1. Introduction
Let
be a bounded and open subset of
. We aim to prove the existence of multiple critical points, in a suitable sense, for a homogeneous Dirichlet problem associated with a functional of the form
under assumptions that do not guarantee any upper growth condition on the principal part
.
If
and
G are smooth and subjected to suitable growth estimates, the functional
f is of class
on some Sobolev space
, and standard variational methods apply (see e.g., [
1,
2]).
The case in which the growth conditions on
G are relaxed, meaning that
f is only continuous or even lower semicontinuous, has been already considered in [
3,
4], but standard growth conditions on the principal part
are still imposed.
However, situations in which there is no upper growth condition on the principal part appear, for instance, in continuum mechanics, and a case in which
is one-dimensional has already been treated in [
5]. On the other hand, to the best of our knowledge, in the multi-dimensional case, only the existence of minima has been proved thus far.
Let us also point out that the fact that each minimum satisfies the associated Euler–Lagrange equation can be not at all obvious. See, e.g., the survey paper [
6]. This problem has also been addressed in [
7], and the assumptions we will impose on
are related to those required in [
7].
In order to prove the existence of minima, the case in which the functional
f is coercive is usually considered. As a first step in the direction of existence results for critical points, we will also consider a coercive case. When standard growth conditions on
and
G are satisfied, the existence of multiple critical points in the coercive case has been obtained, for instance, in [
1,
8]. We will prove a result in the line of ([
1] Theorem 9.10) adapted to our setting.
More precisely, we assume that
satisfy
- (ΨG1)
for every , the function is measurable and, for a.e. , the function is convex;
for every , the function is measurable and, for a.e. , the function is of class ; we set ;
- (ΨG2)
for a.a. and, for every , there exist and such thatfor a.a.and all,
with,
;
- (ΨG3)
for every , there exists such thatmoreover, if , there exist and such that - (G4)
there exist and such thatmoreover, if , there exists also such that
Remark 1. According to ([9] Corollary 2.3), it follows that the function is continuous for a.a. . Moreover, we havewhere is given by assumption (ΨG2). Given
, we can define a functional
by
where we agree that
in the case
.
Remark 2. By standard results, the functionalis convex and lower semicontinuous on (see also the next Corollary cor:lsc), while the functionalis continuous on . However, it is not locally Lipschitz, unless , as we do not have a convenient estimate of . Let us point out that we need to consider the functional on a Lebesgue space such as and not, e.g., on , because is not assumed to be strictly convex and, consequently, it is impossible to prove a Palais–Smale condition related to a norm which requires the strong convergence of .
Definition 1. Let . A function is said to be an energy critical point
of the functional , ifand u is a minimum of the convex functionaldefined on the linear space Remark 3. According to Remark 1, we have that implies . Therefore, we have and .
Let us state our main result.
Theorem 1. Assume that and are even for a.a. and that there exists such that for a.a. and all with .
Then, for every , there exists such that, for every , the functional possesses at least m distinct pairs of energy critical points in with for all .
Since we are mainly interested in the principal part of the functional, in the next examples, we propose the same lower-order term, even if other choices are possible.
Concerning the principal part, since there is no upper bound on , one can consider, in particular, cases with nonstandard growth conditions.
Example 1. The assumptions of Theorem 1 are satisfied by a functional of the formwhere , are such thatand is convex, even and satisfiesLet us point out that, in the case , a possible choice iswith a very different behavior in the the two variables and . Let us also point out that, if , the functionalis continuous on , but not locally Lipschitz, unless further summability conditions on are imposed. Example 2. The assumptions of Theorem 1 are satisfied by the functionalswith as in the previous example and Principal parts of this form appear, for instance, in the study of strongly nonhomogeneous materials and non-Newtonian fluids (see, e.g., [10,11,12] and references therein). Concerning the first case, let us recall that, by Young’s inequality, one haswhenceand assumption (ΨG3) follows. Under a smoothness assumption on , we have that each energy critical point is also a weak solution of the associated Euler–Lagrange equation.
Proposition 1. Let and let be an energy critical point of . Assume also that, for a.e. , the function is of class .
In
Section 2, we recall the tools of nonsmooth critical point theory we need. In
Section 3, we adapt some basic results from [
13] to our setting. The main technical results are contained in
Section 4, where we show how the nonsmooth critical point theory can be applied to a functional such as
. Since we believe that the approach can be useful also when the functional is not coercive, in
Section 4, assumption (
G4) is replaced by more general hypotheses. Finally, in
Section 5, we prove the results stated in the Introduction.
In the following, we will denote by
the usual norm in
. For every
, we also set
2. Nonsmooth Critical Point Theory
In this section, we recall some useful tools. We refer the reader to [
5,
14,
15,
16,
17,
18] for proofs and more details.
