Qualitative properties of positive solutions
of a mixed order nonlinear Schrödinger equation
Abstract.
In this paper, we deal with the following mixed local/nonlocal Schrödinger equation
where , and .
The existence of positive solutions for the above problem is proved, relying on some new regularity results. In addition, we study the power-type decay and the radial symmetry properties of such solutions.
The methods make use also of some basic properties of the heat kernel and the Bessel kernel associated with the operator : in this context, we provide self-contained proofs of these results based on Fourier analysis techniques.
Key words and phrases:
Mixed order operators, regularity theory, power-type decay, heat kernel.2020 Mathematics Subject Classification:
35A08, 35B06, 35B09, 35B40, 35J10.1. Introduction
In this paper, we are concerned with qualitative properties of solutions to the following mixed local/nonlocal Schrödinger equation
(1.1) |
satisfying in and
Here above, for any , the fractional Laplacian is defined as
where is a suitable normalization constant, whose explicit value does not play a role here, and means that the integral is taken in the Cauchy principal value sense.
Moreover, we suppose that and
(1.2) |
We recall that, on the one hand, the Schrödinger equation is the fundamental equation of physics for describing quantum mechanics. Feynman and Hibbs [31] formulated the non-relativistic quantum mechanics as a path integral over the Brownian paths, and this background leads to standard (non-fractional) Schrödinger equation. In this setting, during the last 30 years there have been important contributions to the analysis of the classical nonlinear Schrödinger equation, see e.g. [32, 42, 36, 43, 41, 48].
On the other hand, a nonlocal version of the Schrödinger equation has been introduced by Laskin, see [37, 38, 39]. Laskin constructed the fractional path integral and formulated the space fractional quantum mechanics on the basis of the Lévy flights. In the recent years, the fractional Schrödinger equation has been studied, under different perspectives, by many authors, see in particular [17, 30, 33, 29, 26, 25, 34] and the references therein.
In this framework (see e.g. the appendix in [25]) the classical and the fractional Schrödinger equations arise from a Feynman path integral over “all possible histories of the system”, subject to an action functional obtained as the superposition of a diffusive operator and a potential term: specifically, when the diffusive operator is Gaussian (i.e., as produced by classical Brownian motions), this procedure returns the classical Schrödinger equation, while when the diffusive process is -stable for some (e.g., as induced by Lévy flights), one obtains the fractional Schrödinger equation corresponding to the fractional Laplacian .
In view of this construction, it is also interesting to consider the case in which the diffusive operator in the action functional of the Feynman path integral is of “mixed type”, i.e., rather than possessing a given scaling invariance, it is obtained as the overlapping of two different diffusive operators, say a Gaussian and a -stable one (or, similarly, that the diffusive operator acts alternately in a Gaussian fashion and in a -stable way for “infinitesimal times” with respect to the time scale considered). This superposition of stochastic processes in the quantum action functional give rise precisely to the equation proposed in (1.1).
Our goal in this paper is to investigate existence and qualitative properties of solutions to (1.1). More precisely, we will obtain results of this type:
-
•
We prove the existence of nonnegative weak solutions by exploiting Ekeland’s variational principle (see Theorems 1.1 below).
-
•
We then obtain Hölder continuity and -regularity of weak solutions based on the -theory for the Laplacian and some basic properties of the kernel of the mixed Bessel potential (see Theorems 1.2 and 1.3 below). To overcome the difficulties caused by the fact that the equation involves operators of different orders and therefore is not scale-invariant, we develop a “piecewise” argument presented in Appendices A-B which is new in the literature.
-
•
Based on the -regularity result, we next establish a -regularity result by combining a suitable truncation method and a covering argument (see Theorem 1.4 below).
-
•
Finally, we discuss qualitative properties of classical solutions, such as positivity, power-type decay at infinity, and radial symmetry (see Theorem 1.5 below).
We now state the main results of this paper.
Theorem 1.1 (Existence of nonnegative weak solutions).
There exists a nontrivial and nonnegative weak solution of (1.1).
The result above will be established by exploiting the Mountain Pass geometry of the functional associated with (1.1). Additionally, we will show that the energy of the solution in Theorem 1.1 is precisely at the Mountain Pass level, see Corollary 2.6.
We next establish some regularity results which provide useful auxiliary tools in the analysis of the equation under consideration. For this, we notice that the pseudo-differential operator is the infinitesimal generator of a stochastic process , where is the mixture of the Brownian motion and an independent symmetric -stable Lévy process, and, as such, can be characterized using the Fourier Transform as
Hence, we can define a heat kernel associated with the above operator as
(1.3) |
for and .
In this context, the heat kernel may be viewed as a transition density of the Lévy process and, relying on probabilistic methods, some properties about the transition density of have been established: for example, Song and Vondraček [45] obtained upper and lower bounds on the transition density of by comparing the transition densities of the Brownian motion and the -stable process.
On a related note, Bogdan, Grzywny, and Ryznar [21] obtained general bounds on the transition density of a pure-jump Lévy process based on Laplace transform arguments (by adapting this method appropriately, one can find a sharp upper bound on the transition density of , but no lower bound since the Lévy-Khintchine exponent of the operator is not strictly between and ).
Furthermore, Cygan, Grzywny and Trojan [24] provided the asymptotic behaviors of the transition density of a Lévy process whose Lévy-Khintchine exponent is strictly and regularly varying between the indexes and (from these methods, since in our framework the index of the stochastic process at zero is and at infinity is , only the asymptotic formula for the tail transition density of can be obtained).
Additionally, the Lévy process may be considered as a subordinate Brownian motion, and thus, with the aid of suitable Tauberian Theorems developed by [18], the asymptotic behaviors of the Green function of near zero and infinity were established by Rao, Song and Vondraček [44].
In this paper we will revisit these results, framing them in the setting needed for our purposes, and provide a new set of proofs making use of Fourier analysis techniques (see Theorem 3.1). These properties will serve as an essential step to investigate some basic features of the Bessel kernel , defined as
(1.4) |
In turn, these properties of (see Theorem 3.2) will be pivotal to establish a series of regularity results for the mixed operator which can be described as follows.
