A Faber-Krahn inequality
for mixed local and nonlocal operators
Abstract.
We consider the first Dirichlet eigenvalue problem for a mixed local/nonlocal elliptic operator and we establish a quantitative Faber-Krahn inequality. More precisely, we show that balls minimize the first eigenvalue among sets of given volume and we provide a stability result for sets that almost attain the minimum.
Key words and phrases:
Operators of mixed order, first eigenvalue, shape optimization, isoperimetric inequality, Faber-Krahn inequality, quantitative results, stability.2020 Mathematics Subject Classification:
49Q10, 35R11, 47A75, 49R051. Introduction
At the end of the XIX century, relying on explicit calculations on suitable domains, John William Strutt, 3rd Baron Rayleigh, conjectured that the ball is the minimizer of the first Dirichlet eigenvalue among the domains of a given volume, see [52]. The confirmation of this conjecture entails a number of interesting physical consequences, such as:
-
•
among all drums of a given surface, the circular drum produces the lowest voice,
-
•
among all the regions of a given volume with the boundary maintained at a constant (say, zero) temperature, the one which dissipates heat at the slowest possible rate is the sphere.
Also, the statement with equal volume constraints gives as a byproduct the one with equal perimeter constraint (thanks to the scaling property of the first eigenvalue and the isoperimetric inequality). In this sense, the first attempt to prove Lord Rayleigh’s conjecture dates back to 1918, when Richard Courant established the above claim with equal perimeter constraint, see [19]. Then, using rearrangement methods and the variational characterization of eigenvalues, the original conjecture with volume constraint was established independently by Georg Faber and Edgar Krahn, see [33, 47, 48]. See also [40, Chapter 2] and [44]. We refer to [20] for similar results in the context of composite membranes.
Given that balls are actually established to be the unique minimizers for the first eigenvalue under volume constraint (hence if the first eigenvalue is equal to that of the corresponding ball, then the domain must necessarily be a ball), an intense research activity focused on quantitative versions of the Faber-Krahn inequality: roughly speaking, if the eigenvalue is “close to the one of the ball”, can one deduce that the domain is also “close to a ball”? Classical results in this direction have been obtained by Wolfhard Hansen and Nikolai Nadirashvili in [39] and Antonios Melas in [49], and sharp bounds in terms of the so called Fraenkel asymmetry have been obtained recently by Lorenzo Brasco, Guido De Philippis and Bozhidar Velichkov in [12]. See also [2] for some stability results in space forms.
The goal of this paper is to obtain a Faber-Krahn inequality and a quantitative version of it for an elliptic operator of mixed order. More specifically, for the sake of simplicity, we will focus on operators obtained by the superposition of a classical and a fractional Laplacian, namely of operators of the form
with and
Operators of this type present interesting mathematical questions, especially due to the lack of scale invariance and in view of the combination of local and nonlocal behaviors, see [41, 42, 5, 22, 10, 18, 3, 17, 4, 53, 16, 23, 32, 24, 8, 7, 29, 15, 1, 14, 21, 45]. Moreover, they possess a concrete interest in applications since they model diffusion patterns with different time scales (loosely speaking, the higher order operator leading the diffusion for small times and the lower order operator becoming predominant for large times) and they arise in bi-modal power-law distribution processes, see [51]. Further applications arise in the theory of optimal searching strategies, biomathematics and animal foraging, see [30, 31] and the references therein. See also [11] for further applications.
In our setting, given a bounded open subset of , we consider the first Dirichlet eigenvalue (see Section 3 for a detailed presentation) and we characterize the optimal set by the following result:
Theorem 1.1 (Faber-Krahn inequality for ).
Let be a bounded open set with boundary of class . Let , and let be any Euclidean ball with volume . Then,
(1.1) |
Moreover, if the equality holds in (1.1), then is a ball.
A related Faber-Krahn inequality has been recently obtained for radially symmetric, nonnegative and continuous kernels with compact support in [46]. With this respect, the case treated here of singular kernels seems to be new to the best of our knowledge. Additionally, and more importantly, we establish a stability result for inequality (1.1):
Theorem 1.2 (Quantitative Faber-Krahn inequality for ).
Let . Let be an open, bounded and uniformly convex set with boundary of class .
Then, there exists with the following property: if is a ball with and
(1.2) |
then there exist two balls , such that and
(1.3) |
and
(1.4) |
where is a structural constant.
The proof of Theorem 1.1 is based on the definition of principal eigenvalue of and on the Polya-Szegö inequality.
The proof of Theorem 1.2 is more involved and requires some delicate geometric arguments. For this, a careful analysis of the superlevel sets of the principal eigenfunction is needed, combined with the use of a Bonnesen-type inequality, which is a strengthening of the classical isoperimetric inequality.
We stress that, while in Theorem 1.1 we only require the boundary of to be , in Theorem 1.2 we need to ask that the boundary of is of class . This is due to the fact that, in order to prove Theorem 1.2, we have to use the Faber-Krahn inequality in (1.1) for the superlevel sets of the principal eigenfunction and we are able to prove that these sets are convex and with boundary of class under the condition that has a -boundary, see the forthcoming Lemma 3.7.
