squassina (more precisely, Proposition 3.11 in [7]) and in the last one (4.3). Finally, we conclude by letting \varepsilon \rightarrow 0 .
Remark 4.1. Thanks to Theorem 1.2 it is possible to explicitly compute the limit as s\to 1^- of the constant C_{n, s} as in Theorem 1.1, which coincides with the constant C_n appearing in Corollary 1.3. Indeed, from the definitions of \alpha_{n, s} and c_{n, s} , given in (2.10) and (2.7) respectively, and the property of the Gamma function, it is easy to see that
\lim\limits_{s\to 1^-}\alpha_{n, s} = \lim\limits_{s\to 1^-}c_{n, s} = \pi^{-\frac{n}{2}}\Gamma\left(\frac{n+2}{2}\right). |
Moreover, if B denotes the unitary ball in \mathbb R^n , in view of Theorem 1.2 we have that
\lim\limits_{s\to 1^-}(1-s){{{\rm{cap}}}_s}(B) = \frac{\omega_n}{2}{{{\rm{cap}}}}(B) = \frac{n(n-2)}{2}\omega_n^2, |
see e.g., [27,Theorem 2.8,point (i) ] for the explicit value of {{{\rm{cap}}}}(B) . Thanks to these facts, Remark 3.5 and basic calculus, it is easy to see that
C_n = \lim\limits_{s\to 1^-}C_{n, s} = \max\left\{ 2^{-3}, \frac{\kappa_1^2 C_3\pi^{\frac{n}{2}}}{\Gamma\left(\frac{n+2}{2}\right)}\frac{2}{n(n-2)\omega_n^2} \right\}, |
where
\kappa_1: = \frac{\lambda_1}{4(1+2\lambda_1)}, \quad\text{with}\quad\lambda_1: = \frac{n-2}{10n}, |
and C_3 as in (3.13).
E. Cinti is partially supported by MINECO grant MTM2017-84214-C2-1-P, by Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM), and is part of the Catalan research group 2017 SGR 1392. B. Ruffini acknowledges partial support from the ANR-18-CE40-0013 SHAPO financed by the French Agence Nationale de la Recherche (ANR). R. Ognibene acknowledges support from the MIUR-PRIN project No. 2017TEXA3H. A special thank goes to L. Brasco for several fruitful discussions and suggestions. Eventually, the authors thank Italy national football team for providing a joyful atmosphere during the last stages of the drafting process of this work.
The authors declare no conflict of interest.
[1] |
F. J. Almgren, E. H. Lieb, Symmetric decreasing rearrangement is sometimes continuous, J. Amer. Math. Soc., 2 (1989), 683–773. doi: 10.1090/S0894-0347-1989-1002633-4
![]() |
[2] |
V. Andrievskiĭ, W. Hansen, N. Nadirashvili, Isoperimetric inequalities for capacities in the plane, Math. Ann., 292 (1992), 191–195. doi: 10.1007/BF01444617
![]() |
[3] |
L. Brasco, E. Cinti, S. Vita, A quantitative stability estimate for the fractional Faber-Krahn inequality, J. Funct. Anal., 279 (2020), 108560. doi: 10.1016/j.jfa.2020.108560
![]() |
[4] | L. Brasco, G. De Philippis, Spectral inequalities in quantitative form, In: Shape optimization and spectral theory, Warsaw, Poland: De Gruyter Open Poland, 2017,201–281. |
[5] |
L. Brasco, D. Gómez-Castro, J. L. Vázquez, Characterisation of homogeneous fractional Sobolev spaces, Calc. Var., 60 (2021), 60. doi: 10.1007/s00526-021-01934-6
![]() |
[6] |
L. Brasco, S. Mosconi, M. Squassina, Optimal decay of extremals for the fractional Sobolev inequality, Calc. Var., 55 (2016), 23. doi: 10.1007/s00526-016-0958-y
![]() |
[7] | L. Brasco, E. Parini, M. Squassina, Stability of variational eigenvalues for the fractional p-Laplacian, DCDS, 36 (2016), 1813–1845. |
[8] |
L. Brasco, A. Salort, A note on homogeneous Sobolev spaces of fractional order, Annali di Matematica, 198 (2019), 1295–1330. doi: 10.1007/s10231-018-0817-x
![]() |
[9] | C. Bucur, E. Valdinoci, Nonlocal diffusion and applications, Bologna: Springer Cham, 2016. |
[10] |
X. Cabré, Y. Sire, Nonlinear equations for fractional Laplacians, Ⅰ: Regularity, maximum principles, and Hamiltonian estimates, Ann. Inst. H. Poincaré Anal. Non Linéaire, 31 (2014), 23–53. doi: 10.1016/j.anihpc.2013.02.001
![]() |
[11] |
L. Caffarelli, L. Silvestre, An extension problem related to the fractional Laplacian, Commun. Part. Diff. Eq., 32 (2007), 1245–1260. doi: 10.1080/03605300600987306