The first regularity result that we present deals with the Hölder continuity of weak solutions of problem (1.1):
Theorem 1.2 (Hölder continuity).
Let be a weak solution of (1.1).
Then, for some and
for some , depending only on , , and .
We also establish a -regularity result:
Theorem 1.3 (-regularity).
Let be a bounded weak solution of (1.1).
Then, for any and
for some , depending only on , and .
Moreover,
From Theorem 1.3, we will also deduce the -regularity result below.
Theorem 1.4 (-regularity).
Let be a bounded weak solution of (1.1).
Then, for any and
for some , depending only on , , and .
Finally, we obtain the power-type decay at infinity and the radial symmetry of classical positive solutions to (1.1) by comparison arguments and the method of moving plane, respectively:
Theorem 1.5 (Qualitative properties of positive solutions).
The problem in (1.1) admits a classical, positive, radially symmetric solution .
Furthermore, has a power-type decay at infinity, that is, there exist constants such that, for all ,
2. Existence of weak solutions
The aim of this section is to establish Theorem 1.1. For this, we will first introduce the functional setting and provide some basic definitions. The proof of Theorem 1.1 will then occupy the forthcoming Section 2.2.
2.1. Functional framework
Let and .
As usual, the norm in will be denoted by and the norm in will be denoted by .
We also recall the so-called Gagliardo (semi)norm of , defined as
We point out that the Gagliardo (semi)norm is controlled by the -norm, according to the following observation:
Lemma 2.1.
Let .
Then, there exists a constant , depending only on and , such that, for all ,
Proof.
Utilizing the Hölder inequality, we see that
as desired. ∎
Definition 2.2 (Weak solution).
We say that is a weak solution of (1.1) if, for all ,
We consider the functional defined as
We point out that, in light of Lemma 2.1 and the classical Sobolev embedding, the functional is finite for all .
Furthermore, we have that and, for all ,
As a consequence, being a weak solution of (1.1) according to Definition 2.2 is equivalent to being a critical point of the functional .
Now we introduce an auxiliary problem. For this, we use the notation and , so that , and we consider the following problem
(2.1) |
The corresponding functional is given by
(2.2) |
We point out that and, for all ,
Accordingly, being a weak solution of (2.1) is equivalent to being a critical point of .
The role of this auxiliary functional is made explicit by the following observation:
Lemma 2.3.
Let be a critical point for .
Then,
(2.3) | a.e. in . |
Moreover,
(2.4) | is a critical point for . |
Proof.
We make the following observation:
Lemma 2.4.
We have that and
(2.6) |
Proof.
Since , we have that is unbounded from below and therefore .
Moreover, given with , by the Sobolev inequality one has that
where depends on and .
We remark that is a local minimum of rather than a global minimum. Thus the functional fulfills the geometric hypothesis of the Mountain Pass Theorem, but the Palais-Smale condition is not necessarily satisfied. To overcome this issue, we will make use of the following result, established in [23], which allows us to prove that (and therefore ) has a positive critical value, thus entailing Theorem 1.1.
Lemma 2.5.
([23, Lemma 2.18]) Let . Assume that is bounded in and that there exists such that
Then, in for all .
2.2. Existence of weak solutions
With the preliminary work done so far, we can now complete the proof of Theorem 1.1.
Proof of Theorem 1.1.
We recall the definition of the auxiliary functional in (2.2). In light of Lemma 2.3, in order to find a nonnegative weak solution of (1.1), we focus on finding a nontrivial critical point for the auxiliary functional .
We claim that
(2.7) | the functional has a nontrivial critical point. |
To prove this, we use the Ekeland’s variational principle (see e.g. [40]) and obtain that there exists a sequence such that, as ,
(2.8) |
where is given in (2.5).
We notice that, from (2.8), for large enough,
(2.9) |
It follows that is bounded in , and therefore, up to a subsequence, converges to some weakly in and strongly in for all .
As a result, for any ,
(2.10) |
which entails that .
Now, for every , we set
and we claim that
(2.11) |
Indeed, clearly . Moreover, since converges strongly in to some , we have that .
Hence, it remains to show that . For this, we argue by contradiction and suppose that . Then, we deduce from Lemma 2.5 that in . As a consequence of this, and recalling (2.8), we have that
It now follows from (2.11) that we have that there exists a sequence such that, for all (and possibly taking a subsequence of ),
(2.12) |
In light of (2.12), we define
(2.13) |
and we see that . Therefore,
and, for all ,
and thus, in view of (2.8), we have that, as ,
(2.14) |
Now, using (2.9) with replaced by , we conclude that there exists such that
(2.15) | converges to weakly in and strongly in for all . |
In particular, we can show that the nontrivial critical point for found in the proof of Theorem 1.1 is at the critical level given by (2.5):
Corollary 2.6.
Then,
(2.17) |
Proof.
We first check that
(2.18) |
Indeed, we observe that
and therefore, in light of (2.16), we deduce that .
Moreover, for every ,
and thus, since , we have that as . Hence, there exists some such that .
As a consequence, we consider the path for all and notice that . Thus,
(2.19) |
In addition, let us define for all . In this way, we compute
and we see that has a unique maximum .
Thus, we have that , which implies that . Accordingly, we have that for all . By combining this information with (2.19), we obtain (2.18).
3. Regularity of weak solutions
In this section we focus on the Hölder continuity, and -regularity results for weak solutions of (1.1) amd prove Theorems 1.2, 1.3 and 1.4. These regularity estimates will be the basis for the qualitative analysis we carry out in the next section.
To this end, we start with some basic properties of the heat kernel (1.3) and the Bessel kernel (1.4) which are essential for proving Theorems 1.2 and 1.3, as summarized by the following two results.
Theorem 3.1 (Properties of the heat kernel).
Let and . Let be as defined in (1.3).
Then,
-
•
is nonnegative, radially symmetric and nonincreasing with respect to .
-
•
There exist positive constants and such that
and where and .
Theorem 3.2 (Properties of the Bessel kernel).
Let and . Let be as defined in (1.4).