The rest of this paper is organized as follows. In Section 2 we introduce the basic notation and the setting in which we work, and we provide some regularity results and an Hopf-type lemma for the operator . In Section 3 we introduce the notion of principal Dirichlet eigenvalue for and we give some regularity results for the associated eigenfunction. Section 3 also contains a result (namely, Lemma 3.7) that proves the convexity of the superlevel sets of the first eigenfunction near the boundary of a convex domain: this result, based on a detailed use of the Inverse Function Theorem, highlights an interesting technical difference with respect to the classical case where one can exploit the convexity of all the level sets of the Dirichlet principal eigenfunction, due to its concavity as a function (which is unknown in the fractional case, see [43]).
2. Basic notions and preliminary results
In this section we properly introduce the relevant definitions and notation which shall be used throughout the rest of the paper. Moreover, we review/establish some regularity results concerning our operator .
Let be a bounded open set with boundary of class . We then consider the space defined as follows:
In view of the regularity of , it is well-known that can be naturally identified with ; more precisely, we have (see, e.g., [13, Prop. 9.18])
(2.1) |
where denotes the indicator function of . Throughout what follows, we tacitly identify a function with its “zero-extension” .
We then observe that, as a consequence of (2.1), the set is endowed with a structure of real Hilbert space by the scalar product
The norm associated with this scalar product is
and the (linear) map defined by
turns out to be a bijective isometry between and .
Let . On the space , we consider the bilinear form
moreover, for every we define
(2.2) |
Remark 2.1.
Using the bilinear form , we can give the following definition.
Definition 2.2.
Let . We say that a function is a weak solution of the -Dirichlet problem
if it satisfies the following properties:
-
(1)
;
-
(2)
for every test function , one has
Remark 2.3.
Let and . Since is dense in , we see that is a weak solution of if and only if
Then, by applying the Lax-Milgram Theorem to the bilinear form , one can prove the following existence result (see, e.g., [7, Theorem. 1.1]).
Theorem 2.4.
For every , there exists a unique weak solution of , further satisfying the ‘a-priori’ estimate
Here, is a constant independent of .
Since we aim at proving some global regularity results for , it is convenient to fix also the definition of classical solution of problem . For this, we recall the notation .
Definition 2.5.
Let . We say that a function is a classical solution of if it satisfies the following properties:
-
(1)
;
-
(2)
pointwise in ;
-
(3)
pointwise for every .
Remark 2.6.
(1) We explicitly observe that, if it is possible to compute pointwise for every . Indeed, we have
and the regularity of ensures that the “second-order” different quotient
is in . To be more precise, for every we have
where is a suitable constant and is such that .
(2) Assume that , and let be the (unique) weak solution of , according to Theorem 2.4. If we further assume that , we can compute
Then, a standard integration-by-parts argument shows that is also a classical solution of . Conversely, if is a classical solution of such that , then is also a weak solution of .
The first regularity result that we aim to prove is a global -regularity theorem, which relies on the -theory for developed by Bensoussan and Lions [6].
Theorem 2.7.
Let and let denote the unique weak solution of . Moreover, let us assume that is of class . Then,
(2.5) |
Proof.
Note that for every . We111For instance, in the most delicate case , one can use the setting in [35] with , , , , (with conveniently small), , and . Alternatively to [35], one can use [6, Theorem 3.2.3]. utilize [35, Theorem 3.1.22] to obtain -regularity for some . By combining this with the Sobolev embedding theorem, we obtain (2.5). ∎
Furthermore, we recall that the -regularity of when can be deduced by applying [35, Theorem 3.1.12]. We refer to Appendix B for an explicit proof.
Theorem 2.8.
Let and . Suppose that is of class . If and if denotes the unique weak solution of (according to Theorem 2.4), then
In particular, is a classical solution of .
We also recall that a regularity result up to the boundary for eigenfunctions of mixed operators has been recently obtained in [25] for radially symmetric, nonnegative and continuous kernels with compact support.
We close this section by stating, for future reference, the following Hopf-type lemma for our operator . Similar statements for this type of operators have been proved in [35, Theorem 3.1.5], but with the additional assumption that .
Theorem 2.9.
Let and . Let be such that
(2.6) |
Let . Assume that on and that in . Suppose also that and that is of class . Finally, let be the outer unit normal to at . Then,
(2.7) |
Proof.
Without loss of generality, we suppose that and that coincides with the origin. In particular,
(2.8) |
Also, by the regularity of , we suppose that there exists such that touches the boundary of at the origin. We remark that, by assumption, there exists a point such that , and therefore, by continuity, there exists a ball such that in . In light of this and (2.8), we can find sufficiently small such that
Now, we claim that there exists such that
(2.9) |
To prove this, we suppose by contradiction that for every there exists
such that . This implies that, for every ,
(2.10) |
Notice that as , and therefore, by the Dominated Convergence Theorem, one sees that
for some independent of . Plugging this information into (2.10), we obtain that, for every ,
(2.11) |
Now we claim that there exists such that
(2.12) | in and in . |
To this end, we consider a -diffeomorphism such that
-
(a)
for a suitable ;
-
(b)
;
-
(c)
.
Possibly reducing we can also assume that
For all we define and we notice that is a function in and along . Thus, we define
We observe that is continuous across and that
This gives that is a function in . Moreover, we claim that
(2.13) | is a function in . |
To check this, it suffices to consider and with and observe that
thus establishing (2.13).
Hence, for every we define and we infer from (2.13) that . Furthermore, if , then
and, as a result, we have that . If instead , then . Hence, letting
and using that , we get in this case that
These observations complete the proof of (2.12).
By (2.12) it follows that
Hence, if ,
and accordingly
This and (2.11) entail that , and we thereby obtain a contradiction by choosing conveniently small. This completes the proof of (2.9).
