Asymptotic expansion
of a
nonlocal phase transition energy
Abstract.
We study the asymptotic behavior of the fractional Allen–Cahn energy functional in bounded domains with prescribed Dirichlet boundary conditions.
When the fractional power , we establish the first-order asymptotic development up to the boundary in the sense of -convergence. In particular, we prove that the first-order term is the nonlocal minimal surface functional. Also, we show that, in general, the second-order term is not properly defined and intermediate orders may have to be taken into account.
For , we focus on the one-dimensional case and we prove that the first-order term is the classical perimeter functional plus a penalization on the boundary. Towards this end, we establish existence of minimizers to a corresponding fractional energy in a half-line, which provides itself a new feature with respect to the existing literature.
Key words and phrases:
Gamma-convergence, nonlocal phase transitions, fractional Allen–Cahn equation2010 Mathematics Subject Classification:
82B26, 35R11, 49J45, 35A01.1. Introduction
1.1. Higher-order expansions and boundary effects
In this paper, we initiate the study of the asymptotic behavior as of the nonlocal phase coexistence model
(1.1) |
where , is a bounded domain, is a double-well potential (see (1.2)), is a small phase parameter, and , take into account the scaling properties of the energy functional in terms of the fractional parameter (see (1.3)). In particular, the effect of the potential energy, which favors the pure phases of the system, is balanced by the kinetic energy that takes into account the long-range particle interactions and is ferromagnetic.
The energy (1.1) was considered in [SV-gamma, SV-dens]. In the former, the authors prove that, under appropriate scaling, the energy -converges to the classical perimeter functional when and to the nonlocal perimeter functional when .
In this work, we initiate the analysis of the higher-order asymptotic behavior as . This is important in the study of slow motion of interfaces in the corresponding evolutionary problem and showcases improved convergence rates for the minimum values. Different than the interior result in [SV-gamma], we also take into account the nonlocal Dirichlet boundary conditions by restricting the class of minimizers to those satisfying a.e. on for a given function (see (1.16)). In this sense, the novelties provided by our work cover two distinct but intertwined aspects: on the one hand, we analyze the effect of the boundary data on the limit energy functional; on the other hand, we provide a higher-order expansion of the limit functional with respect to the perturbation parameter .
Let us now precisely describe our setting mathematically and state our main results.
We let be a bounded domain of class . Roughly speaking, this domain is the “container” in which the phase transition takes place. Notice in (1.1) that the set
collects all the couples for which at least one entry lies in : that is, takes into account all the possible long-range interactions which involve particles inside the container. As is customary, denotes the complement of in .
To write the total energy of the system in a convenient notation, given a measurable function and a measurable set , we write
In particular, if with and measurable subsets of , we write
We assume that is a double-well potential satisfying
(1.2) |
Following [SV-gamma, SV-dens], the scaling in the fractional Allen–Cahn energy (1.1) is given by
(1.3) |
As , the functional in (1.1) recovers the classical Allen–Cahn phase transition energy.
Let be a given Lipschitz continuous function. We define
(1.4) |
We establish a first-order asymptotic expansion of for any fixed .
For , we prove a first-order expansion in any spatial dimension. Morally speaking, we show that approximates the fractional perimeter of a nonlocal minimal surface in with prescribed Dirichlet boundary data as . See the seminal paper [CRS] for more on nonlocal minimal surfaces. The precise result goes as follows.
Theorem 1.1.
Let , and be a bounded domain of class . It holds that
(1.5) |
where
Here, the infimum is taken over measurable sets .
Here above and in what follows, denotes the characteristic function associated to a measurable set .
Notably, for and in dimension , we can find a specific scenario in which does not have a meaningful second-order expansion and, in fact, intermediate fractional powers of could be needed for a meaningful expansion (see Theorem 1.9 and Corollary 1.10 below).
The case is fundamentally different as the long-range interactions are lost in the limit as . In this paper, we provide the first-order asymptotic expansion of in dimension . Since the proof is already quite involved, we leave the general case and the higher-order asymptotic expansion of as future work.
We use the standard notation to denote the (classical) perimeter of a set in an open set , see [Giusti].
For a function such that for some set with , the trace of along , given by
(1.6) |
is well-defined. As an abuse of notation in this setting, we often write in place of , even when .
When , we also take an extra assumption to control the behavior of the functions considered near the boundary. This is a technical assumption that we aim at relaxing in a forthcoming work, but, roughly speaking, it is used in our setting to prevent additional oscillations of minimizing heteroclinics.
This technical hypothesis goes as follows. Suppose that . Let . Consider the class of measurable functions such that, for ,
(1.7) |
We point out that when , the set consists of all the measurable functions .
Consider
(1.8) |
In this setting, the first-order of the asymptotic expansion of the nonlocal phase coexistence functional goes as follows:
Theorem 1.2.
Let , be a bounded interval, and . Suppose that on .
It holds that
where
Here, the infimum is taken over measurable sets such that and the constant depends only on and .
Also, is a suitable function, whose explicit definition will be given in (7.3) relying on the energy of a suitable minimal layer solution.
Roughly speaking, for any and , the penalization functional mentioned in Theorem 1.2 is given by
(1.9) |
where is the class of functions such that , , and in , and similarly for . That is, for is the minimum energy connecting the interior value (in the trace sense) to the boundary value .
Thus, in order to construct the penalization function , we need to construct the heteroclinic function connecting to . This construction is also somewhat delicate and does not follow trivially from the constructions of the heteroclinic function connecting to put forth in [CABSI, PAL]. Indeed, in our setting, cutting methods and monotonicity considerations are not available. Moreover, since is not an equilibrium for the potential , a bespoke analysis is needed.
The case is even more delicate, due to the presence of infinite energy contributions. To overcome this difficulty, the minimal layer solution when will be obtained by a limit procedure from the cases .