Then,
-
(a)
is positive, radially symmetric, smooth in and nonincreasing with respect to .
-
(b)
There exist positive constants and such that, if ,
and, if ,
-
(c)
There exists a positive constant such that, if ,
-
(d)
If , then for all . If , , then for all .
-
(e)
If and , then the function is a solution of
For the reader’s convenience, we provide the detailed proofs of Theorems 3.1 and 3.2 by using Fourier analysis techniques in Appendices A and B.
3.1. Hölder continuity of weak solutions
We devote this section to the proof of Theorem 1.2, which will achieved by combining -theory and Theorem 3.2.
3.1.1. -regularity of weak solutions
We shall explore the -regularity theory for weak solutions of linear equations, which is a pivotal step towards the proof of the Hölder continuity for the nonlinear equation, which will be obtained combining the -theory and a localization trick.
Lemma 3.3.
Let and . Let be a solution of
Then, and
for some constant depending on , and .
Proof.
The gist of the proof relies on checking that satisfies an equation of the type , for some function , in order to apply the classical Calderón–Zygmund regularity theory to this equation. For this, the core of the argument will be to check that (and that ).
The technical details of the proof go as follows. From [46, Theorem 3, page 135], we know that we can identify the Sobolev space with the space
(3.1) |
In light of this, it suffices to show that .
We first claim that
(3.2) |
To this end, we point out that, in the distributional sense,
where the kernel is given in (1.4). Since (thanks to (b) of Theorem 3.2), the Young’s convolution inequality gives that
for some constant . This estaslishes (3.2).
Moreover, we remark that satisfies
where |
Here above and in what follows denotes the Dirac’s delta at .
Now we define
We claim that
(3.3) |
Once the claim in (3.3) will be established, the proof of Lemma 3.3 can be completed as follows. One notices that , and therefore
From this, one concludes that and that , as desired.
Hence, to complete the proof of Lemma 3.3, we now focus on the proof of (3.3). For this, we define, for every , , and ,
and we observe that
(3.4) |
Also, for any , we notice that
(3.5) |
We split as , where
We will now prove in Step 1 that and in Step 2 that . Thus, Step 1 and Step 2 will give the desired claim in (3.3).
Step 1. We prove that by exploiting the rescaling property (3.4). To this end, we observe that for every . Thus, we pick and use the Fourier Inversion Theorem for the radial function (see e.g. [20, Chapter II]). In this way, we find that
(3.6) |
where denotes the Bessel function of first kind of order .
We claim that
(3.7) |
Indeed, from (3.6) it follows that
where is the Bessel function of the third kind and Re denotes the real part of .
We consider the straight line
and, using the Residue Theorem, we find that
Furthermore, we observe that, for any and ,
(3.8) |
Employing (LABEL:estimate_of_H), we see that
(3.9) |
By combining (LABEL:H11) with (3.9), one obtains that, if and ,
for some constant . As a result, for any , there exists independent of such that, for every and ,
which establishes (3.7), as desired.
By combining (3.4) with (3.7), we know that there exists depending on and such that, when and ,
Owing to this and (3.5), we conclude that, for every ,
and, for every ,
From the last two formulas we obtain that , as desired.
Step 2. We now prove that . This will be a byproduct of the rescaling property (3.4). The full argument goes as follows. Since , one has that . We thus use again the Fourier Inversion Theorem for the radial function , finding that
We now claim that
(3.10) |
Indeed, using a similar argument as in the proof of Step 1, for any and , one has
3.1.2. -regularity of weak solutions
We dedicate this part to show the -regularity of weak solutions based on Lemma 3.3 and the usual iteration technique.
For this, we first make the following observation:
Lemma 3.4.
For all , let
Then, for every there exists with such that
Proof.
Denote by
and suppose by contradiction that, for all ,
Also, notice that , and therefore, for all , we have that . Hence, the following limit exists
and moreover, by construction, .
In addition,
and thus, taking the limit in ,
Solving for , we find that , which gives the desired contradiction. ∎
Proof of Theorem 1.2.
We consider a sequence of cut off functions for any satisfying
(3.11) |
Let solve
(3.12) |
Then
Moreover,
where is given by (1.4).
Let now be either equal to the critical exponent if or any real number in if . Let also .
Moreover, we set . We observe that , thanks to (1.2). Hence, in light of Lemma 3.4, we can define
and, for every ,
In this way, we have that
(3.13) |
Also, for every , we set .
In this setting, we claim that, for every ,
(3.14) |
The proof of this claim is by induction over .
We first check that (3.14) holds true when . For this sake, we notice that , and therefore, since , one has that . From this and Lemma 3.3, applied to the equation for in (3.12), it then follows that and
(3.15) |
up to renaming . The last inequality in (3.15) uses the Sobolev Embedding (see also [22, Corollary 9.11] when ). This establishes (3.14) when .
Now we suppose that (3.14) holds true for all the indexes in with and we prove it for the index . For this, we observe that, since by construction, we have that and .
Moreover, owing to the Hölder inequality (used here with exponents and ) and the smoothness of away from the origin and its decay (recall (a) and (b) of Theorem 3.2), we have that, for every ,
This and the triangle inequality give that
From this and the Sobolev Embedding, we deduce that
up to renaming , depending on , , and . Hence, using the inductive assumption,
Using again the Sobolev Embedding and renaming once more,
Hence, we have that and
We can therefore use Lemma 3.3 for the equation (3.12) for . In this way, we obtain that and
for some constant , depending on , , and . This completes the proof of the claim in (3.14).
Using (3.14) with , we find that
(3.16) |
Now, in view of (3.13), we distinguish two cases, either or .
If we use the Sobolev Embedding Theorem, by choosing
and we see that . In particular, from (3.16), it follows that
(3.17) |
Moreover, using the smoothness of away from the origin, when one has that
By combining this and (3.17), we deduce that, for any , with ,
and
The ball is centred at the origin, but we may arbitrarily move it around . Covering with these balls, we conclude that for some .
This completes the desired result when and we now focus on the case .