From (2.6) and (2.9) we deduce that there exists such that
Moreover, we have in . Indeed, if there exists a point such that , then by the Maximum Principle (see [7, Theorem 1.4]) we would have that in , which would contradict our hypothesis. As a consequence, we are in the position of applying the Hopf Lemma for the classical Laplacian (see e.g. [36, Lemma 3.4]) thus obtaining the desired result in (2.7). ∎
Remark 2.10.
We stress that
(2.14) |
As a counterexample, let be the minimizer of
among the functions in such that outside , for a given smooth function such that
By the Sobolev Embedding Theorem, we know that is continuous in and, in view of its minimality property, it satisfies in . Thus, let be such that . We claim that
(2.15) |
Indeed, if not, we would have that in . Since outside we know from the weak maximum principle (see e.g. [7, Theorem 1.2]) that
and consequently vanishes identically in . In particular, the origin would provide an interior maximum for . Accordingly, by the strong maximum principle (see e.g. [7, Theorem 1.4]), we gather that vanishes identically in . Since this is a contradiction with the values of in , the proof of (2.15) is completed. Now, from (2.15), we deduce that
(2.16) |
We also take such that . By construction, we have that is a minimizing point for in and that in . Thus, equation (2.16) proves the observation in (2.14).
3. The principal Dirichlet eigenvalue for
In this section we introduce the notion of principal Dirichlet eigenvalue for , and we prove some regularity results for the associated eigenfunctions. In what follows, we tacitly inherit all the assumptions and notation introduced so far; in particular, and is a bounded open set with boundary.
To begin with, we give the following definition.
Definition 3.1.
We define the principal Dirichlet eigenvalue of in as
(3.1) |
where is the quadratic form defined in (2.2).
Remark 3.2.
Before proceeding we list, for a future reference, some simple properties of which easily follow from its very definition.
-
(1)
For every , one has
-
(2)
For every , one has
-
(i)
if ;
-
(ii)
if .
-
(i)
-
(3)
Let be the principal Dirichlet eigenvalue of in , i.e.,
Then, one has the bound
-
(4)
There exists a positive constant such that
Then, by using standard arguments of the Calculus of Variations (see, e.g., the approach in [54, Proposition 9]), it is possible to prove the following result.
Theorem 3.3.
The infimum in (3.1) is achieved. More precisely, there exists a unique such that:
-
(1)
and ;
-
(2)
a.e. in .
Furthermore, is a weak solution of the following problem
(3.2) |
namely
(3.3) |
Definition 3.4.
We refer to given by Theorem 3.3 as the principal eigenfunction of (in ).
We devote the rest of this section to prove some regularity properties of . To begin with, we establish the following global boundedness result.
Theorem 3.5.
The principal eigenfunction is globally bounded on .
The proof of Theorem 3.5 relies on the classical method by Stampacchia, and is essentially analogous to that of [7, Theorem 4.7] (see also [28, 55]). However, we present it here with all the details for the sake of completeness.
Proof of Theorem 3.5.
To begin with, we arbitrarily fix and we set
Moreover, for every , we define and
We explicitly point out that, in view of these definitions, one has
-
(a)
;
-
(b)
(since );
-
(c)
and (since ).
We now observe that, since , we have ; furthermore, since a.e. in , one also has
and thus (remind that is bounded). We are then entitled to use the function as a test function in (3.3), obtaining
(3.4) |
To proceed further we notice that, for a.e. , one has
(3.5) |
Moreover, taking into account the definition of , we get
(3.6) |
Gathering together (3.4), (3.5) and (3.6), we obtain
(3.7) |
We then claim that, for every , one has
(3.8) |
Indeed, if , we have ; as a consequence,
which is exactly the claimed (3.8). By combining (3.8) with (3.7), and taking into account that a.e. in , for every we get
(3.9) |
We now turn to estimate from below the left-hand side of (3.9). To this end we first observe that, owing to the very definition of , we have
As a consequence, we obtain
(3.10) |
Using the Hölder inequality (with exponents and , being the Sobolev exponent), jointly with the Sobolev inequality, from (3.9)-(3.10) we obtain the following estimate for every
(3.11) |
where is the Sobolev constant and .
Now, if is sufficiently regular, by combining Theorem 3.5 with Theorems 2.7-2.8 we obtain the next key result.
Theorem 3.6.
Let and be such that
Let us suppose that is of class . Denoting by the principal eigenfunction of in (according to Theorem 3.3), we have
Moreover, is a classical solution of (3.2), that is,
(3.12) |
Furthermore, satisfies the following additional properties:
-
(1)
pointwise in ;
-
(2)
denoting by the external unit normal to , we have
(3.13)
Proof.
First of all we observe that, setting , from Theorem 3.5 we infer that ; as a consequence, since is the (unique) weak solution of
(and is of class ), we deduce from Theorem 2.7 that
In particular, . In view of this fact, and taking into account our assumptions on and on , we can apply Theorem 2.8 to : therefore,
and is a classical solution of , that is, satisfies (3.12). To complete the proof, we now turn to prove the validity of properties (1)-(2).
We want to stress that one can get almost everywhere positivity of by means of the strong maximum principle for weak solutions proved in [14, 9].
Finally, by exploiting Theorem 3.6, we can prove Lemma 3.7 below. We stress that in the purely local regime, the following lemma has been proved in [57] without any restriction on . On the contrary, the same property is not known to hold in the purely fractional case. As a matter of fact, convexity and superharmonicity properties for fractional eigenfunctions can reserve surprisingly severe difficulties, see e.g. [43].
