Symmetry and Monotonicity of Positive Solutions for a Class of Choquard Equation with Hardy Potential ()
1. Introduction
The following Schrödinger equation with convolution nonlinearity
(1)
where
with
as usual, is called as Choquard equation. The Choquard equation, as a significant class of nonlinear partial differential equations, plays a crucial role in self-gravitational quantum field theory. It can characterize some special physical phenomena, such as interactions among particles. When the potential function
is singular, the Equation (1) involving singular potential can simulate singular interactions of the motions of particles in an environment with a singular potential field. That is one of reasons why such Schrödinger equations containing such potentials in
have been paid great attention by many scholars in recent years.
When
and
, Moroz and Van Schaftingen [1] respectively discussed existence, nonexistence and sharp decay estimates of super-solutions to (1) in exterior domain with four different types of potentials: unperturbed Laplacian
, fast decay potentials
with
and
, Hardy potentials
with
and slow decay potentials
with
and
. As for symmetry results involving such Choquard-type nonlinearity, it should be noted that Le [2] established the classification of solutions to the integral equation
where
and
, by the method of moving planes in integral form.
For
with
and
, Zhu and Tang [3] considered existence and asymptotic behavior of solutions for (1) by virtue of variational methods and some analysis techniques. And analogous cases may be seen in [4] [5]. Moreover, Guo and Tang [6] addressed the indefinite case in which
is 1-periodic.
As
, the Equation (1) degenerates to
(2)
There are many noted works involving so-called Hardy-type potentials in the Whole space
. For instance, the study on entire solutions for (2) with
has been concerned for a long time. Based on variational method for minimizing problems together with sophisticated versions of the moving plane technique, Terracini [7] derived different results concerning positive solutions of (2), where
,
, and
is a super-linear function. The author first proved there is no any positive solution in
, provided that
and the first eigenvalue of the associated linear problem is non-positive. And it was shown that there can be solutions in
if and only if
as
and there is a positive solution
which is homogeneous of degree
. For the case of
, it was proved that there is a constant
such that if
there are two positive solutions in
, where one is radially symmetric and another is not. Felli and Pistoia [8] used the perturbation methods to obtain the singular or blowing-up solutions for the critical case about the Equation (2). For the power-nonlinearity case, Cîrstea and Fărcăşeanu [9] studied the existence and non-existence, uniqueness and non-uniqueness and the behaviors near zero or at infinity of solutions in
when
with
and
with
whose parameters are under distinct ranges. When
, Chen and Zou [10] founded the existence and symmetry results of the doubly critical Schödinger system. Kang [11] studied the existence of radial solutions for elliptic systems involving critical power-nonlinearity. And Esposito, López-Soriano, and Sciunzi [12] completed the classification of positive solutions for such systems with the singular-critical power nonlinearity by a suitable version of moving plane method. The case of normalized solutions can be found in [13] and references therein. Such Schrödinger-Hardy equations were also discussed in [3] [14]-[16] about the case involving nonlinearities satisfying Berestycki-Lions type conditions.
In addition, when
serves as a generalization of Hardy-type potentials, there are many noted works arising from (2). Felli [17] dealt with the critical case with potential term formed
with
and proved the existence of positive solutions by means of minimizing the associated Rayleigh quotient. Franca and Sfecci [18] studied the corresponding dynamical systems of (2) in the case where
is radial and
is radial function with different sign on the interior and exterior of a ball by applying to the Fowler transformation.
The aim of this paper is to study the symmetry of solutions to the following Schrödinger equation involving Hardy potential and Choquard-type nonlinearity, as one of special cases of (1),
(3)
where
,
and
. And the pair of parameters
lies in the set
Clearly, the set is a part of a ray in 2-dimensional plane, which lies in a region where allow the subsolution of (3) exists, see [1].
It should be pointed out that this is the first time the moving plane method has been applied to address partial differential equations involving convolution-nonlinear term in weak or integral sense, in order to obtain the symmetry and monotonicity of solutions. Indeed, many noted works about Choquard-type equations involving such singular potentials like Hardy potentials in
focus only on existence or non-existence of solutions. Especially, there are no optimal existence results about equation (3) under different ranges of parameters
,
and
, albeit with the fact that there are partial results discussed in [1]. However, this paper only concentrates on the behavior (more explicitly, symmetry) of solutions under suitable range of those parameters. That is, althogh the range of
in this paper is not optimal for the existence result of solutions, it is sharp for the symmetry result of solutions about
as (3) possesses a positive solution in
.
And this work is motivated by the study of symmetry of solutions in the works above, particularly the spirit of [7] which was faced with the same issue that all of difficulties resulting from the lack of the
-regularity of some possible solutions force one to replace the pointwise decay estimates on the equation itself by integral estimates as well as make the moving plane procedure work in weak or distributional sense. However, when dealing with the symmetric property, one prefer constructing suitable test functions in the concrete calculations rather than making Kelvin transform as [7] to get rid of the decay hypothesis on solutions at the infinity, since such transform would make the calculations more complex and add more extra assumptions on the integrablility of solutions in the presence of convolution nonlinearity. Furthermore, the equation cannot be reduced to ordinary differential equations like [7]. So it is hard in this paper to directly derive the symmetry results unless combining necessary integral estimates about the convolution term and moving plane technique briefly mentioned in [7]. It is also necessary for the application of strong maximum principle about weak solutions. Meanwhile, in order to circumvent the possible impact on the singularity of origin and infinity point, some standard cutoff functions in analysis will be also utilized subtly.