We stress that the construction of this minimal heteroclinic connection is an important ingredient for the recovery sequence in the -convergence expansion that we propose and introduces a novelty with respect to the available methods in the nonlocal setting. Indeed, while the interior estimates of the -limits rely on the existing literature (see [SV-gamma]), the control of the boundary terms requires an ad-hoc analysis and an appropriate interpolation that involves this new heteroclinic connection.
We now describe in detail the mathematical setting related to this layer solution. Towards this end, for any , we define
and
where .
Given , we define to be the closure, with respect to the -seminorm, of the set
Loosely speaking, is the family of functions in that connect at to in .
With this notation, we can establish the existence of minimizers to the energy in the class :
Theorem 1.3.
Given and , there exists a unique global minimizer of the energy .
Moreover, is strictly increasing in , for all , for all , and solves
(1.10) |
Furthermore, there exist constants , such that
(1.11) |
and, for all ,
(1.12) |
If , then we additionally have and there exist , such that
(1.13) |
A similar result holds when is replaced by .
We faced several difficulties in proving the existence result in Theorem 1.3. First, if , then minimizers may instead prefer to connect to in the sense that . For instance, due to the boundary condition in , classical sliding methods as used in [PAL] are ineffectual in proving both monotonicity and uniqueness. To solve the issue, we impose the additional constraint in the class . Then, since the energy on the right-hand side of (1.9) is finite for , existence is proved using compactness. However, for , the energy is infinite and must be defined through an appropriate rescaling. In the latter case, we give a novel approach to proving existence by sending . For this, we first prove uniform asymptotics in near for the heteroclinic orbits connecting at to at . See Sections 6 and 7 for full details.
1.2. Asymptotic development
The notion of asymptotic development in the sense of -convergence was introduced by Anzellotti and Baldo in [Baldo1], see also [Baldo2]. We will present the definitions and details surrounding asymptotic development in our setting. For reference, let us first recall the definition of -convergence, see [Baldo1, Definition 1.1].
Definition 1.4.
For a topological space , let be a family of functions with parameter . We say that is the -limit of if the following hold.
Let as .
-
(1)
For any sequence in as , it holds that
-
(2)
Given , there exists a sequence in as such that
In this case, we write in .
Consider the space of all the measurable functions such that the restriction of to belongs to . We endow , as well as its subspaces, with the metric of :
(1.14) |
Notice that the values outside are neglected in this procedure.
We comprise the Dirichlet datum inside the functional by defining
(1.15) |
and
(1.16) |
Notice that is the functional space used in the construction of in (1.4), so that
We also set and .
Following [Baldo1, Baldo2] (see in particular [Baldo1, pages 109–110]), the asymptotic development of order , written as
holds in the sense of -convergence if
-
(1)
in ;
-
(2)
for any , we have that in , where
with and
One can show (see [Baldo1, pages 106 and 110]) that
and
(1.17) |
Therefore, the asymptotic development provides a refined selection criteria for minimizers of .
In this setting, we prove the following asymptotic behavior of in (1.1) in the sense of -convergence.
Theorem 1.5.
Let , and be a bounded domain of class . For all , it holds in the sense of -convergence that
where
For , we recall (1.7) and define
(1.18) |
Here, is used in the construction of in (1.8). In this setting, we prove the following first-order asymptotic of .
Theorem 1.6.
Let , be a bounded interval, and . Suppose that on .
It holds in the sense of -convergence that
where
and
Remark 1.7.
Remark 1.8.
In this framework and recalling (1.17), Theorems 1.1 and 1.2 are a consequence of Theorems 1.5 and 1.6, respectively.
For , notice in Theorems 1.1 and 1.5 that the nonlocal energies with exterior boundary conditions in give rise to a nonlocal energy with exterior boundary conditions. In contrast, for , we see in Theorems 1.2 and 1.6 that the limiting energy localizes both in the interior and at the boundary, in the sense that the penalization energy only sees on .
For and in a specific one-dimensional setting, we prove by direct calculation that does not have a meaningful asymptotic development of order or in fact any non-integer order . We present the precise result here. An analogous statement is expected to hold in a more general setting and is left for future work.
Theorem 1.9.
Let , , , and . Then, there exists a sequence such that in as , and
In particular,
and, for all ,
Corollary 1.10.
In the setting of Theorem 1.9, we have that
Since the inaugural works [Baldo1, Baldo2] there have been several papers devoted to asymptotic development in the local setting. See for instance [Braides] for additional details. See also [MR759767, MR2971613, MR3385247], where a second-order -convergence expansion is produced for a classical, one-dimensional, Modica-Mortola energy functional, and [DalMaso, Leoni], for the higher-dimensional case.
To the best of our knowledge, we are the first to consider asymptotic development in the nonlocal setting. Our main obstacle to overcome is understanding the penalization function and its role at the boundary for .
1.3. Organization of the paper
The rest of the paper is organized as follows. First, in Section 2, we prove -convergence to as described in Remark 2.2. Then, we set notation for first-order -convergence in Section 3. The proof of Theorem 1.5 and a discussion surrounding Theorem 1.9 for is in Section 4. After that, we will assume for the remainder of the paper that and set notation in Section 5. Background and preliminaries on heteroclinic connections are provided in Section 6. Section 7 contains the proof of Theorem 1.3 and the construction of the penalization function . To prove Theorem 1.6 for , we establish the -inequality in Section 8 (see Proposition 8.3) and the -inequality in Section 9 (see Proposition 9.10). Lastly, we collect some auxiliary energy estimates in Appendix A.
2. Computation of
In this section, we will establish the zero-th order term in Theorems 1.5 and 1.6 in a general setting. Let be as in (1.15) and as in (1.16) for all (i.e. for ).
Lemma 2.1.