Lemma 3.7.
Let and let us suppose that is uniformly convex, and that is of class . Then, denoting by the principal eigenfunction of in , there exists such that, for every fixed , the set
Moreover, is of class .
Proof.
We split the proof into three steps.
Step I. In this first step we prove that, if is sufficiently small, the superlevel set can be realized as a suitable “deformation” of . To this end we first notice that, owing to the assumptions on , Theorem 3.6 ensures that
-
(a)
, for every such that , and is a classical solution of (3.2);
-
(b)
pointwise in ;
-
(c)
for every .
Moreover, since on , we have
Now, since property (c) implies that on , we get
(3.15) |
as a consequence, since , again by property (c) we have
(3.16) |
Hence, it is possible to find some such that
(3.17) |
Furthermore, bearing in mind that is of class , by possibly shrinking we can also suppose that (see, e.g., [36, Section 14.6])
-
(i)
is of class on ;
-
(ii)
for every there exists a unique such that
For a fixed , we then consider the function
Since is of class on , by (3.15)-(3.16) we have
as a consequence, again by shrinking if necessary, we obtain
(3.18) |
In particular, is strictly increasing on , and
(3.19) |
To proceed further, we consider the set and we notice that, since in , we have
(3.20) |
Hence, we define
(3.21) |
Taking into account (3.19), it is immediate to see that, for every , there exists a unique point such that
then, we set
see Figure 1.

Before proceeding we highlight, for future reference, the following fact: since and is strictly increasing on , we have
(3.22) |
We now turn to prove that, if , one has
(3.23) |
To this end we first observe that, by the very definition of , we have on ; moreover, since is strictly increasing on , one also has
and thus out of . Finally, we show that
Let then be fixed. If , by (3.20)-(3.21) we have
If, instead, , one has , and thus . By property (ii) of , we know that there exists a unique such that
on the other hand, since , we necessarily have
This, together with the strict monotonicity of on , implies that
and completes the proof of (3.23).
Step II. We now turn to prove that, if is as in Step I and , the boundary of is of class . To this end we observe that, by crucially exploiting identity (3.23) and the very definition of , one has
(3.24) |
This, together with (3.17), shows that
and thus is of class (actually, it is of class ).
Step III. In this last step we prove that, if (where is as in Step I), the set is convex. As in Step II, we use in crucial way identity (3.23).
To begin with, we consider a covering of of finitely many small open balls
such that for every we can write as a graph in some coordinate direction (where is the concentric ball of with twice the radius of ). For sufficiently small, we can suppose that
In this setting, it suffices to check that can locally be written as a graph of a convex function for all . Without loss of generality, we argue for and assume that
for some satisfying
(3.25) |
Thus, by (3.24), in this chart can be locally parameterized by
(3.26) |
where we used the notation and belongs to a domain of . We now introduce the function
Let also . We notice that, as in (3.18), for small enough,
Also is locally a function and . As a consequence, we find that is locally a function. We also observe that, by (3.26), the set can be locally parameterized by
(3.27) |
Now we set
Observe that is locally a function. Moreover, by (3.22) we have
(3.28) |
and, up to a rotation, we can focus our analysis at a point for which
(3.29) |
Therefore
(3.30) | the Jacobian matrix of at is the identity |
and we can exploit the Inverse Function Theorem, denote by the inverse function of in the vicinity of and have that is also a function. Using the notation and , we have that and thus
Additionally,
and
Accordingly, recalling (3.27), we can locally parameterize as
with
For this reason, to complete the proof of the convexity property of , it suffices to check the convexity of the function
(3.31) |
that describes its graph. To this end, we remark that, for each , ,
that is
Setting , we thus deduce from (3.30) that
Also, by (3.29), we have that the gradient of at vanishes. With these items of information, we find that
and
Besides, owing to (3.28), we have
and therefore
As a result, recalling the definition of in (3.31),
By the uniform convexity of the domain in (3.25), we obtain that
This, together with the fact that is a function in , gives
in a neighborhood of , and proves that is uniformly convex, as desired. ∎
4. A Faber-Krahn inequality and proofs of Theorems 1.1 and 1.2
This section is devoted to show a quantitative Faber-Krahn inequality for and in particular to prove Theorems 1.1 and 1.2. In what follows, denotes the principal Dirichlet eigenvalue of (in ), as given by Definition 3.1, and the corresponding principal eigenfunction (according to Theorem 3.3).
We begin by proving Theorem 1.1.
Proof of Theorem 1.1.
Let be the (unique) Euclidean ball with centre and volume . If is the principal eigenfunction of in , we define
as the (decreasing) Schwarz symmetrization of . Now, since , it follows from a well-known theorem by Polya-Szegö that
(4.1) |
furthermore, by [34, Theorem A.1] we also have
(4.2) |
Gathering together all these facts and recalling (1) in Theorem 3.3, we then get
(4.3) |
From this, reminding that is translation-invariant (see Remark 3.2-(1)), we derive the validity of (1.1) for every Euclidean ball with volume .
To complete the proof of Theorem 1.1, let us suppose that
for some (and hence, for every) ball with . By (4.3) we have
In particular, by (4.1)-(4.2) we get
We are then entitled to apply once again [34, Theorem A.1], which ensures that must be proportional to a translate of a symmetric decreasing function. As a consequence of this fact, we immediately deduce that
must be a ball (up to a set of zero Lebesgue measure). ∎
Now that we have established Theorem 1.1, we can finally prove the main result of this paper, namely, the quantitative version of inequality (1.1) presented in Theorem 1.2.