And the main result is as below.
Theorem 1. Let
,
and
. And let
satisfiy
with
. Suppose that
be a solution to (3), no matter whether
is singular at the origin or not. Then it follows that
is symmetric with respect to the hyperplane
and increasing in the
-direction in
. Further, if
is of
, we have
In particular, the solution is radial and radially decreasing about the origin.
Remark. As is known that moving plane technique firstly proposed by Axelander [19] and Serrin [20] and subsequently developed by many others, see e.g. Sciunzi for [21] and Chen and Li for [22], etc. In addition, unfortunately, there is no any available approach to address the case of
due to the deficiency of the analysis techniques used in this paper and therefore it is still open for this case.
The organizations in this paper are as follows. Some notations and preliminary results are stated in section 2 while the proof of main results are arranged in section 3. Moreover, the whole paper always denotes
or
by any constant whose value may be distinct from line to line, and, as for
, only the related dependence is specified in what follows.
2. Notations and Preliminary Results
In this section, some useful preliminary results will be given as below. To begin with, for any fixed
, to set
And the reflection of a point
though the hyperplane
is defined by
and
denotes by the reflection operator with respect to
. In other words, the change of the parameter
corresponds to the movement of the plane
. In the meantime,
Owing to the lack of decay estimates of solutions to (3), it is necessary to find a function to “cut off” the infinity to ensure that integral estimates mentioned below make sense. And note that the solution
has no definition at the point zero, which makes it possible for
to be a non-removable singular point of
. Hence, we must also truncate any small neighborhood of this possible isolated singularity so as to make the following estimates meaningful.
It is easy to see that there exist
satisfying
(i)
on
and
on
,
where
;
(ii)
on
;
(iii)
;
(iv)
and
satisfying.
(i)
on
and
on
,
where
;
(ii)
on
;
(iii)
.
Here
is arbitrary sufficient small positive number and
is any large number as well as both
and
are independent with each other.
For convenience, Hardy’s and Hardy-Littlewood-Sobolev inequalities respectively are as follows.
Lemma 2. (Hardy’s inequality) ([8] [23] [Lemma 1.1]) If
, then
and
where
is optimal and not attained.
Lemma 3. (The (weighted) Hardy-Littlewood-Sobolev inequality) ([24] [25]) Let
,
and
satisfying
. Assume that
and
, then there exists a positive constant
not depending on
and
such that
where
and
.
And the reader can refer [26] [27] and the references therein for more explicit details about cut-off functions.
We say
solves (3) if
fulfills
(4)
where
and the parameters
and
are supposed as mentioned in the introduction.
3. Main Results and Their Proofs
In this section, the proof of the main result will be shown later. Firstly, note that
is the solution of
(5)
for any test function
.
Lemma 4. Under the assumption of Theorem 1, for
, we have that
(6)
where
,
. Apparently,
for any
. Here
is the characteristic function of set
.
Remark. The auxiliary set
proposed in the term
in the formula (6) is not necessary, unless 0 is a non-removable singular point of
. In addition, without loss of generality, we may assume
for simplicity, anyway.
Proof. For
, without loss of generality, we simplistically denote by
in the subsequent statements.
Define
. By the density argument, one can plug
as a test function into (4) and (5), after subtracting (5) from (4), we have
(7)
In order to get the summability of
in
, it suffices to estimate
and
term by term. First, via Young’s and Hölder’s inequalities, we have
for
small enough with
, where we used the fact that
and
. And
is a fixed number between 0 and
. Likewise,
Since
It follows from Hardy’s inequality that
At last, we estimate
, being similar to the calculation as above.
Write
, and the Lagrange mean value theorem yields that
for fixed
, where
is a number depending only on
between
and
. By virtue of Hardy-Littlewood-Sobolev (short for HLS) inequality, we have
Next, set
and
Since
similar to the estimate of
by mean value theorem, HLS inequality and Hölder’s inequality, then we have
Summing up all estimates of
and
above, we get
Using Lebesgue’s dominated convergence theorem to take the limit on the right-hand side (and Fatou’s lemma on the left-hand side), it yields that
Namely,
This completes the proof.
With the above lemma, we give the next lemma, which makes sure the moving plane can start from negative infinity along
-axis.
Lemma 5. Under the assumption of lemma 4, for any
with
sufficiently large, we have
where
Proof. Analogous to the proof in the Lemma 4, taking the same test function
. We give more explicit estimates for
and
. At first, for
and
, choosing the parameter
, then Young’s and Hölder’s inequalities yields that
(8)
and
(9)
And
implies
as
due to the decay property of Lebesgue’s integral. Next, by virtue of Hardy’s inequality, we get
(10)
while
Thus, putting it into (10), it follows that
(11)
Write the last term in the curly braces as
as below, i.e.