Let , and be a bounded domain of class . It holds that where
Proof.
Take a sequence .
We first show that if is a sequence in with as , then
(2.1) |
We recall that the notion of convergence in is the one in (1.14).
Notice that if
the claim in (2.1) is obvious. Hence, we suppose that
In this case, we take a subsequence realizing the above limit.
Moreover, we recall (1.1) and (1.16) and we observe that
We also point out that, as , we have that when and when , that is
(2.2) |
Furthermore, by (1.14), we take a further subsequence, still denoted by , that converges to a.e. in . Hence, by Fatou’s Lemma,
which proves (2.1).
Now, to complete the proof of Lemma 2.1, we show that for every there exists a sequence which converges to as and such that
(2.3) |
To construct such a recovery sequence, we perform several surgeries, such as truncations, mollifications and cutoffs (roughly speaking one smooths a bit the function and glues it to smoothly near the boundary). The arguments are, in a sense, of elementary nature, but they require some delicate quantifications.
For this, we fix . For any , we define
Notice that a.e. in as and . Therefore, by the Dominated Convergence Theorem, we have that
In particular, we can find such that
(2.4) |
Now, for any , we define to be the set of all the points of which are at distance larger than from . We set .
Notice that
(2.5) |
and, again by the Dominated Convergence Theorem, we have that
Therefore, we can find sufficiently small such that
(2.6) |
and also that
(2.7) |
Now we perform a mollification argument. Let . For any , we define
By construction, and
(2.8) |
In addition, by (2.5),
(2.9) |
We also claim that, if ,
(2.10) | is supported in . |
To check this, suppose that , that is
This implies that there exists such that
In particular, we have that . Accordingly, for any , we have that . So, we find that
which says that , thus establishing (2.10).
Now we perform a cutoff argument. Namely, we take , with in and we set
By construction,
(2.12) | is a Lipschitz function, |
which coincides with outside , and such that
In particular,
(2.13) |
Furthermore, by (2.9), we have that, in ,
Therefore, by (2.7),
Combining this with (2.4), (2.6) and (2.11), we find that
(2.14) |
Now we remark that, for any ,
(2.15) |
On the other hand, for any fixed , we have that , thanks to (2.12). From these observations, we conclude that there exists such that, for any ,
(2.16) |
Now we define
We observe that and that is strictly increasing. So, by a linear interpolation, we can extend to a piecewise linear function on , which is strictly increasing and therefore invertible on its image, with strictly increasing.
As a consequence, the inequality in (2.16) holds true for every , that is, equivalently, for every . Hence, in particular, if we set
we have that
(2.17) |
Now we observe that
(2.18) |
Indeed, suppose not. Then, since is increasing, the limit above exists and
In particular, there exists such that for any we have that . As a consequence, for any ,
and so
Sending , we obtain a contradiction and so (2.18) is proved.
3. Set up for first-order expansion
4. Asymptotic development for and proof of Theorem 1.5
This section is devoted to the proofs of Theorem 1.5 and Theorem 1.9. We assume throughout that and (unless otherwise stated) are fixed.
4.1. Computation of
Here, we establish the first-order term in Theorem 1.5.
Lemma 4.1.
Let . It holds that where is given by
Proof.
We take a sequence .
First, we show that if is a sequence in with in as , then
(4.1) |
To this aim, we may assume that
(4.2) |
otherwise (recall (3.1)) and we are done.
In addition, we may suppose that
(4.3) | a.e. in , |
since, if not,
and so, by Fatou’s Lemma,
Accordingly, by (4.3) we know that for some . By (4.2) and Fatou’s Lemma,
This proves (4.1).
Now we show that for every there exists a sequence which converges to as and such that
(4.4) |
For this, we may suppose that for some . Otherwise , and we are done. In particular, we have that
So, we define
Then, and we have that
This proves (4.4).
4.2. Proof of Theorem 1.5
4.3. Computation for
The rest of this section is devoted to the second-order asymptotic development when . In light of Lemma 4.1, we have that
The set of minimizers is denoted by
and the functionals are given by
Now we notice that -converges to in .
Lemma 4.2.
Assume , and fix . Let be such that in and as .
Then
Proof.
As in (4.2), we may assume that . Suppose first that for some measurable . By Fatou’s Lemma and the definition of ,
Consequently,
Now, suppose that . In this case,
By Fatou’s Lemma,
and consequently,
This completes the proof. ∎
We now consider -convergence of in in dimension . In particular, we establish Theorem 1.9. The proof is broken up into several small observations.
Lemma 4.3.
Let and . Let and . Let also be the minimizer of among measurable functions with in .
Then,
and
where
Proof.
Given with in , we define and
Since vanishes identically in and for all , we have that
From this, we obtain the desired result (we stress that means with ). ∎
Lemma 4.4.
In the notation of Lemma 4.3, we have that .
Proof.
Due to the minimality of , we know that . Also, a minimizer satisfies the Euler-Lagrange equation in and therefore it is necessarily continuous at the origin, ruling out the possibility for to attain the minimum. ∎
Lemma 4.5.
In the notation of Lemma 4.3, we have that is a minimizer of among measurable functions with in and in .
Proof.
We know that a minimizer can be taken to be monotone (see [Alberti, Theorem 2.11]) and odd symmetric (see [MR3596708, Lemma A.1]). Since the minimizer is constrained to take values in , the desired result follows. ∎
Proof.
Corollary 4.7.
Proof.
We consider the function
and we observe that which is positive if and only if and the desired result follows. ∎
5. Notation for
Throughout the remainder of the paper, we assume that .
We use the following notation for measurable sets , :
(5.1) | ||||
We recall that and are defined in (3.2). We point out that in contrast to , the energy only depends on the values of in .