The proof of Theorem 1.2 is based on the following technical lemma.
Lemma 4.1.
Let . Let be bounded open set with boundary of class . Let , and let be any Euclidean ball with volume . Moreover, let be the principal eigenfunction of in , let be as in Lemma 3.7, and let
(4.4) |
Let also .
Then, there exists a small , only depending on and , with the following property: if is such that
(4.5) |
then we have the estimate
Proof.
First of all we observe that, since is of class , from Theorem 3.6 we derive that the principal eigenfunction of in satisfies
(actually, if ); moreover, since , by Lemma 3.7 we know that is of class . We then consider the function
Since and on , it is readily seen that ; as a consequence, since is a weak solution of (3.2), we have
where we have used Hölder’s inequality and the fact that .
On the other hand, since in and (by the choice of and the definition of ), an application of Minkowski’s inequality gives
Gathering together all these estimates, and reminding (4.5), we obtain
(4.6) |
Using Remark 3.2-(2) and the Faber-Krahn inequality in Theorem 1.1 (notice that we are in the position of applying Theorem 1.1 for the set in light of the regularity result in Lemma 3.7, and notice also that ), we get
(4.7) |
By combining (4.6) with (4.7), we then get
(4.8) |
Finally, if is sufficiently small and , we obtain
From this and (4.8), we obtain the desired result. ∎
Now we provide a convexity result that turns out to be useful for the proof of Theorem 1.2:
Lemma 4.2.
Let be open, bounded and convex. Then, there exists with the following property: if there exists a ball such that
(4.9) |
then there exists a ball , which is concentric to , such that and
(4.10) |
Here, the positive constant depends only on .
Proof.
Up to a translation, we can assume that is centered at the origin. We assume that has radius and we take maximizing the distance to the origin among points in . Let also . By construction, we have that (since contains ) and that is contained in the ball of radius . Since , up to a rotation we can suppose that . We also consider the convex hull of and . By construction, lies in the closure of . Let be the right circular cone obtained by intersecting and the halfspace . Notice that the height of the cone is equal to and we denote by the radius of its basis. By triangular similitude (see the triangles and in Figure 2 on page 2), we see that
As a consequence,
(4.11) |
for some depending only on .
On the other hand, in view of (4.9),
Combining this and (4.11), we have that
(4.12) |
Also, by (4.9) we know that, for sufficiently small,
for some depending only on . From this and (4.12), up to renaming we conclude that
(4.13) |
In spite of some comments appeared in the literature, we believe that the exponent in formula (4.10) of Lemma 4.2 is optimal, as remarked in Appendix A.
Proof of Theorem 1.2.
Along the proof, constants depending only on , and may change passing from a line to another. Nevertheless, to avoid a cumbersome notation, we will keep the same symbol for all of them.
Let be the decreasing Schwarz symmetrization of the first eigenfunction (given by Theorem 3.3). We recall from Theorem 3.3 that we can assume , and hence as well. We define the sets
and
We further define and the function
We recall that
(4.17) |
and, thanks to the classical isoperimetric inequality,
(4.18) |
We also recall that is the boundary of the set and is the boundary of a ball with volume equal to . Therefore,
(4.19) |
where, as usual, denotes the -dimensional Lebesgue measure of the unit ball.
Now we take to be as in Lemma 4.1 and as in Lemma 4.2. We also define , where is as in Lemma 3.7. We stress that , and are small quantities depending only on the structural parameters of the problem and we suppose that the parameter in the statement of Theorem 1.2 satisfies
(4.20) |
Step I. We first prove that, if and (1.2) is satisfied, then
(4.21) |
To this aim, we use (4.17) and (4.18) to observe that
Consequently, using (1.2) and (4.2), we get that
which is precisely (4.21).
Step II. We prove that if and (1.2) is satisfied, then there exists a structural constant and sufficiently small such that
(4.22) |
For this we take
(4.23) |
and we point out that, as a consequence of (4.20), the condition in (4.4) is satisfied. This allows us to use Lemma 4.1, yielding that
Therefore, because of the above choice for and exploiting the Cauchy-Schwarz inequality, we also find that
and thus there exists some such that
(4.24) |
Hence, using (4.21) and (4.24), and recalling (4.18), we deduce that
which gives (4.22).
We notice that, by the very definition of infimum, and recalling the choice of in (4.23), there exists such that
(4.25) |
for such that (1.2) is satisfied.
Step III. We let such that (1.2) is satisfied, and we take such that (4.25) holds true. We prove that there exists a structural constant such that
(4.26) |
For this, recalling that (4.18) and (4.19) hold for every , and using (4.25), we have that
From this, eventually modifying the constant , we further obtain that
(4.27) |
Furthermore, since , recalling (4.20), we have that the condition in (4.4) is satisfied, and therefore we can exploit Lemma 4.1. In this way, we obtain that
thanks to (4.20). Plugging this information into (4.27), we obtain (4.26).
Step IV. We are now ready to finish the proof of (1.3). Once this is established, the existence of a ball for which (1.4) holds follows from the convexity of and Lemma 4.2 (notice that we are in the position of exploiting Lemma 4.2 thanks to (4.20)).
