1. Introduction
It is well-known that important geometric inequalities, for example, the Sobolev inequality, Moser–Trudinger inequality, etc., and the existence of extreme functions play a key role to study partial differential equations. For a bounded domain
with
, we have
,
for
by the calssical Sobolev embedding theorem. Particularly, for
,
,
. But
. For the borderline case
, the Moser–Trudinger inequality is the perfect replacement. In 1971, Moser [
1] proved the sharpening of Trudinger’s inequality as follows:
for
and
stands for area of the
-sphere. Moreover,
is sharp, which means that if
, then the inequality (
1) can no longer hold. Inequality (
1) is the so-called Moser–Trudinger inequality, the extremal of which is related to the existence of solutions of some semi-linear Liouville-type equations.
As far as we know, there have been many important studies related to the Moser–Trudinger inequality, for example, [
2,
3,
4,
5,
6,
7,
8], etc. In the references listed above, readers can see the Moser–Trudinger inequality in
and in hyperbolic spaces, the existence of an extremal function for the Moser–Trudinger inequality, etc. These important geometric inequalities play a key role in geometry analysis, calculus of variations, and PDEs; we refer to [
9,
10,
11,
12,
13,
14] and references therein. And recently, the authors of [
15] studied a system of Kirchhoff type driven by the
Q-Laplacian in the Heisenberg group
. They obtained the existence of solutions via variational methods based on a new Moser–Trudinger-type inequality for the Heisenberg group
. Moreover, in [
16], the authors also focus on a Kirchhoff-type problem and establish the existence of a radial solution in the subcritical growth case by the Moser–Trudinger inequality and minimax method.
Let
. Clalnchi and Ruf [
17,
18] proved the weighed Moser–Trudinger-type inequality involving the radical functions in unit ball
B:
for any
,
. Moreover, the constant
is sharp, i.e., if
, the supremum in (
2) will be infinite. We note that the authors applied Leckband’s inequality [
19] to prove the weighed Moser–Trudinger-type inequality (
2). Note that when
, by the Pólya-Szegö principle, (
2) recovers the classical Moser–Trudinger inequality (
1). Furthermore, Roy [
20] proved the existence of an extremal function for inequality (
2).
Recently, many researchers have intended to establish anisotropic Moser–Trudinger-type inequalities. Let
be a nonnegative and convex function, the polar
of which represents a Finsler metric on
. By
, a Finsler–Laplacian operator
is defined by
where
. In Euclidean modulus,
is nothing but the common Laplacian. The Finsler–Laplacian operator is closely related to the Wulff shape, which was initiated in Wulff’ work [
21]. More details about the properties of
and
can be seen in
Section 2.
For a bounded smooth domain
, Wang and Xia [
22] proved that, for
,
where
denotes the volume of a unit Wulff ball in
.
In this paper, we intend to establish the anisotropic Moser–Trudinger-type inequality with logarithmic weight. We believe that these sharp inequalities will be the key tools to study the existence of solutions for some quasi-linear elliptic equations, such as the Finsler–Laplacian equation. For
, we let
which is the weight of logarithmic type defined on a unit Wulff ball
. And
represents the functions of completion of
with respect to the norm
Let be the subspace of of all radial functions with respect to F. In this paper, radial functions with respect to F means that , where .
In the following, for convenience, we denote
We now state our main results.
Theorem 1. For anywe have . Moreover, this constant is sharp, i.e., if , is infinite. Next, we prove the existence of an extremal function for the anisotropic Moser–Trudinger-type inequality with logarithmic weight.
Theorem 2. There exists such that, for , is attained.
Finally, we establish the Lions-type concentration-compactness property, which can be seen as an improvement of the anisotropic Moser–Trudinger-type inequality in Theorem 1 for some situations.
Theorem 3. Let be a sequence in such that and in . Then we havefor any . 2. Preliminaries
In this section, we give preliminaries involving the Finsler–Laplacian, co-area formula with respect to F and convex symmetrization of u with respect to F.
Let
be a function that is
, convex, and even. And
is a homogenous function, that is, for any
,
Furthermore, we assume for any ,
By the homogeneity property of
F, we can find two positive constants
such that
The operator
is called Finsler–Laplacian, which was studied by many mathematicians. For some important works involving the Finsler-Laplacian, we refer to [
22,
23,
24,
25,
26] and the references therein.
is the support function of
, which is defined as
, where
. Then we can check that
is also a function that is
. And
is also a convex and homogeneous function. What is more,
is dual to
in the sense that
Denote the unit Wulff ball of center at origin as
and
which is the volume of a unit Wulff ball
. Also, we denote
as the Wulff ball of center at origin with radius
r, i.e.,
For later use, by the assumptions of
, we can obtain some properties of the function
; see also [
25,
27,
28].
Lemma 1. We have
- (i)
;
- (ii)
, and for some and ;
- (iii)
for ;
- (iv)
, for ;
- (v)
for .
Now, we give the co-area formula and isoperimetric inequality with respect to
F, respectively. For a domain
,
, let
, which we denote as a function of bounded variation. The anisotropic bounded variation of
u with respect to
F is defined by
and the anisotropic perimeter of
G with respect to
F is defined by
where
is the characteristic function defined on the subset
G. Then we have the co-area formula (see [
26])
and the isoperimetric inequality
Furthermore, (
9) becomes an equality if and only if
G is a Wulff ball.