Also, let , and assume that , are such that in . It is a straightforward computation to show that
(5.2) |
Lastly, we note how the energies in (5.1) scale. First, if for , then it is easy to check that
(5.3) |
and similarly for . In the special case (i.e. ), we have
(5.4) |
6. Heteroclinic connections
In this section, we give background and preliminaries on heteroclinic connections that connect at to at . As proved in [SV-gamma], these are used to construct the recovery sequence in the interior (i.e. in compact subsets ) for -convergence. We will also use the heteroclinic connections as barrier when studying connections at the boundary.
Recalling (5.1), we set
(6.1) |
Existence and uniqueness (up to translations) of minimizers of over the class of -functions that connect at and at was established in [PAL, CABSI]. After fixing the value of the minimizer at the origin, the authors also show that the unique minimizer is in the class and satisfies
(6.2) |
Moreover, by [PAL, Theorem 2], for any , there exist constants , such that
(6.3) |
6.1. Limiting behavior near
We will need the following results on the limiting behavior of the minimizers as .
For clarity, let denote the minimizer of given in (6.2). Since , as a consequence of [CS, Theorem 26], there exist and such that, for all ,
(6.4) |
Fix now such that
(6.5) |
We see that the asymptotic behavior of in (6.3) is uniform in .
Lemma 6.1.
There exist constants , such that, for all ,
The proof relies on the barrier constructed in [SV-dens]. More precisely, one can check that the following version, in which all constants are uniformly bounded in in compact subsets of , holds.
Lemma 6.2 (Lemma 3.1 in [SV-dens]).
Let . Fix and .
There exists such that, for any and all , there exists a rotationally symmetric function
with
such that
and
(6.6) |
for all .
Several times in the paper, we will use for that, for ,
(6.7) |
Proof of Lemma 6.1.
We will prove that there exist , such that, for all ,
(6.8) |
The uniform decay at is proved similarly.
Fix small and . Let be such that .
We claim that
(6.9) | can be bounded from above by some , independent of . |
In the following, denotes a positive constant, independent of .
We may assume that , otherwise we are done. For ease in notation, we write and consider the sets and . Let be such that
Note that the Lipschitz constant of in can be taken independently of . Define the function
Since in and is a minimizer of , it holds that
(6.10) |
Now, we notice that in , and therefore there exists such that
On the other hand, since in , there exists such that
Therefore, from (6.10), we have that
(6.11) |
We next show that there exists such that
(6.12) |
For this, for any , , we use the fact that
to note that
From this and (6.4), and since is Lipschitz in , for any ,
Recalling (6.5), it follows for all that
(6.13) |
Next, we check the interactions in by first writing
Since and with (6.7), we estimate, for ,
Taking gives the same estimate.
On the other hand, we use (6.14) and estimate as in (6.13) for all to find
Combining the previous two displays gives
(6.17) |
By (6.11) and (6.12), there exists such that, for all ,
Consequently, is bounded uniformly in , thus giving the claim in (6.9).
We are now ready to show (6.8). The proof follows the lines of [PAL, Proposition 3] except that we have to carefully track the dependence on . For the sake of the reader, we sketch the idea.
First, note that there exists some such that
For this particular and for , let be the barrier in Lemma 6.2 and take . As in the proof of [PAL, Proposition 3] (note that all the constants in [PAL, Corollary 4 and Lemma 9] can be made uniform in in compact subsets ), there exist , (possibly depending on ) such that
Moreover, one can show that there exists such that, for all ,
Recall from above that is such that and that there exists some such that for all . Since is strictly increasing, it must be that and
Corollary 6.3.
There exists a subsequence such that
Proof.
By (6.4), there exist a subsequence and a measurable function such that
Note that is non-decreasing, , and solves
7. Connections at the boundary and proof of Theorem 1.3
We recall the setting of stated on page 1.2 and define the functional
(7.1) |
Let denote the minimizer of in the class (whose definition is obtained replacing with in the definition of ) and
(7.2) | denote the minimizer in the class described in Theorem 1.3. |
The penalization function is then defined as
(7.3) |
We are left to prove Theorem 1.3 for for a fixed . The proof is split into several lemmata. We first establish properties of non-decreasing solutions to the PDE in (1.10). Then, we prove existence of minimizers for and show that they satisfy (1.10). By sending , we prove existence of minimizers for . Next, we collectively show asymptotic behavior at for all and finally prove uniqueness.
For clarity, let us state the definition of local and global minimizers in our setting.
Definition 7.1.
We say that a function is a local minimizer of in if in and
We say that is a global minimizer of if is a local minimizer of in for all and if .
7.1. Properties of solutions
For later reference, we start with the following properties of solutions to the PDE in (1.10).
Lemma 7.2.
Let and . Let be a non-decreasing solution to (1.10) such that for all .
It holds that for any , is strictly increasing in , and .
If , then additionally .
Proof.
The interior regularity (and for ) follows from [ROSerra, Theorem 1.1] and the global regularity follows from [RosOtonBook, Proposition 2.6.15].
Now, we show that . For this, suppose by contradiction that there exist a point such that . Since , it must be that , so we have
Since , it follows that in , a contradiction.
Lastly, we show that is strictly increasing in . Suppose, by way of contradiction, that there exist such that . Observe that
Therefore,
Since is non-decreasing and , the integrand is nonnegative. Consequently, for all , which is false (for example take ). Therefore, is strictly increasing in . ∎
7.2. Existence of solutions for
Here, we prove existence of global minimizers in Theorem 1.3 for and then show that, after translating, the heteroclinic function in Section 6 is above and touches at the origin (and therefore it can be used as a barrier).
Proposition 7.3.
Let and .
There exists a global minimizer of such that , is strictly increasing in , for all , and solves (1.10).
Proof.