![]() |
[12] |
A. Cianchi, N. Fusco, Functions of bounded variation and rearrangements, Arch. Rational Mech. Anal., 165 (2002), 1–40. doi: 10.1007/s00205-002-0214-9
![]() |
[13] |
M. Cicalese, G. P. Leonard, A selection principle for the sharp quantitative isoperimetric inequality, Arch. Rational Mech. Anal., 206 (2012), 617–643. doi: 10.1007/s00205-012-0544-1
![]() |
[14] |
A. Cotsiolis, N. K. Tavoularis, Best constants for Sobolev inequalities for higher order fractional derivatives, J. Math. Anal. Appl., 295 (2004), 225–236. doi: 10.1016/j.jmaa.2004.03.034
![]() |
[15] | S. Dipierro, M. Medina, E. Valdinoci, Fractional elliptic problems with critical growth in the whole of \mathbb{R}^n, Pisa: Edizioni della Normale, 2017. |
[16] | L. C. Evans, R. F. Gariepy, Measure theory and fine properties of functions, Boca Raton, FL: CRC Press, 2015. |
[17] |
A. Figalli, F. Maggi, A. Pratelli, A mass transportation approach to quantitative isoperimetric inequalities, Invent. Math., 182 (2010), 167–211. doi: 10.1007/s00222-010-0261-z
![]() |
[18] |
L. E. Fraenkel, On the increase of capacity with asymmetry, Comput. Methods Funct. Theory, 8 (2008), 203–224. doi: 10.1007/BF03321684
![]() |
[19] |
R. L. Frank, R. Seiringer, Non-linear ground state representations and sharp Hardy inequalities, J. Funct. Anal., 255 (2008), 3407–3430. doi: 10.1016/j.jfa.2008.05.015
![]() |
[20] |
N. Fusco, F. Maggi, A. Pratelli, The sharp quantitative isoperimetric inequality, Ann. Math., 168 (2008), 941–980. doi: 10.4007/annals.2008.168.941
![]() |
[21] | N. Fusco, F. Maggi, A. Pratelli, Stability estimates for certain Faber-Krahn, isocapacitary and Cheeger inequalities, Ann. Sc. Norm. Super. Pisa Cl. Sci., 8 (2009), 51–71. |
[22] |
N. Fusco, V. Millot, M. Morini, A quantitative isoperimetric inequality for fractional perimeters, J. Funct. Anal., 261 (2011), 697–715. doi: 10.1016/j.jfa.2011.02.012
![]() |
[23] |
R. R. Hall, W. K. Hayman, A. W. Weitsman, On asymmetry and capacity, J. Anal. Math., 56 (1991), 87–123. doi: 10.1007/BF02820461
![]() |
[24] |
W. Hansen, N. Nadirashvili, Isoperimetric inequalities in potential theory, Potential Anal., 3 (1994), 1–14. doi: 10.1007/BF01047833
![]() |
[25] | W. Hansen, N. Nadirashvili, Isoperimetric inequalities for capacities, In: Harmonic analysis and discrete potential theory (Frascati, 1991), Plenum, New York, 1992,193–206. |
[26] | G. H. Hardy, J. E. Littlewood, G. Pólya, Inequalities, 2 Eds., Cambridge University Press, 1952. |
[27] | J. Malý, W. P. Ziemer, Fine regularity of solutions of elliptic partial differential equations, Providence, RI: American Mathematical Society, 1997. |
[28] | E. Mukoseeva, The sharp quantitative isocapacitary inequality (the case of p-capacity), Adv. Calc. Var., 2021, doi: 10.1515/acv-2020-0106. |
[29] |
G. de Philippis, M. Marini, E. Mukoseeva, The sharp quantitative isocapacitary inequality, Rev. Mat. Iberoam., 37 (2021), 2191–2228. doi: 10.4171/rmi/1259
![]() |
[30] |
O. Savin, E. Valdinoci, Density estimates for a nonlocal variational model via the Sobolev inequality, SIAM J. Math. Anal., 43 (2011), 2675–2687. doi: 10.1137/110831040
![]() |
[31] |
Y. Sire, S. Terracini, S. Vita, Liouville type theorems and regularity of solutions to degenerate or singular problems part Ⅰ: even solutions, Commun. Part. Diff. Eq., 46 (2021), 310–361. doi: 10.1080/03605302.2020.1840586
![]() |
[32] |
M. Warma, The fractional relative capacity and the fractional Laplacian with Neumann and Robin boundary conditions on open sets, Potential Anal., 42 (2015), 499–547. doi: 10.1007/s11118-014-9443-4
![]() |
1. | Dario Mazzoleni, Benedetta Pellacci, Calculus of variations and nonlinear analysis: advances and applications, 2023, 5, 2640-3501, 1, 10.3934/mine.2023059 |