Hence we focus on the proof of (1.3). We let be the inradius of and let us consider the ball whose radius is and such that . We recall that the convexity of the set ensures that the following Bonnesen-type inequality holds, see e.g. [27, 38, 50]:
(4.28) |
Also, from (4.26), we have that
Therefore,
(4.29) |
Furthermore, using the isoperimetric inequality we see that
As a result, we obtain that
(4.30) |
Now, combining (4.28), (4.29) and (4.30), we find that
Recalling that has radius , hence and
it thus follows that
(4.31) |
We also point out that
(4.32) |
for sufficiently small and therefore
up to renaming in the last inequality.
Appendix A Convex sets, remarks on the literature, and optimality of Lemma 4.2
In the literature, it seems to be suggested (see e.g. the end of Section 2 in [49]) that the result in Lemma 4.2 could be improved, for instance by posing the following natural question:
Problem A.1.
Let . If is bounded and convex, contains the unit ball and
(A.1) |
is it true that there exists a ball such that and
(A.2) |
for some constant ?
Here we show that the answer to Problem A.1 is negative. We provide two counterexamples, one closely related to the proof of Lemma 4.2, and one in which we additionally assume that the set is uniformly convex and its boundary is of class .

We claim that (A.1) holds true. Indeed, considering Figure 2, we have that and . As a result, we see that and
We also remark that the triangles , and are similar and accordingly
These identities entail that
and |
As a result, if
we find that .
Now, consider a ball such that . Since
we have that the diameter of is at least , hence the radius of is at least and thus, using (A.1),
This yields that (A.2) is not satisfied in this case, since otherwise
which is a contradiction.
Counterexample 2. For simplicity, we take here (the case can be obtained by rotations of the example constructed in a given plane). Let be small and
We construct here a counterexample of Problem A.1 even under the additional assumption that is uniformly convex with boundary of class . For this, let with and
with
We consider the set
We observe that for all , hence , and
(A.3) | whenever . |
We let be the curvature of at the point . Then,
Hence, if
we have that
(A.4) |
and
From this relation and (A.4), we conclude that
hence is uniformly convex.
Appendix B -regularity for
For completeness, we present here an explicit proof of Theorem 2.8 in the case . In this situation the action of the fractional operator is better behaved since it does not produce boundary singularities on functions that are smooth (or even just Lipschitz) up to the boundary and with zero external datum. This fact makes the proof technically easier since it allows one to “reabsorb” the fractional operator into the source term of the equation. For this reason, we thought it could be of some interest, at least for some readers, to find here a self-contained result with its own proof. The precise statement is the following:
Theorem B.1.
Let and be such that
(B.1) |
Suppose that is of class . If and if denotes the unique weak solution of (according to Theorem 2.4), then
In particular, is a classical solution of .
Proof.
We split the proof into different steps.
Step I. We consider the functions space defined as follows:
Then, we claim that there exists a constant such that
(B.2) |
In fact, since and since, by assumption (B.1), , it is not difficult to recognize that , and
As a consequence, since one obviously has , we are entitled to apply the result in [56, Prop. 2.1.7]: this gives and
(B.3) |
where is a constant independent of . To complete the proof of (B.2), we then turn to estimate the -norm of in terms of .
First of all we observe that, on account of (B.3), for every one has
(B.4) |
where is chosen in such a way that
On the other hand, since and , we have the estimate
(B.5) |
where is another constant which does not depend on . Gathering together (B.3), (B.4) and (B.5), we finally obtain
which is exactly the claimed (B.2). We explicitly point out that the constant only depends on and (as the same is true of ).
Step II. In this second step, we establish the following facts:
-
(1)
for every ;
-
(2)
there exists a constant such that
(B.6)
As regards assertion (1), it is a direct consequence of (B.2), together with the fact that if . We now turn to prove assertion (2).
To this end we first notice that, since is of class , we are entitled to apply [36, Thm. 6.14]: there exists a constant such that
(B.7) |
for every function satisfying on . Then, by combining (B.2) with (B.7), we obtain the following chain of inequalities:
(B.8) |
holding true for every . Now, owing again to the regularity of , we can invoke the global interpolation inequality contained in, e.g., [36, Chap. 6]: there exists a constant , independent of , such that
This, together with (B.8), immediately gives the desired (B.6).
Step III. In this last step, we complete the proof of the theorem by using the so-called method of continuity (see, e.g., [36, Thm. 5.2]). To this end, we first notice that is endowed with a structure of Banach space by the norm
Moreover, for every we define
Owing to (1)-(2) in Step II, we derive that maps into , and
(B.9) |
where is a suitable constant independent of and . On the other hand, by carefully scrutinizing the proof of [7, Thm. 4.7], it is easy to see that
(B.10) |
where is another constant independent of and . Thanks to (B.9)-(B.10), we are then entitled to apply the method of continuity in this setting: indeed,
-
•
and are Banach spaces;
-
•
are linear and bounded from into (see (B.2));
-
•
there exists a constant such that
As a consequence, since is surjective, we deduce that also is surjective: for every there exists a (unique) such that
(B.11) |
We explicitly notice that, since , one has and on ; thus, by (B.11) we derive that is a classical solution of . In view of these facts, to complete the proof we are left to show that
To this end we first notice that, since and since on , we surely have ; this, together with (2.1), implies that
(B.12) |
Since is a classical solution of , from (B.12) and Remark 2.6 we infer that is also a weak solution of problem ; on the other hand, since is the unique weak solution this problem, we conclude that a.e. in , as desired. ∎
Acknowledgments. We would like to thank Professor Lorenzo Brasco for pointing to our attention the paper [2] and for some interesting and pleasant discussions.