By Lemma A.2, there exists a function such that for some . Consequently,
(7.4) |
Consider the set of functions
We claim that
(7.5) |
Since , we have that . For the reverse inequality, first note that is decreasing under truncations at , so it is not restrictive to minimize among all functions satisfying such that .
It remains to check that
(7.6) | is decreasing under monotone rearrangements. |
This is a consequence of [Alberti, Theorem 2.11] (see also [AlmgrenLiem, Theorem 9.2] for symmetric decreasing rearrangements). Indeed, consider the superlevel sets of
We claim that
(7.7) |
For this, note that
If , then
which gives a contradiction. This establishes (7.7).
We can now apply [Alberti, Theorem 2.11] to find that the increasing rearrangement of satisfies , and therefore (7.6) holds true.
Consequently, , and so the proof of (7.5) is complete.
In light of (7.5), we now let be a minimizing sequence for , namely
As a consequence of [Hitchhikers, Theorem 7.1], up to a subsequence, there exists a measurable function such that almost everywhere as . By construction, in , , and is non-decreasing. Also, by Fatou’s Lemma,
(7.8) |
which shows that is a (global) minimizer of .
Furthermore, since is non-decreasing and bounded, the limit
exists. If , then
which contradicts (7.8). Therefore, and we have that .
We will now use the solution to (6.2) as a barrier for . Towards this end, let be the unique point at which and set
Via the sliding method (see [PAL]), we show the following.
Lemma 7.4.
Let and .
It holds that
(7.9) |
Proof.
First, since in and in , we have that, for all , there exists such that
Take as large as possible so that
(7.10) |
and, for all , there exist and such that and
(7.11) |
We claim that
(7.12) | is a bounded sequence in . |
Indeed, if there existed a subsequence such that , then from (7.11) and the continuity of and ,
a contradiction. If instead there existed a subsequence such that , then, similarly,
a contradiction. Hence, (7.12) is proved.
Thus, thanks to (7.12), up to a subsequence, there exists a point such that
Therefore, taking the limit in in (7.11) and recalling (7.10), we have that
(7.13) |
Now, if for some , by the strict monotonicity of , (7.10) and (7.13),
a contradiction. Therefore, for all .
Now, suppose that there exists a subsequence such that for all . By (7.13) and since , we have that, for sufficiently small,
Since exists and is continuous, we have that
Therefore, taking the limit in (7.10) and (7.13) as gives (7.9) in this case.
Consider now the scenario in which there exists such that for all . Define the function
Note that, in light of (7.10) and (7.13), for all and . Also, for all ,
Consequently, since we have the equation for at ,
(7.14) |
We point out that
(7.15) | is also bounded from below. |
Indeed, if there existed a subsequence such that , then
(7.16) |
Since , there exists such that
(7.17) |
By (7.16), there exists such that, for all , it holds that . Therefore, we can employ (7.17) and find that
(7.18) |
Thanks to (7.15), up to a subsequence, there exists a point such that
We claim that is a bounded sequence in . Indeed, if there existed a subsequence such that , then from (7.13)
a contradiction. If instead there existed a subsequence such that , then
a contradiction.
Therefore is bounded, and so, up to a subsequence, there exists such that
We thus send in (7.10) and (7.13) to obtain
(7.19) |
Accordingly, the desired result in (7.9) is proved if we show that
(7.20) |
Hence, we now focus on the proof of (7.20).
Suppose by contradiction that . Sending in (7.14) gives
From (7.19), we know that the integrand is nonnegative. Therefore, it must be that for all , which is false for . Therefore, we have that .
Moreover, since is invertible and , we have that , which completes the proof of (7.20), as desired. ∎
7.3. Existence of solutions for
Now, we establish existence of global minimizers in Theorem 1.3 for . For this, we will show that a subsequence in of the minimizers given by Proposition 7.3 converges as to a minimizer for .
For , let denote the global minimizer of from Proposition 7.3.
Lemma 7.5.
Let . There exist and such that, for all ,
(7.21) |
Furthermore, there exists a subsequence that converges locally uniformly as to a non-decreasing function such that in and (7.9) holds for .
Proof.
Since , to prove (7.21), we only need to check the uniform bound on the -seminorm. By Proposition 7.3 and [ServadeiValdinoci, Theorem 1], is also a viscosity solution to
Note that there exists such that, for all ,
(7.22) |
Consider a ball . By [CS, Theorem 26] with (7.22), there exist and such that, for every ,
On the other hand, by [ROSduke, Proposition 1.1], there exist and such that, for every ,
The previous two displays yield (7.21).
By (7.21), there exists a function such that, up to a subsequence,
Moreover, , in , and is non-decreasing.
For all , let denote the corresponding solution to (6.2). Let be such that and set . In particular, solves
(7.23) |
As in Corollary 6.3, we can show that, up to taking another subsequence,
(7.24) |
Note here that
(7.25) | the sequence is bounded uniformly in . |
Indeed, if there existed a subsequence such that as , then, by Lemma 6.1,
On the other hand, since for all ,
thus giving the desired contradiction and proving (7.25).
From (7.25) we deduce that as . Moreover, we claim that
(7.26) |
To check this, we observe that
Hence, taking the limit as and using the continuity of and the uniform convergence in (7.24), we obtain (7.26).
It remains to check that (7.9) holds for . To this end, we recall that, for , Lemma 7.4 can be written as
(7.27) |
Moreover, for all ,
Thus, using the uniform convergence in (7.24) and the continuity of , we obtain that, for all ,
Recalling (7.26), this gives that, for all ,
We can therefore pass to the limit the inequalities in (7.27) and obtain that
Proposition 7.6.
The function in Lemma 7.5 is a global minimizer of .
Furthermore, for all , is strictly increasing in , for all , and solves (1.10) for .
Proof.