Recalling (3.1) and using [46, Theorem 3, page 135], we see that . Furthermore, owing to [46, formula (40), page 135], we know that , where
Thus, from [30, Theorem 3.2], it follows that and
(3.18) |
for some constant .
Using the decay properties of as given by Theorem 3.2, we obtain that, for any ,
(3.19) |
By combining (3.16), (3.18) and (3.19), we have that
This gives that and therefore, by Lemma 3.3, we deduce that . Since , we observe that . Hence, using the Sobolev Embedding Theorem, we have that , with
and
(3.20) |
3.2. -regularity of weak solutions
We aim here to establish the -regularity of weak solutions of (1.1) based on the -theory and the smoothness of the kernel defined by (1.4).
Proof of Theorem 1.3.
We consider the cut off function satisfying the properties in (3.11) with . Let solve
Then
Moreover, we deduce that
We recall that , thanks to Theorem 1.2.
Now, owing to the Hölder inequality and the smoothness of away from the origin and its decay (recall Theorem 3.2), we have that, for every ,
and
(3.21) |
In the light of these estimates, we can focus on the regularity of . For this, we observe that, for any ,
(3.22) |
That is, for any . From Lemma 3.3, it then follows that for any , and in particular one can take any . Accordingly, from the Sobolev Embedding, we deduce that .
We point out that the ball is centred at the origin, but we may arbitrarily move it around . Covering with these balls, we conclude that for any . This completes the proof of Theorem 1.3. ∎
3.3. -regularity of weak solutions
The goal of this section is to establish the -regularity result for solutions of problem (1.1), as stated in Theorem 1.4. For this, we combine a suitable truncation argument for the solution with the -regularity argument.
We point out that Theorem 1.4 can be obtained by appropriately modifying [47, Theorem 1.6]. For the convenience of the reader, we sketch the proof in the following subsections.
To begin with, we introduce some notations. For , , and , we denote the interior norms as follows:
3.3.1. A mollifier technique and a truncation argument
Let solve (1.1), and let be a standard mollifier. For every , and , we denote by
Also, we set
Then, we have that
(3.23) |
Moreover, the following regularity estimates on and follow as a direct consequence of their definitions:
Lemma 3.5.
Let . Then, and
If in addition , then, for every ,
We now use a cut off argument for to get the a -estimate for . Consider a cut off function satisfying
and let
We point out that
With this notation, one obtains the following result:
Lemma 3.6.
([47, Lemma 5.4]) Let and . Let be a solution of
Then, there exists such that satisfies
In particular,
(3.24) |
for some positive constant .
As a consequence of Lemma 3.6, setting , we have that there exists such that satisfies
In particular, employing (3.24), one finds that
(3.25) |
We also observe that, since , for all , there exists such that
where
Thus, combining this with [35, Theorem 4.6], one deduces that for all there exists such that
3.3.2. Interior -regularity
The estimate in (3.26), coupled with the following statement, will allow us to obtain that the -norm of is bounded uniformly in some ball with respect to and thus use Arzelà-Ascoli Theorem to complete the proof of Theorem 1.4. The technical details go as follows:
Proposition 3.7.
([47, Proposition 4.3]) Let , and . Suppose that, for any , there exists such that, for any and any , we have that
(3.27) |
Then, there exist constants , , depending only on , and , such that
Proof of Theorem 1.4.
We will use Proposition 3.7 with , so that, for every ,
Moreover, we notice that the estimate in (3.26) tells us that formula (3.27) is verified in our setting with replaced by , and
Therefore, we are in a position of exploiting Proposition 3.7, thus obtaining that, for every ,
From the Arzelà-Ascoli Theorem, we obtain that , for every , and
Hence, a covering argument and Theorem 1.3, give that
where the constant depends on , , and .
The ball is centered at the origin, but we may arbitrarily move it around . Covering with these balls, we obtain the desired result. ∎
4. Qualitative properties of positive solutions
In this section, we are concerned with the positivity, the decay at infinity and the radial symmetry of classical solutions of (1.1), as stated in Theorem 1.5.
4.1. Power-type decay of classical positive solutions
In this part, we shall apply the Maximum Principle and comparison arguments to obtain the decay at infinity of classical positive solutions of (1.1).
4.1.1. Existence of classical positive solutions
We devote this part to establish the existence of classical positive solutions.
Theorem 4.1.
Let and .
Then, problem (1.1) has a classical solution, which satisfies in . Moreover,
(4.1) |
Proof.
Hence, we can now focus on the proving that in . For this, we argue by contradiction and we assume that there exists a global minimum point at which . Accordingly, we have that and . As a result, we deduce from (1.1) that
which is a contradiction.
4.1.2. Power-type decay of classical positive solutions
In this part, we exploit suitable barriers constructed using the Bessel kernel to establish the following result:
Theorem 4.2.
Let and . Let be a positive classical solution of (1.1).
Then, there exist constants such that, for every ,
To prove this result, we start with the following two lemmata, which construct suitable subsolutions and supersolutions. We start with the analysis of the subsolution:
Lemma 4.3.
There exists a function satisfying
(4.2) |
and, for every ,
(4.3) |
for some constant .
Proof.
We consider the function , where is the characteristic function of the ball .
Also, in light of the smoothness of (see e.g. (c) of Theorem 3.2), we have that .
The supersolution is constructed exploting Theorem 3.2 with a parameter in place of :
Lemma 4.4.
Let and . Let and
Then,
-
(a)
is positive, radially symmetric, smooth in and nonincreasing with respect to .
-
(b)
There exist positive constants and such that, if ,
and, if ,
-
(c)
There exists a positive constant such that, if ,
-
(d)
If , then for all . If , , then for all .
-
(e)
There exists a positive constant such that, if ,
Exploting the kernel given by Lemma 4.4, we are able to construct the supersolution as follows:
Lemma 4.5.
There exists satisfying
Also, when we have that
for some constant .
Proof.
We consider the function , where is defined in Lemma 4.4 with .
Thus, and it satisfies
Moreover, from Lemma 4.4, when we have that
for some positive constant , that may change from step to step.
Furthermore, using the smoothness of (see e.g. (a) of Lemma 4.4), we have that . We claim that for any .