References
- [1] N. Abatangelo, M. Cozzi, An elliptic boundary value problem with fractional nonlinearity, SIAM J. Math. Anal. 53(3) (2021), no. 3, 35770–3601.
- [2] A.I Ávila, Stability results for the first eigenvalue of the Laplacian on domains in space forms, J. Math. Anal. Appl. 267 (2002), no. 2, 760–774.
- [3] G. Barles, E. Chasseigne, A. Ciomaga, C. Imbert, Lipschitz regularity of solutions for mixed integro- differential equations, J. Differential Equations 252 (2012), no. 11, 6012–6060.
- [4] G. Barles, E. Chasseigne, A. Ciomaga, C. Imbert, Large time behavior of periodic viscosity solutions for uniformly parabolic integro-differential equations, Calc. Var. Partial Differential Equations 50 (2014), no. 1-2, 283–304.
- [5] G. Barles, C. Imbert, Second-order elliptic integro-differential equations: viscosity solutions’ theory revisited, Ann. Inst. H. Poincaré Anal. Non Linéaire 25 (2008), no. 3, 567–585.
- [6] A. Bensoussan, J.L. Lions, Impulse Control and Quasi-Variational Inequalities, Gauthier-Villars, Paris, 1984.
- [7] S. Biagi, S. Dipierro, E. Valdinoci, E. Vecchi, Mixed local and nonlocal elliptic operators: regularity and maximum principles, Comm. Partial Differential Equations, DOI: 10.1080/03605302.2021.1998908.
- [8] S. Biagi, S. Dipierro, E. Valdinoci, E. Vecchi, Semilinear elliptic equations involving mixed local and nonlocal operators, Proc. Roy. Soc. Edinburgh Sect. A 151 (2021), no. 5, 1611–1641.
- [9] S. Biagi, D. Mugnai, E. Vecchi, Global boundedness and maximum principle for a Brezis-Oswald approach to mixed local and nonlocal operators, Commun. Contemp. Math., DOI:10.1142/S0219199722500572.
- [10] I. H. Biswas, E. R. Jakobsen, K. H. Karlsen, Viscosity solutions for a system of integro–PDEs and connections to optimal switching and control of jump-diffusion processes, Appl. Math. Optim. 62 (2010), no. 1, 47–80.
- [11] D. Blazevski, D. del-Castillo-Negrete, Local and nonlocal anisotropic transport in reversed shear magnetic fields: Shearless Cantori and nondiffusive transport, Phys. Rev. E87 (2013), 063106.
- [12] L. Brasco, G. De Philippis, B. Velichkov, Faber-Krahn inequalities in sharp quantitative form, Duke Math. J. 164 (2015), no. 9, 1777–1831.
- [13] H. Brezis, Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011.
- [14] S. Buccheri, J.V. da Silva, L.H. de Miranda, A System of Local/Nonlocal -Laplacians: The Eigenvalue Problem and Its Asymptotic Limit as , preprint, arXiv: 2001.05985
- [15] X. Cabré, S. Dipierro, E. Valdinoci, The Bernstein technique for integro-differential equations, Arch. Ration. Mech. Anal.
- [16] X. Cabré, J. Serra, An extension problem for sums of fractional Laplacians and -D symmetry of phase transitions, Nonlinear Anal. 137 (2016), 246–265.
- [17] L. Caffarelli, E. Valdinoci, A priori bounds for solutions of a nonlocal evolution PDE, Analysis and numerics of partial differential equations, 141–163, Springer INdAM Ser., 4, Springer, Milan, 2013.
- [18] A. Ciomaga, On the strong maximum principle for second-order nonlinear parabolic integro-differential equations, Adv. Differential Equations 17 (2012), no. 7-8, 635–671.
- [19] R. Courant, Beweis des Satzes, dass von allen homogenen Membranen gegebenen Umfanges und gegebener Spannung die kreisförmige den tiefsten Grundton besitzt, Math. Z. 1 (1918), 321–328.
- [20] G. Cupini, E. Vecchi, Faber-Krahn and Lieb-type inequalities for the composite membrane problem, Commun. Pure Appl. Anal. 18(5) (2019), 2679–2691.
- [21] J.V. da Silva, A.M. Salort, A limiting problem for local/non-local -Laplacians with concave-convex nonlinearities, Z. Angew. Math. Phys. 71 (2020), Paper No. 191, 27pp.
- [22] R. de la Llave, E. Valdinoci, A generalization of Aubry-Mather theory to partial differential equations and pseudo-differential equations, Ann. Inst. H. Poincaré Anal. Non Linéaire 26 (2009), no. 4, 1309–1344.
- [23] F. del Teso, J. Endal, E. R. Jakobsen, On distributional solutions of local and nonlocal problems of porous medium type, C. R. Math. Acad. Sci. Paris 355 (2017), no. 11, 1154–1160.
- [24] F. Dell’Oro, V. Pata, Second order linear evolution equations with general dissipation, Appl. Math. Optim. 83 (2021), no. 3, 1877–1917.
- [25] L. M. Del Pezzo, R. Ferreira, J. D. Rossi, Eigenvalues for a combination between local and nonlocal -Laplacians, Fract. Calc. Appl. Anal. 22 (2019), no. 5, 1414–1436.
- [26] E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker’s guide to the fractional Sobolev spaces, Bull. Sci. Math. 136 (2012), no. 5, 521–573.