We first check that
(7.28) | is a local minimizer of . |
For this, fix and let be a measurable function with and such that . We will show that
(7.29) |
It is enough to consider the case in which is smooth with . Since is a local minimizer of for ,
(7.30) |
By writing out the expression for the energies, we see that (7.30) is equivalent to
(7.31) |
By the Dominated Convergence Theorem, since is smooth and bounded,
and since and, for close to , ,
Moreover, thanks to (7.21), there exists some , independent of , such that, for close to ,
and
Therefore, by the Dominated Convergence Theorem,
and
Collecting the above limits, we take the limit as of (7.30) written as (7.31) and obtain that (7.29) holds. Namely, is a local minimizer, thus completing the proof of (7.28).
We next show that
(7.32) | is a global minimizer of . |
In light of (7.28), in order to establish (7.32) we only need to check that
For this, let be as in Lemma A.2. Fix (with as given in (6.5)). Notice that the function belongs to (recall the definition of on page 1.2) and therefore, since is a minimizer of in , we have that
(7.33) |
Let now . We point out that if is such that , then one can check that
(7.34) |
Using this with , since in and in , we have that
Therefore, exploiting (7.34) with and recalling also (7.33),
With this and Lemma A.2, we have that
Therefore, with Fatou’s Lemma,
and we have that
as desired. Hence, is a global minimizer of and (7.32) is thereby established.
Finally, we check the remaining properties in the statement of Proposition 7.6. Recall in (7.23) and that, by Lemma 7.5, the estimate (7.9) holds. Since is strictly decreasing, this implies that in and also
With this and the minimization property, solves (1.10) for . The remaining properties follow from Lemma 7.2. ∎
7.4. Asymptotics
We now prove the asymptotic behavior at of global minimizers.
Lemma 7.7.
Proof.
We begin by proving (1.11). By (6.3) together with (7.9) (which holds also for , recall Lemma 7.5), there exist , such that
In particular, (1.11) holds with some , depending on .
Let . We prove the first estimate in (1.13) for . For this, let , be given by (1.11), fix and set . Let . One can readily check that satisfies
By the interior regularity theory for the fractional Laplacian, see for example [ROSerra],
(7.35) |
Since , we have that
(7.36) |
Moreover,
(7.37) |
Combining (7.35), (7.36), and (7.37), and using that , up to renaming ,
In particular,
(7.38) |
and therefore the first estimate in (1.13) holds for with in place of .
7.5. Uniqueness
Finally, we show that minimizers in are unique.
Lemma 7.8.
Let and . The global minimizer of is unique.
Proof.
Let be the minimizer found in Propositions 7.3 and 7.6, and let be another minimizer. Since cutting at level decreases the energy, we have that in .
Set . Since and solve
we have that
where since .
Since , , we have that . Moreover, since ,
Fix . By [RosOtonWeidner, Corollary 6.3], there exists , depending only on , such that
Sending , we find that which implies that is constant in . Therefore,
for some . Since , , it must be that . Consequently, and thus the minimizer is unique. ∎
7.6. Proof of Theorem 1.3
With the work done so far, we can now complete the proof of Theorem 1.3.
8. Bounding the energy from below for
Here, we establish the inequality for the first-order convergence in Theorem 1.6.
8.1. Interpolation near the boundary
We begin by proving an adaptation of [SV-gamma, Proposition 4.1] near . For and , consider . Set
and, for small , define
where denotes the distance from a point to .
Recall also the notation (5.1).
Proposition 8.1.
Fix small and . Let as , and let be a sequence in and be a sequence in such that
Then, there exists a sequence such that
and
Proof.
For ease in the proof, we assume that
and take . The general setting follows along the same lines, and the proof for is similar (see [SV-gamma, Proposition 4.1]). Furthermore, we drop the notation for the subscript .
Also, we may assume that there exists some such that
(8.1) |
otherwise there is nothing to show. Since
(8.2) |
we obtain from (8.1) that
(8.3) |
The rest of the proof is broken into four steps. First, we will partition the set into a finite sequence of intervals so that one of the intervals satisfies an estimate similar to (8.3) with right-hand side small for all . Then, we perform a second partition to find a subinterval such that the difference between and is sufficiently small there. Next, we use the first two steps to appropriately partition and construct the desired function . Lastly, we estimate errors.
Step 1. (Partition ). Fix small. For , to be determined, set
We partition into disjoint intervals , so that
From (8.3), and since , we have that
For sufficiently large, there exists some such that
Fix such a and set
Since , we have that
(8.4) |
Step 2. (Partition ). Let be the integer part of and let be the left endpoint of . For , we define
and observe that .
Denote by
(8.5) |
the distance from to the boundary of . As in the proof of [SV-gamma, Proposition 4.1], we can show that there exists some such that
(8.6) |
From now on such an is fixed once and for all.
Step 3. (Partition and construct ). We partition into the following six disjoint regions:
Note that
(8.7) |
Let be a smooth cutoff function such that on , on , and . Define by
By construction, it is clear that in , in and in . It remains to show that satisfies the energy estimate in Proposition 8.1.
Starting with the kinetic energy, we will show that
(8.8) |
Since , one can check (see [SV-gamma]) that
With this and noting that
in order to prove (8.8), it is enough to show that
(8.9) |
Step 4. (Error estimates). We will show that there exists such that
(8.11) |
Notice that (8.11), together with (8.10), gives the desired result in (8.9) (and then in (8.8)). Hence, we now focus on the proof of (8.11).
First note that
and, since ,
Therefore, (8.4) implies that
(8.12) |
Following the proof of [SV-gamma, Proposition 4.1], we use (8.4), (8.6), and (8.12) to find that
(8.13) |
We are left to show that
For this, observe that
Therefore, recalling that and for ,
Since is Lipschitz continuous and , ,
Therefore,
8.2. Contribution near the boundary
We now establish an energy estimate near .
Proposition 8.2.