Indeed, we point out that for all and also we see that
Then,
where .
Since , we have that for all and
These facts and Lemma 3.3 imply that for all . By the Sobolev Embedding, we obtain that for any and also
Proof of Theorem 4.2.
We consider the function given by Lemma 4.3. Utilizing the positivity and continuity of and in , we can find a constant such that in .
We claim that
(4.4) |
Indeed, for the sake of contradiction, we assume that there exists a global strictly negative minimum point . Since , we have and , and thus
This contradiction implies (4.4).
We now focus on the bound from above. Owing to (4.1), we can find some such that
In this case, we make use of the function given by Lemma 4.5. From the positivity and continuity of and in , there exists such that in .
4.2. Radial symmetry of positive solutions
Our aim is now to prove the radial symmetry of the positive solution found in Theorem 4.1. The main statement of this section is the following:
To establish Theorem 4.6, we will exploit the moving planes method. To begin with, we recall some notation. For any , we set
Moreover, for any , we set and .
With this notation, we first establish the following:
Lemma 4.7.
If and in , then in .
Proof.
Suppose by contradiction that there exists such that . In the light of Theorem 1.4, we know that . We thus have , and therefore
which implies that in . This is in contradiction with the assumption and therefore the desired result is established. ∎
Now, we define
Lemma 4.8.
We have that .
To check this, we will need the following estimate:
Lemma 4.9.
Let . Then, for any measurable set , there exists , depending on and , such that, for all measurable functions ,
The proof of Lemma 4.9 follows the same line as the one of [30, Lemma 5.2]. We provide here the details for the facility of the reader, since some exponents need to be adjusted to our setting.
Proof of Lemma 4.9.
Notice that if is unbounded, the claim is obviously true. Hence, we can suppose that .
From (d) of Theorem 3.2 we know that if , then for and if , then for .
We point out that, since , setting , we see that when . Thus, using the Hölder inequality with exponents and , we have that, for every ,
In addition, by Young’s convolution inequality,
As a result, we conclude that
as desired. ∎
We will also need the following intermediate estimate:
Lemma 4.10.
Let . Let
(4.6) |
Then,
(4.7) |
Proof.
Owing to the fact that is radial symmetric (recall (a) of Theorem 3.2), we deduce that
Moreover, we point out that for every , . Collecting these pieces of information and recalling that is nonincreasing with respect to , we conclude that, for every ,
Therefore, by the positivity of and the Mean Value Theorem, we derive that
From this and the Hölder inequality, we obtain (4.7).
Proof of Lemma 4.8.
We recall the definition of in (4.6) and we observe that, for every ,
Thus, there exists sufficiently negative such that for all . Therefore, using this information into (4.7), we find that, for all ,
This implies that, for all , we have that , and therefore, for all , we see that .
Now, from Lemma 4.7 it follows that, for all ,
(4.8) | either in or in . |
We claim that
(4.9) | if in for some , , then . |
Indeed, suppose that and let . Then, for all ,
That is, is periodic in the first coordinate of period , violating the fact that is non-trivial and with finite -norm. This establishes (4.9).
Lemma 4.11.
We have that .
Proof.
With this preliminary work, we can prove the following:
Lemma 4.12.
We have that for all .
Proof.
We claim that there exists , depending on , , and , such that
(4.12) | in , for all . |
Indeed, we recall the definition of in (4.6) and that, for any , the estimate in (4.7) holds true, namely
(4.13) |
Now we pick large enough such that, for all ,
(4.14) |
Moreover, from (4.11) we deduce that, for small enough, we can find such that
In addition, by the continuity of , up to taking smaller, we have that, for all ,
This implies that, for all ,
Thus, for all , we have that
As a result, up to taking even smaller, we find that, for all ,
As a consequence of this and (4.14), for all ,
Pluggin this information into (4.13), we thereby obtain that
This, in turn, implies that for all , which gives the desired claim in (4.12).
Owing to (4.12) and Lemma 4.7, we conclude that, for all , either in or in . This and (4.9) give that the second possibility cannot occur, and therefore, for all , we have that in .
This is a contradiction with the definition of and thus the desired result is established. ∎
Appendix A Properties of the heat kernel
In this section, we focus on the fundamental properties of the heat kernel introduced in (1.3), with the aim of establishing Theorem 3.1. For this, we observe that the main step to establish Theorem 3.1 is to investigate the inverse Fourier transform and the asymptotic behavior of by scaling techniques. We thus need to overcome the additional difficulties caused by the fact that the operator that we take into account is not scale invariant, and therefore our analysis cannot rely entirely on either the purely classical or the purely fractional counterparts, as given in [19, 30].
To start with, let us define a “two-scales” function for , , and as
(A.1) |
which satisfies the following rescaling properties
(A.2) |
We notice that if , then , while if , then . As a consequence of this, we shall take -scaling for and take -scaling for in order to discuss the properties of the heat kernel .
A.1. Nonnegativity of heat kernel
We perform some auxiliary analysis on the kernel defined in (1.3). For this sake, we recall a result contained in [30, Lemma A.2].
Lemma A.1.
If and are radially symmetric, nonnegative and decreasing in , then is radially symmetric and decreasing in .
Then, we have:
Lemma A.2.
Let and . Let be defined as in (1.3).
Then, is nonnegative, radially symmetric, and nonincreasing with respect to .
Proof.
We point out that, being the Fourier transform of a radially symmetric function, is radially symmetric in . To prove the nonnegativity of , we adapt the arguments in [30].
For all , we define the radially symmetric nonnegative functions as
and, for any ,
where and are such that and , respectively.
In this way, we have that
Let also
Then, it is immediate to check that
where if .
We now define, for all ,
the convolution product being taken times.
By the properties of the Fourier transform with respect to the convolution product, one has that, for all ,
(A.3) |
We define, for all ,
and, by combining (A.3) and the fact that , we see that
We note that the right-hand-side of the above equation converges to pointwise as . Furthermore, since, for all ,
this convergence also holds in , that is, as ,
As a result, taking the inverse Fourier transform, we see that converges in to for any .