- [27] A. Dinghas, Bemerkung zu einer Verschärfung der isoperimetrischen Ungleichung durch H. Hadwiger. Math. Nachr. 1 (1948), 284–286.
- [28] S. Dipierro, M. Medina, I. Peral, E. Valdinoci, Bifurcation results for a fractional elliptic equation with critical exponent in , Manuscripta Math. 153 (2017), no. 1-2, 183–230.
- [29] S. Dipierro, E. Proietti Lippi, E. Valdinoci, Linear theory for a mixed operator with Neumann conditions, Asymptot. Anal., DOI: 10.3233/ASY-211718.
- [30] S. Dipierro, E. Proietti Lippi, E. Valdinoci, (Non)local logistic equations with Neumann conditions, Ann. Inst. H. Poincaré Anal. Non Linéaire, DOI: 10.4171/AIHPC/57.
- [31] 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, Phys. A 575 (2021), 126052.
- [32] S. Dipierro, E. Valdinoci, V. Vespri, Decay estimates for evolutionary equations with fractional time-diffusion, J. Evol. Equ. 19 (2019), no. 2, 435–462.
- [33] G. Faber, Beweis, dass unter allen homogenen Membranen von gleicher Fläche und gleicher Spannung die kreisförmige den tiefsten Grundton gibt, Sitzungsber. Bayer. Akad. Wiss. München, Math.-Phys. Kl. (1923), 169–172.
- [34] R.L. Frank, R. Seiringer, Non-linear ground state representations and sharp Hardy inequalities, J. Funct. Anal. 255 (2008), 3407–3430.
- [35] M.G. Garroni, J.L. Menaldi, Second order elliptic integro-differential problems, Chapman & Hall/CRC Research Notes in Mathematics, 430. Chapman & Hall/CRC, Boca Raton, FL, 2002. xvi+221 pp.
- [36] D. Gilbarg, N.S. Trudinger, Elliptic Partial Differential Equations of Second Order, Springer-Verlag, Berlin, (2011).
- [37] E. Giusti, Direct Methods in the Calculus of Variations, World Scientific Publishing Co., Inc., River Edge, NJ, 2003.
- [38] H. Hadwiger, Die isoperimetrische Ungleichung im Raum, Elem. Math. 3 (1948), 25–38.
- [39] W. Hansen, N. Nadirashvili, Isoperimetric inequalities in potential theory, Potential Anal. 3 (1994), 1–14.
- [40] A. Henrot, Extremum problems for eigenvalues of elliptic operators. Frontiers in Mathematics. BirkhäuserVerlag, Basel, 2006.
- [41] E. R. Jakobsen, K. H. Karlsen, Continuous dependence estimates for viscosity solutions of integro–PDEs, J. Differential Equations 212 (2005), no. 2, 278–318.
- [42] E. R. Jakobsen, K. H. Karlsen, A “maximum principle for semicontinuous functions” applicable to integro-partial differential equations, NoDEA Nonlinear Differential Equations Appl. 13 (2006), 137–165.
- [43] M. Kassmann, L. Silvestre On the superharmonicity of the first eigenfunction of the fractional Laplacian for certain exponents, preprint, http://math.uchicago.edu/luis/preprints/cfe.pdf
- [44] S. Kesavan, Some remarks on a result of Talenti, Ann. Scuola Norm. Sup. Pisa Cl. Sci. 15 (1988), 453–465.
- [45] P. Garain, J. Kinnunen, On the regularity theory for mixed local and nonlocal quasilinear elliptic equations, preprint, arXiv:2102.13365
- [46] D. Goel, K. Sreenadh, On the second eigenvalue of combination between local and nonlocal -Laplacian, Proc. Amer. Math. Soc. 147 (2019), no. 10, 4315–4327.
- [47] E. Krahn, Über eine von Rayleigh formulierte Minimaleigenschaft des Kreises, Math. Ann. 94 (1925), 97–100.
- [48] E. Krahn, Über Minimaleigenschaften der Kugel in drei und mehr Dimensionen, Acta Comm. Univ. Tartu (Dorpat) A9, (1926), 1–44.
- [49] A.D. Melas, The stability of some eigenvalue estimates, J. Differential Geom. 36 (1992), 19–33.
- [50] R. Osserman, Bonnesen-style isoperimetric inequalities, Amer. Math. Monthly 86 (1979), no. 1, 1–29.
- [51] 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), no. 1, 137–167.
- [52] J.W.S. Rayleigh, The theory of sound, London (1894/96) pp. 339–340 (Edition: Second).
- [53] X. Ros-Oton, J. Serra, Nonexistence results for nonlocal equations with critical and supercritical nonlinearities, Comm. Partial Differential Equations 40 (2015), no. 1, 115–133.
- [54] R. Servadei, E. Valdinoci, Variational methods for non-local operators of elliptic type, Discrete Contin. Dyn. Syst. 33 (2013), 2105–2137.
- [55] R. Servadei, E. Valdinoci, Weak and viscosity solutions of the fractional Laplace equation, Publ. Mat. 58 (2014), no. 1, 133–154.
- [56] L.E. Silvestre, Regularity of the obstacle problem for a fractional power of the Laplace operator, Thesis (Ph.D.)–The University of Texas at Austin (2005), 95 pp.
- [57] I.M. Singer, B. Wong, S.-T. Yau, S. S.-T. Yau, An estimate of the gap of the first two eigenvalues in the Schrödinger operator, Ann. Sc. Norm. Super. Pisa Cl. Sci. (4) 12 (1985), 319–333.