Let
as mentioned above, and taking the double limit, it follows that
. Moreover, the second and third term in the curly braces are both infinitesimals as
and
, respectively. Indeed, we have the estimates as before:
and
where
is a constant depending only on dimension
. Up to now, there is still term
left. Based on the estimates for
and
in the preceding lemma, together with the Sobolev’s embedding theorem, we have
(12)
where
is the Sobolev embedding constant from
to
. Likewise, we have
(13)
Inserting (8)-(13) into (7) and making it up, then we deduce that
(14)
Employing Fatou’s lemma again, it follows (14) that
Note that
and
as
with
As a result, there exists
such that for
, it holds that
This means that
for any
. Namely,
. Observe that
on
, so
for any
due to the continuity. This is as desired.
Before the proof of the main result, we give a weaker version of strong comparison principle whose proof is similar to the classical strong maximum principle for elliptic equations or inequalities with divergence structure.
Lemma 6. Let
be a weakly superharmonic function in
. If
is continuous over
, then either
in
or
in
.
Proof. Since
in
, if
in
, we have done. Otherwise, setting
and it means
. Thus, it suffices to prove that
. And
weakly solves the following differential inequality
That is, for any nonnegative
, it holds
On the other hand, note that
is relatively closed to
due to the continuity of
. Let
with
small enough satisfying
, then
for any nonnegative
. So
restricted to
is also a weak superharmonic function in
. And
is still nonnegative with a minimum point
. Whence, it follows from the weak Harnack inequality [[28], Lemma 2.113] (in the case of
and
) that
which implies that
in
and thus
. That is,
is a interior point of
. Consequently,
is also open. Concluding all above,
and thereby
in
. This accomplishes the proof of the lemma.
With crucial lemmas all above at hand, next, we take up proving Theorem 1 with moving the hyperplane
from negative infinity along the
-axis.
Proof of Theorem 1. Now, we take up the proof of Theorem 1. And it is split into three steps.
Step1: We have
in
for
with
large enough, which contributes to Lemma 5.
Thus, define
where
Step2: We assert that
. Arguing by contradiction and assume that
. By the continuity, we see that
in
. Note that
and
for any nonnegative
. It suffices to take
, via Lemma 6, then we immediately deduce that
in
as
. Indeed,
would be undefined at
as long as
, which is a contradiction. Now, we manage to obtain the contradictory conclusion resulting from the fact that there exists some small
such that, for any
, it holds
in
whenever
. Actually, for arbitrary
, there are
and a compact set
(depending on
and
) so that
and
for every
. Define
on the compact
. So it follows that
and
for any
with some
. Also, we can assume that
. Since
and
, we assert that there is
such that
and thereby
for all
. Next, we repeat the proof of Lemma 5 but using test function
Hence, after recovering the first and last inequalities in (14), for any
, we have
(15)
Since
and
both vanish in a neighborhood of
, via the previous construction. Now, fix
then
combining with (15), we deduce
for every
. Hence,
for every
, and it means that
in
, which demonstrates the claim of step 2.
Step3. Conclusion. Similar to the step 1 and step 2, we may define
and
Simultaneously, we consider another test function , where . Continuing the proceeding procedure, we see which
is increasing along
-direction in
and symmetry with respect to
. This completes the proof of Theorem 1.■
Remark. Actually, since in
the
-axis is a direction that can be taken arbitrarily in the proof, we may get the radial symmetry of solutions.
4. Conclusions and Suggestions
The symmetry and monotonicity with respect to the plane
of solutions to a class of Choquard with Hardy potential is derived in this paper via applying the moving plane method. The key is to compare the size between solution
and its reflection
with respect to hyperplane
. By moving
to the limit position to get
, we prove
must be zero as the limit position is reached, i.e.
must stop at the position of
. We can get the opposite monotonicity of the solution towards to opposite direction along the according coordinate axis, thereby resulting in the symmetry of the solution. That is the idea of the moving plane technique. Of cause, there are two many mathematical techniques to deal with the problem in this paper which are partly distinct from others’ works introduced in the introduction.
As mentioned in the introduction, if there exists a solution of this class of Choquard equation within the range of parameters assumed in this paper, it must be symmetric about some hyperplane throughout the oringin point. And there are still some interesting open problems for further investigations in Mathematics or Physics. For instance, as for the case that potentials
is nonnegative and
, the physicial explanations can be found in [29]. However, is there any physical significance about
in such Choquard-type equations with Hardy potential ? In Mathematics, can the symmetry results be generalized to the system coupled by over two Choquard-type equations with Hardy potentials in different domains including
? When
does not belong to the set mentioned in the introduction, is there any symmetric positive solution to the equation? Namely, more difficultly, what is the sharp range of
for such problems? Thess points are quite fascinating in both Mathematics and Physics.