Since is nonnegative for all and , we deduce the nonnegativity of .
By scaling and exploiting the continuity of with respect to , one derives the nonnegativity of for all .
The monotonicity of the heat kernel can be also deduced by the fact that and are nonincreasing in and Lemma A.1. ∎
A.2. Asymptotic formulae for the heat kernel
To obtain the bounds on , we point out two asymptotic formulae for as defined in (A.1). These asymptotics are described in the next Lemmata A.3 and A.4.
Lemma A.3.
Let , , and be as defined by (A.1).
Then, for any , there exists independent of such that, for every and ,
That is,
and the limit is uniform with respect to .
Proof.
From the definition of , one has that
By using the Fourier Inversion Theorem for radial functions, see e.g. [20, Chapter II], and the Bessel functions in [28], it is immediate to obtain the -dimensional integral representation of , namely that
(A.4) |
where denotes the Bessel function of first kind of order .
We also recall that, for all ,
and
Consequently,
(A.5) |
As a result,
(A.6) |
where is the Bessel function of the third kind and Re denotes the real part of .
We now choose a straight line running from to in the upper half-plane and making a angle with the positive real axis, that is
(A.7) |
Furthermore, applying the Residue Theorem, we find that
(A.8) |
where denotes the arc
We claim that, for every and ,
(A.9) |
Indeed, since , one has that
(A.10) |
Also, from the fact that
one obtains that
(A.11) |
By combining (A.11) with (LABEL:step_1_estimate_of_H1), we conclude that
which implies (A.9), as desired.
Now we claim that for any there exists independent of such that, for every and ,
Indeed, we first make use of the properties of on defined in (A.7) to prove that the integral in the right-hand-side of (LABEL:x_H(x,1,eta)) is uniformly bounded with respect to and .
Let , from the definition of (see e.g. [28, page 21]), one has
(A.13) |
Furthermore, we derive that, for every and ,
(A.14) |
By combining (LABEL:x_H(x,1,eta)) with (A.14), one has
(A.15) |
Let us estimate the term . For some large enough, one has
(A.16) |
From (A.14), we observe that
As a result, for any , taking
(A.17) |
one has that
(A.18) |
We now turn to estimating . For this, we take as in (A.17) and we remark that
Note also that, for any , large enough and ,
(A.19) |
By combining (A.14) with (LABEL:R_0_estimate), we see that for every , there exists such that, if ,
which implies that
(A.20) |
Recalling (A.16), by combining (A.18) with (A.20), we deduce that, for every , there exists , depending only on , and , such that, for every ,
Regarding , from (LABEL:A1_A2), we see that, for every , there exists , depending only on , and , such that, for every ,
We take and we use (LABEL:A1_A2) to find that, for every ,
That is,
(A.21) |
where the limit is uniform with respect to .
Applying the Residue Theorem again to the region composed of and the straight line
(A.22) |
we obtain that
where
Lemma A.4.
Let , , and be as defined by (A.1).
Then, for any , there exists independent of such that, for every and ,
That is
(A.23) |
and the limit is uniform with respect to .
Proof.
As in the proof of Lemma A.3, first of all, the 1-dimensional integral representation of follows from (A.5), that is
It thus follows that
Exploiting the Residue Theorem once again, and referring to (LABEL:L1) and (A.9), we infer that
Moreover, owing to (A.14), we observe that, for every ,
(A.24) |
We now focus on the term . For some large enough, one deduces that
(A.25) |
As for , we notice that, utilizing (A.14), for every ,
Therefore, for any , picking
(A.26) |
one finds that
(A.27) |
As regards , if is as defined in (A.26) we have that
We notice that, for any , large enough and ,
(A.28) |
Utilizing (LABEL:R_1_estimate), we obtain that for every , there exists such that, for every ,
As a consequence of this and (A.14), one obtains that
(A.29) |
Recalling (A.25), by combining (A.27) with (A.29), we deduce that for every , there exists , depending only on , and , such that, for every ,
Regarding , thanks to (LABEL:Q1_Q2), we see that for every , there exists , depending only on , and , such that, for every ,
Taking , it follows from (LABEL:Q1_Q2) that, for every and ,
That is,
and the limit is uniform with respect to .
A.3. Bounds on the heat kernel
In this section, we establish the following result related to the upper and lower bounds on the heat kernel by using the two asymptotic formulae in Lemmata A.3 and A.4.
Lemma A.5.
Let and . Let be as defined in (1.3).
Then, there exist positive constants and such that
(A.30) |
(A.31) |
where and .
Proof.
We observe that, by the definition of in (1.3), for every ,
(A.32) |
for some constant depending only on and .
In addition, in the light of Lemma A.3, we know that there exists independent of such that, for all and ,
(A.33) |
where .
Furthermore, for every , one has that . Hence, by combining (A.2) with (A.33), one concludes that
That is,
Exploiting (A.2), one sees that, if and ,
(A.34) |
Similarly, for every , we have that . In this way, using the nonnegativity of , by combining (A.2) with Lemma A.4, we know that there exists independent of such that, for every , and ,
(A.35) |
Gathering (A.32), (A.34) and (A.35), we obtain (A.30), as desired.
Let us now estimate (A.31). We first claim that for any , there exists such that, for every and ,
(A.36) |
Indeed, from the definition of it follows that
Accordingly, recalling the fact that and taking , one derives that
Subsequently, applying the Residue Theorem, we observe that
and we obtain (A.36), as desired.
Appendix B Properties of the Bessel kernel
The main aim of this section is to prove Theorem 3.2 by exploiting the properties of the heat kernel defined in (1.3).
To start with, in view of the definition of , we observe that, from the nonnegativity, radial symmetry and monotonicity of for all , it follows that the kernel is nonnegative, radially symmetric and nonincreasing in .
B.1. Decay of the Bessel kernel
In this part, our goal is to employ the lower and upper bounds on the heat kernel (see e.g. (A.30) and (A.31)) to obtain the decay of , as given by the following Lemmata B.1 and B.2.
Lemma B.1.
Let and . Let be as defined in (1.4).