3. Anisotropic Moser–Trudinger-Type Inequality with Logarithmic Weight
In this section, we prove Theorem 1. Firstly, we give a useful formula involving the change in functions in a unit Wulff ball
. For
and any
, we let
Then we have the following lemma.
Lemma 2. Let with . Define v by (10); then we have . Proof. By the property of
in Lemma 1, we have
Hence, by the co-area Formula (
8), we have
□
Next, in this paper, we frequently need to change the variable in the following way.
For
, we change the variable as follows:
and set
Then we have
. By Lemma 1 and co-area Formula (
8), we can transform the norm as follows:
The functional changes as follows:
and
where
.
Now it is easy to prove Theorem 1 by Lemma 2.
Proof of Theorem 1. Let
with
. Define
v by (
10). By Lemma 2, we have
for
. By the definition of
, we obtain
Since (
15) holds for
with
, then we have
for any
. Hence, we obtain that the function
is decreasing on
. Thus, by the anisotropic Moser–Trudinger-type inequality (
3), we obtain
.
Now we prove the constant
is sharp. We need to show that, if
,
is infinite. By (
14), we only need to test
where
.
Consider the family of functions of Moser’s type
By direct computation, we have
. However, as
,
The proof of Theorem 1 is completed. □
4. Existence of the Extremal Function
In this section, we complete the proof of Theorem 2. Firstly, we give a uniform bound for
. For
, we denote by
the value of
with
. By the Hölder inequality and co-area Formula (
8), for any
and
, we have
In particular, when
, for any
and
, we have
The definition of
in (
11) and (
12) shows that the anisotropic norm changes as
and (
13) shows that the functional
and
changes as
For
, we define
Then the existence of an extremal function in Theorem 1 reduces to find
such that
Let
be a maximizing sequence of (
19), that is,
. Since
then there exist a subsequence (still denoted by
) and a function
such that
Next, we give an inequality and we will use it several times. For any
and
, by the Hölder inequality, we have
Now we give a lemma involving concentration-compactness alternative, by which we only need to prove that the maximizing sequence
in (
20) does not concentrate at 0, and then we can pass to the limit in the functional. Firstly, we give a definition.
We say a sequence of functions
concentrates at
, denoted by
if
and any
,
.
Lemma 3. [Concentration-compactness alternative] For any sequence , , such that in , then up to a subsequence (still denoted by ), either (i) , or (ii) concentrates at .
Proof. We assume that
does not hold; then we only need to show that
holds. Since
does not hold, then there exist
and
such that for sufficiently large
k,
where we use the variable of change
Since for any
k,
we have for
,
T sufficiently large,
We note that in (
23), we applied the inequality if
,
. Then, for
large enough,
.
We split the integral
, where
and
Since converges pointwise to , then also converges pointwise to v. Then, by and the dominated convergence theorem, we have that .
By (
23), we have for any small
and
T large enough,
which is smaller than
. Then
, that is,
. □
The following lemma is proved in [
29]. For
,
, let
Lemma 4 ([
29]).
For each and , we havewhere and . The inequality tends to an equality if , and . Let
such that
concentrates at 0, that is,
,
. Define
from
by the same transformation as in (
22). Then, since
concentrates at 0, we have that
in
and converges pointwise to 0.
Lemma 5. Let be as above. Then one of the following alternatives holds:
(i) We can find points such that (ii) If such does not exist, then What is more, if the first alternative (i) holds, we can find to be the first point in satisfying (26) and satisfying as . Proof. Since
, then if
,
. However, if
,
, then
, which implies that we cannot find
satisfying (
26) in
.
Now we assume
does not hold. Then we have
,
. Furthermore, we have
Define the dominating function as follows:
Then, by the dominated convergence theorem, we obtain that .
Let
hold. We choose the first
satisfying (
26). We now prove that
as
. For any large number
M, we need to prove that there exist
, such that for any
,
. Firstly, we choose
small, such that
Now, since
concentrates, we have for
and any
,
Then we obtain for any , . □
Now we define the concentration level at 0,
We can give the estimate for the concentration level.
Lemma 6. For , we have that Proof. To prove the lemma, it is sufficient to assume the sequences
satisfy the first alternative in Lemma 5, because if
satisfy the second alternative, we can obtain the inequality (
27) by Lemma 5.
Firstly, we show that
where
and
are as in Lemma 5. Since
and (
21), we have that
uniformly on compact subsets of
. Then for any
,
, we obtain
for
and
k large enough. By the property of
, that is, for
,
, we obtain
Now, as
and
, we have
Set
. Then, by (
21) with
and
, we have
Define the function
. Then
By
, we have
Now, applying Lemma 4 with
and
, we obtain
where
and
. Therefore, it is left to show that
We make use of the Maclaurin series expansion. Firstly,
for some positive constant
C, which depends only on
. Thus, we have
To estimate
, we first use the binomial expansion of
to obtain
. Now, using (
30) and (
33), we obtain
Then we have completed the proof of the Lemma. □
Proof of Theorem 2. We assume
does not converge to
, where
,
is as in (
20). Thus, by Lemma 6, we obtain
If we can find some
such that
then clearly
and thus, we obtain a contradiction.
Consider the function
as follows:
where
. It has been proved in [
29] that
and
Set
for
. Then
Now we can choose
sufficiently close to 1 such that
. Let us estimate the term
. Since
for
, we have
Now, by direct calculation, we obtain
and
where
. Note that by the above estimates, we have
Thus, we can choose , depending only on n, such that . Thus, we have finished the proof of the Theorem. □