Assume that on and let . Let be such that
(8.15) | |||||
(8.16) | and | either or , for all , |
for some and some .
Then,
Proof.
Without loss of generality, we assume that
and that
(8.17) |
Consequently, we deduce from (8.16) that
(8.18) |
We then want to show that
For this, we set
We take a sequence which is infinitesimal as and such that
(8.19) |
Then, for any fixed , we take a sequence which is infinitesimal as and such that
(8.20) |
We point out that, up to extracting a subsequence, we can suppose that
(8.21) |
Since we need to extract subsequences in a delicate and appropriate way, given (to be taken large in the sequel), we use (8.19) to find such that if then
Then, fixed , we use (8.20) to find such that for any we have that
In particular, if and ,
(8.22) |
Now, we define
(8.23) |
where is given by Theorem 1.3 with (recall the notation in (7.2)). In particular,
(8.24) | if , then . |
Since uniformly as , by the monotonicity of , we have that
Hence, for any , , there exists such that
(8.25) |
Now, let be such that for all . If , we have that
(8.26) |
so, by (8.15),
Since is Lipschitz continuous, there exists such that, for any and any ,
Therefore, if , we have that, for any , any and any ,
(8.27) |
Now we define
(8.28) |
Notice that both and diverge as . We also set
Note that is infinitesimal as , thanks to (8.21).
From (8.27), we have that that in , as . An obvious consequence of (8.24) is that also in , as . From (8.25), we have that in , as . On the other hand, if , then by (8.26) and (8.28),
and so, thanks to (8.17),
That is, for any and any , there exists such that, for any and any ,
In particular, taking large enough to guarantee that , it holds that
and as . Therefore, by collecting these pieces, we find that a.e. in , as .
Without loss of generality, let us take now take . Fix and set
By (8.18), we have that in . We are now in a position to apply Proposition 8.1 to guarantee the existence of such that
and satisfying
(8.29) |
where . We recall that is infinitesimal as .
Let . Since and outside , in particular outside , we are in the position of applying (5.2) (with and ), obtaining that
Thus, rescaling as in (5.4) (with ) and using the fact that is a global minimizer of , we find that
Consequently with (8.29) and scaling as in (5.3)-(5.4), we have that
Since is arbitrary, we obtain, recalling also (7.1) and (7.3), that
(8.30) |
For the case , using here that is a local minimizer in , we similarly estimate (since in ),
so that
Consequently,
(8.31) |
8.3. Proof of the inequality
We recall that is the space of all the measurable functions such that the restriction of to belongs to . Moreover, is endowed with the metric of , as made clear in (1.14). Also, the spaces and are defined in (1.15) and (1.7), respectively.
Proposition 8.3.
Assume that on and let . Let be such that in and as .
Then,
where is given by
Proof.
Notice that we can assume that , otherwise (recall the setting in Section 3) and we are done.
Also, in light of [SV-gamma, Proposition 3.3], we can assume that a.e. in , with of finite perimeter. In particular, can be defined along in the trace sense (see e.g. [Giusti]). Moreover, since and in , it must be that .
For clarity in the proof, we assume that . Consider a covering of small intervals of , with infinitesimal as , and with
(8.33) | either or in , with . |
We can satisfy (8.33) since has finite perimeter.
Now we fix an interval . Then, for large , the balls lie outside , and, by inspection, one sees that
(8.34) |
By [SV-gamma, Proposition 4.5], it holds that
With this and Proposition 8.2, we find in (8.34) that
We remark that Proposition 8.2 can be used in this framework, since (8.15) and (8.16) are consequences of the fact that , while (8.15) follows from (8.33). The desired result now follows by taking arbitrarily close to and noticing that
9. Bounding the energy from above for
In this section, we establish the -inequality in Theorem 1.6 by constructing a recovery sequence corresponding to some infinitesimal sequence . To this aim, we let for short and we take which is infinitesimal as and such that, recalling (3.2),
(9.1) |
The quantity will play a crucial role in the detection of the recovery sequence near the boundary of . We will use the notation . Notice that as , thanks to (9.1). We also denote by quantities that are infinitesimal as .
Now, let be such that
(9.2) |
and let denote the signed distance to to with the convention that in . Then, by [SV-gamma, Proposition 4.6], for any , it holds that
(9.3) |
and depends only on and . Here above and in the rest of this section, is the unique solution of (6.2).
We denote by the signed distance from , with the convention that in . Set to be the projection along the boundary of . In particular, we have that
Also, we set
We fix sufficiently small so that . Moreover, since (9.2) holds, up to taking smaller, we can suppose that
(9.4) | does not change sign in . |
For clarity and ease in notation, we also define the trace of in by setting
Finally, for sufficiently large (so that ), we define the function by
(9.5) |
where is as given by (9.3) and
Recall the notation in (7.2) for .
We will show that in (9.5) is the so-called recovery sequence.
Let be fixed. Moreover, for clarity, assume that
(9.6) |
and note that is a finite disjoint covering of for . We also have for sufficiently large that
where
Recalling the definition of the set in (1.7), we assume that (or equivalently the set ) satisfies, for ,
Consequently, .
We now estimate the kinetic and potential energies for separately.
9.1. Kinetic energy estimates
We first estimate the kinetic energy for as
(9.7) | |||
(9.8) | |||
(9.9) | |||
(9.10) | |||
(9.11) | |||
(9.12) |
Under the scaling (3.2), we will use (9.3) (with the corresponding potential energy in ) to estimate the quantity in line (9.8). We then show that the boundary energy arises in-part from lines (9.9) and (9.10), and that the remaining terms (9.11) and (9.12) vanish as .
First, we estimate from above the kinetic energy contributions of near the boundary of , coming from inside . To this end, for , we define
(9.13) |
where, recalling (1.3), we set
(9.14) |
This quantity is well-defined for a fixed and, since is a global minimizer of , recalling the notation in (7.3),
Lemma 9.1.