Then, if ,
(B.1) |
and, if ,
for some positive constants and depending on and
Proof.
We shall make use of the nonnegativity of and (A.31) to prove the strict positivity of and a lower bound on the behaviour of .
Lemma B.2.
Let and . Let be as in (1.4).
Then, is positive and
for some constants and depending only on and
B.2. Smoothness of the Bessel kernel
In this section, we focus on the smoothness of the kernel given by (1.4). These properties of are useful in the proofs of our regularity results.
Lemma B.3.
Let and . Let be as in (1.4).
Then, there exists , depending only on and , such that, for all ,
Proof.
Since is radially symmetric, with a slight abuse of notation we write , with . Furthermore, recalling the definition of and employing (A.4), we infer that
Taking the derivative of the expression above with respect to , one finds that
(B.6) |
Concerning , owing to (B.1), one has that, for every ,
We now estimate . For this, integrating by parts in , we see that
Furthermore, integrating by parts in ,
As a consequence,
(B.7) |
From (A.1) and (A.30), it follows that, for any ,
(B.8) |
where the constant depends only on and .
Let us now estimate the integral . For this, for every , and we set
Using again the Fourier Inversion Theorem for radial functions, we see that
(B.9) |
where denotes the Bessel function of first kind of order . Furthermore, we observe that satisfies the following scaling properties:
(B.10) |
By the definition of in (B.7), we see that
We now focus on estimating and .
Step 1. We estimate . Let us start by calculating the value of for every . From (B.9), we infer that
Thus, owing to the fact that is the Bessel function of the third kind, we have that
Using again the Residue Theorem, we deduce that
where is defined in (A.7).
We now claim that for any there exists independent of such that, for every and ,
For this purpose, we notice that, for large enough,
To estimate , we note that, by the definition of (see e.g. [28, page 21]),
As a consequence, given , taking
(B.11) |
one has that
(B.12) |
We now consider where is defined in (B.11). Recalling (A.14), we have that
It follows that for any there exists such that, for every ,
which implies that
By combining this and (B.12), we deduce that for any there exists , depending only on , and , such that, for every ,
That is,
where the limit is independent of .
Moreover, applying the Residue Theorem again to the region composed of and the straight line given in (A.22), and using the modified Bessel function of the third kind (see e.g. [28, page 5]), it follows that
(B.13) |
From this, we infer that there exists , depending only on and , such that, for every ,
From (B.10), we deduce that, if and ,
(B.14) |
Moreover, according to the definition of , one has that
(B.15) |
Owing to (B.14) and (B.15), we conclude that, for every ,
(B.16) |
and, for every ,
As a consequence of this and (B.16), we have that, for every ,
(B.17) |
Step 2. We now estimate the . To this end, we remark that
where is defined in (B.13) and the limit is independent of .
Recalling (B.10), we see that there exists , depending only on and , such that, for all and ,
(B.18) |
Owing to (B.15) and (B.18), we conclude that, if ,
(B.19) |
and, if ,
By (LABEL:phi_2_x>m), one sees that, for every ,
(B.20) |
Finally, recalling (B.6), we have that
Owing to the decay of , by combining (B.8) with (B.21), we deduce that, for every ,
for some constant depending only on and .
Using similar arguments, we infer that, for every ,
for some constant depending only on and . ∎
B.3. Proof of Theorem 3.2
We observe that, gathering together the results of Sections B.1 and B.2, we obtain the desired claims in (a)-(d) of Theorem 3.2. Thus, it suffices to prove property (e) of Theorem 3.2. To this end, we introduce the following two ancillary results:
Lemma B.4.
Let be in the Schwartz space and . Then, the function
belongs to .
Proof.
We have that
which is finite. ∎
Lemma B.5.
Let be in the Schwartz space. Then, the function
belongs to and
(B.22) |
Proof.
We have that
Therefore, by Fubini-Tonelli’s Theorem,
which is finite. ∎
With this preparatory work, we can now check that the Bessel kernel is the fundamental solution of the mixed order operator , as clarified by the following result:
Lemma B.6.
The Fourier transform of equals in the sense of distribution.
More explicitly, for every in the Schwartz space, we have that
(B.23) |
Proof.
By (1.4) and Fubini-Tonelli’s Theorem (whose validity is a consequence here of Lemma B.2), we have that
Hence, by (1.3),
Proof of (e) of Theorem 3.2.
Let and . Since is dense in , there exists a sequence such that
(B.24) |
From Lemma B.6, it follows that the function is a solution of
Namely, for every ,
(B.25) |
Moreover, owing to (B.24) and taking into account the fact that , we infer that in
By passing to the limit in (B.25), we conclude that the function is a solution of
References
- [17] C. J. Amick and J. F. Toland. Uniqueness and related analytic properties for the Benjamin-Ono equation—a nonlinear Neumann problem in the plane. Acta Math., 167(1-2):107–126, 1991. ISSN 0001-5962. URL https://doi.org/10.1007/BF02392447.
- [18] N. H. Bingham, C. M. Goldie, and J. L. Teugels. Regular variation. Number 27. Cambridge university press, 1989.
- [19] R. M. Blumenthal and R. K. Getoor. Some theorems on stable processes. Trans. Amer. Math. Soc., 95:263–273, 1960. ISSN 0002-9947. URL https://doi.org/10.2307/1993291.
- [20] S. Bochner and K. Chandrasekharan. Fourier Transforms. Annals of Mathematics Studies, No. 19. Princeton University Press, Princeton, NJ; Oxford University Press, London, 1949. ix+219 pp.
- [21] K. Bogdan, T. Grzywny, and M. Ryznar. Density and tails of unimodal convolution semigroups. J. Funct. Anal., 266(6):3543–3571, 2014. ISSN 0022-1236.
- [22] H. Brezis. Functional analysis, Sobolev spaces and partial differential equations. Universitext. Springer, New York, 2011. ISBN 978-0-387-70913-0. xiv+599 pp.
- [23] V. Coti Zelati and P. H. Rabinowitz. Homoclinic type solutions for a semilinear elliptic PDE on . Comm. Pure Appl. Math., 45(10):1217–1269, 1992. ISSN 0010-3640. URL https://doi.org/10.1002/cpa.3160451002.