If , then
Proof.
Without loss of generality, assume that
(9.15) |
Consequently,
(9.16) |
For ease, set
(9.17) |
With this and recalling (9.4) and (9.5), we use the change of variables and to find that
Recall that . Thus, when , we have that and the lemma holds with equality.
If instead , recall that . Notice that, for large,
(9.18) |
Therefore,
as desired. ∎
Now we address the counterpart of Lemma 9.1, in which we estimate from above the mixed energy contributions of near the boundary of , coming from the interactions between and its complement. To this aim, we set
(9.19) |
The reader may compare (9.13) and (9.19) to appreciate the different interactions coming from inside the domain and the ones coming from the inside/outside relations. Note that
Lemma 9.2.
If , then
Proof.
As in the proof of Lemma 9.1, assume, without loss of generality, that (9.15), (9.16), and (9.17) hold. We recall (9.5) to find that
Consequently,
Now, we use the following elementary inequality: if , , then
so that
(9.20) |
Hence, we find that
for sufficiently large and for some .
This gives that
where we use the substitutions and .
For , we have that
as desired.
It remains to estimate the error terms in lines (9.11) and (9.12). For this, we first prove the following regularity estimate away from .
Lemma 9.3.
Let . If , , then, for any , there exists a constant such that
Moreover, if , , there exists such that
Proof of Lemma 9.3.
Recall the notation in (9.6) and notice that, without loss of generality, we can assume that .
In this setting, we write that for all . In light of (9.4), we also have that for all .
Assume also that and set .
We now break the proof into cases based on the definition of and , according to (9.5).
Case 2. Suppose that . Since , , we write
We use (1.12) to estimate
Also, as a consequence of (9.4), we have that for for sufficiently large. Hence, by (6.3), we have that
With this, we obtain that
(9.21) |
Therefore,
Case 4. Suppose that . In this case, and , so we write
Recalling (1.12) and using that , we find that
Case 5. Suppose that . This is the case in which and . Estimating as in (9.21) and using that , we obtain that
Case 6. Suppose that and . This case is trivial since :
We have exhausted all cases and the proof of Lemma 9.3 is complete. ∎
Lemma 9.5.
It holds that
(9.22) |
and
(9.23) |
Proof.
For any , we use that to note that
(9.24) |
With this, we first estimate
and similarly find that . This proves (9.22).
Regarding (9.23), we first estimate the long-range interactions in the same way to find that
and similarly find that .
It is left to show that
We only prove the estimate around since the estimate near is similar. For this, we assume, without loss of generality, that , so that
We thus want to prove that
(9.25) |
For this, we split the remaining error around as
(9.26) |
Using again (9.24), we see that
Recalling (9.7), we combine Lemmata 9.1, 9.2, and 9.5 to obtain following estimate for the kinetic part of the energy:
Lemma 9.6.
It holds that
9.2. Potential energy estimates
We now estimate the potential energy. For this, recall that and write
Near the boundary , we have the following.
Lemma 9.7.
Proof.
We now check the error terms.
Lemma 9.8.
It holds that
Proof.
Combing the previous two lemmata, we have the following:
Lemma 9.9.
9.3. Proof of inequality
We are finally prepared to prove the inequality in Theorem 1.6. Let be as in (7.3). Recalling (9.13) and (9.19), notice that
(9.29) |
We recall that is the space of all the measurable functions such that the restriction of to belongs to . Moreover, is endowed with the metric of , as made clear in (1.14). Also, the space is defined in (1.7).
Proposition 9.10.
Assume that on . Let and be a measurable set.
Then, for any sequence , there exists a sequence such that in and
We stress that Proposition 9.10 holds also for .
9.4. Proof of Theorem 1.6
We can now complete the proof of Theorem 1.6.
Appendix A Some auxiliary results
Here, we collect some auxiliary results that are used in Section 7.
First, we show that minimizers of in (7.1) must be strictly less than in a left-sided neighborhood of the origin.
Lemma A.1.
Let and . Suppose that is such that . Assume that in for some and define
Then, .
Proof.
By a change of variables and using that in , we have that
and
Therefore, the kinetic energy for satisfies
On the other hand, since , we have that . In particular, , so that
The result follows from the previous two displays. ∎
Next, we present a useful competitor for energy estimates in Section 7.
Lemma A.2.
Let and as in (7.1). Define the function
Then, there exists a constant such that, for all ,
(A.1) |
Moreover, there exists a constant such that, for all ,
(A.2) |
Proof.
To start, we calculate some integrals that will be used to prove both (A.1) and (A.2). In the following, is an arbitrary constant, independent of both and .
Observe that
(A.3) | ||||
and
(A.4) | ||||
Next, we find
(A.5) | ||||
Finally, we estimate
(A.6) | ||||
We now prove (A.1). We begin with the interactions in . Using (A.3), (A.4), and (A.5), we find that
Since
we arrive at
Next, we consider the interactions in . For this, we use (A.6) to estimate
Lastly, for the potential energy, we simply notice that
Combining the last three displays gives (A.1).
Next, we prove (A.2). Beginning with the interactions in the container , we use (A.3), (A.4), and (A.5) to find
(A.7) |
Observe that, exploiting (6.7) (with and and with and ),
This and (A.7) give that
(A.8) |
Acknowledgments
It is a pleasure to thank Hidde Schönberger for fruitful discussions. SD has been supported by the Australian Future Fellowship FT230100333 “New perspectives on nonlocal equations”. SP has been supported by the NSF Grant DMS-2155156 “Nonlinear PDE methods in the study of interphases.” EV and MV have been supported by the Australian Laureate Fellowship FL190100081 “Minimal surfaces, free boundaries and partial differential equations.”