- [24] W. Cygan, T. Grzywny, and B. Trojan. Asymptotic behavior of densities of unimodal convolution semigroups. Trans. Amer. Math. Soc., 369(8):5623–5644, 2017. ISSN 0002-9947. URL https://doi.org/10.1090/tran/6830.
- [25] J. Dávila, M. del Pino, S. Dipierro, and E. Valdinoci. Concentration phenomena for the nonlocal Schrödinger equation with Dirichlet datum. Anal. PDE, 8(5):1165–1235, 2015. ISSN 2157-5045. URL https://doi.org/10.2140/apde.2015.8.1165.
- [26] J. Dávila, M. del Pino, and J. Wei. Concentrating standing waves for the fractional nonlinear Schrödinger equation. J. Differential Equations, 256(2):858–892, 2014. ISSN 0022-0396. URL https://doi.org/10.1016/j.jde.2013.10.006.
- [27] S. Dipierro and E. Valdinoci. Elliptic Partial Differential Equations from an Elementary Viewpoint. A Fresh Glance at the Classical Theory. World Scientific, 2024.
- [28] A. Erdélyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi. Higher transcendental functions. Vols. I, II. McGraw-Hill Book Co., Inc., New York-Toronto-London, 1953. xxvi+302, xvii+396 pp. Based, in part, on notes left by Harry Bateman.
- [29] M. M. Fall and E. Valdinoci. Uniqueness and nondegeneracy of positive solutions of in when is close to 1. Comm. Math. Phys., 329(1):383–404, 2014. ISSN 0010-3616. URL https://doi.org/10.1007/s00220-014-1919-y.
- [30] P. Felmer, A. Quaas, and J. Tan. Positive solutions of the nonlinear Schrödinger equation with the fractional Laplacian. Proc. Roy. Soc. Edinburgh Sect. A, 142(6):1237–1262, 2012. ISSN 0308-2105. URL https://doi.org/10.1017/S0308210511000746.
- [31] R. Feynman and A. Hibbs. Quantum mechanics and path integrals. McGraw-Hill, New York, 1965.
- [32] A. Floer and A. Weinstein. Nonspreading wave packets for the cubic Schrödinger equation with a bounded potential. J. Funct. Anal., 69(3):397–408, 1986. ISSN 0022-1236. URL https://doi.org/10.1016/0022-1236(86)90096-0.
- [33] R. L. Frank and E. Lenzmann. Uniqueness of non-linear ground states for fractional Laplacians in . Acta Math., 210(2):261–318, 2013. ISSN 0001-5962. URL https://doi.org/10.1007/s11511-013-0095-9.
- [34] R. L. Frank, E. Lenzmann, and L. Silvestre. Uniqueness of radial solutions for the fractional Laplacian. Comm. Pure Appl. Math., 69(9):1671–1726, 2016. ISSN 0010-3640. URL https://doi.org/10.1002/cpa.21591.
- [35] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. ISBN 3-540-41160-7. xiv+517 pp. Reprint of the 1998 edition.
- [36] M. K. Kwong. Uniqueness of positive solutions of in . Arch. Rational Mech. Anal., 105(3):243–266, 1989. ISSN 0003-9527. URL https://doi.org/10.1007/BF00251502.
- [37] N. Laskin. Fractional quantum mechanics and Lévy path integrals. Phys. Lett. A, 268(4-6):298–305, 2000. ISSN 0375-9601. URL https://doi.org/10.1016/S0375-9601(00)00201-2.
- [38] N. Laskin. Fractional Schrödinger equation. Phys. Rev. E (3), 66(5):056108, 7, 2002. ISSN 1539-3755. URL https://doi.org/10.1103/PhysRevE.66.056108.
- [39] N. Laskin. Principles of fractional quantum mechanics. In Fractional dynamics, pages 393–427. World Sci. Publ., Hackensack, NJ, 2012.
- [40] J. Mawhin and M. Willem. Critical point theory and Hamiltonian systems, volume 74 of Applied Mathematical Sciences. Springer-Verlag, New York, 1989. ISBN 0-387-96908-X. xiv+277 pp. URL https://doi.org/10.1007/978-1-4757-2061-7.
- [41] K. McLeod. Uniqueness of positive radial solutions of in . II. Trans. Amer. Math. Soc., 339(2):495–505, 1993. ISSN 0002-9947. URL https://doi.org/10.2307/2154282.
- [42] K. McLeod and J. Serrin. Uniqueness of positive radial solutions of in . Arch. Rational Mech. Anal., 99(2):115–145, 1987. ISSN 0003-9527. URL https://doi.org/10.1007/BF00275874.
- [43] P. H. Rabinowitz. On a class of nonlinear Schrödinger equations. Z. Angew. Math. Phys., 43(2):270–291, 1992. ISSN 0044-2275. URL https://doi.org/10.1007/BF00946631.
- [44] M. Rao, R. Song, and Z. Vondraček. Green function estimates and Harnack inequality for subordinate Brownian motions. Potential Anal., 25(1):1–27, 2006. ISSN 0926-2601. URL https://doi.org/10.1007/s11118-005-9003-z.
- [45] R. Song and Z. Vondraček. Parabolic Harnack inequality for the mixture of Brownian motion and stable process. Tohoku Math. J. (2), 59(1):1–19, 2007. ISSN 0040-8735. URL http://projecteuclid.org/euclid.tmj/1176734744.
- [46] E. M. Stein. Singular integrals and differentiability properties of functions. Princeton Mathematical Series, No. 30. Princeton University Press, Princeton, NJ, 1970. xiv+290 pp.
- [47] X. Su, E. Valdinoci, Y. Wei, and J. Zhang. On some regularity properties of mixed local and nonlocal elliptic equations. 2023. Preprint.
- [48] X. Wang. On concentration of positive bound states of nonlinear Schrödinger equations. Comm. Math. Phys., 153(2):229–244, 1993. ISSN 0010-3616. URL http://projecteuclid.org/euclid.cmp/1104252679.