Let
X be a metric space endowed with the distance
d. We denote by
the open ball of center
u and radius
. We will also consider the set
endowed with the distance
Let
be a function and let
When
f is real-valued and continuous, the next notion has been independently introduced in [
15,
16] and in [
18], while a variant has been developed in [
17]. By means of a suitable device, also the general case was considered in [
15,
16]. Here, we follow the equivalent approach of [
14].
Definition 2. For every with , we denote by the supremum of the σ’s in such that there exist and a continuous mapsatisfyingwhenever and . The extended real number is called the weak slope of f at u.
When f is real-valued and continuous, the next result provides a simple estimate.
Proposition 2. Let be a continuous function. Assume there exist , , and a continuous mapsatisfyingwhenever and . Then, we have .
Proof. It is a simple consequence of ([
14] Proposition 2.2). □
Remark 4. Let X be an open subset of a normed space and let be of class . Then, we have for all .
Definition 3. We say that is a (lower) critical point of f if and . We say that is a (lower) critical value of f if there exists such that and .
Definition 4. Let . A sequence in X is said to be a Palais–Smale sequence at level
c(-sequence,
for short) for f, if We say that f satisfies the Palais–Smale condition at level c(, for short), if every -sequence for f admits a convergent subsequence in X.
Now, let us see, following [
15,
16], how the case of a general
f can be reduced, to some extent, to the continuous case, taking advantage of the function
introduced in [
19].
Define a function
by
. It is readily seen that
is Lipschitz continuous of constant 1, whence
for all
.
Proposition 3. For every with , we have Proof. See ([
14] Proposition 2.3). □
We aim to reduce the study of a general f to that of the continuous function . In view of the natural correspondence , a key point is to have a control on pairs with .
Definition 5. Let . We say that f satisfies condition , if there exists such that Remark 5. If is continuous, then whenever .
Proof. See ([
16] Proposition 2.3). □
Several results of critical point theory can be extended to the nonsmooth case, by means of the previous concepts. In view of our purposes, let us mention an extension of D.C. Clark’s theorem (see [
1,
8] when
f is smooth).
Theorem 2. Let X be a Banach space, let be a lower semicontinuous function and let . Assume that:
- (a)
The function f is even and bounded from below;
- (b)
There exists an odd and continuous map ψ from the -dimensional sphere to X such that - (c)
For every , the function f satisfies and .
Then, there exist at least m distinct pairs of critical points of f with for all .
When dealing with the weak slope , an auxiliary concept is sometimes useful. From now on in this section, we assume that X is a normed space over and is a function.
The next notion has been introduced in [
14].
Definition 6. For every with , and , let be the infimum of r’s in such that there exist and a continuous mapsatisfyingwhenever and . Let us recall that the function is convex, lower semicontinuous and positively homogeneous of degree 1.
Definition 7. For every with , we set Remark 6. If f is convex, then agrees with the subdifferential of convex analysis. If f is locally Lipschitz, then and agree with Clarke’s notions [20], while, in general, , where denotes Clarke’s subdifferential. The subdifferential we have recalled is suitably related to the weak slope because of the next result.
Theorem 3. For every with , the following facts hold:
- (a)
;
- (b)
.
Proof. See ([
14] Theorem 4.13). □
4. The Variational Approach
Throughout this section, we consider two functions, and G, satisfying (ΨG1), (ΨG2) and (ΨG3). Moreover, we assume that:
- (G5)
If , there exist and such that
On the other hand, if
, it follows from (Ψ
G2) that, for every
, there exists
such that
In any case, the function
satisfies (
L1), (
L2) and, according to Corollary 1, we can define a lower semicontinuous functional
as in (
1), namely,
According to the Introduction, for every
, we set
Lemma 1. If , for every , there exist and such thatfor a.a and all and with . Proof. Let
. Since
from (Ψ
G3), we infer that
Combining this fact with (Ψ
G2), we deduce that there exist
and
such that
Since
there exist also
and
such that
and the assertion follows. □
Lemma 2. Let and let . Then, we have and, for everythere exists such thatwhenever , and . Proof. Assume, first, that
. From Lemma 1, we infer that
. Assume now, for a contradiction, that there exist
and
in
satisfying
whence
,
and
Then, from Lemma 1 and the (generalized) Fatou lemma, we infer that
and a contradiction follows.
In the case , the proof is similar, taking into account assumption (ΨG2). □
Theorem 5. For every with , the following facts hold:
- (a)
We have , and - (b)
If , then we have that , and u is a minimum of the convex functionaldefined on the convex set - (c)
If , then we have for all .
Proof. Let with . From assumptions (ΨG3) and (G5), it readily follows that and .
Actually, Lemma 2 yields
. Given
we infer from Lemma 2 that there exists
such that
whenever
,
and
.
On the other hand, also
satisfies (
L1) and (
L2), meaning we may assume that
whenever
and
by Corollary 1.
Taking into account the convexity of
, it follows that
whenever
,
and
. Given
, we may also assume that
. Then, if we set
it follows that
whenever
and
. According to Definition 6, we have that
and (
2) follows by the arbitrariness of
,
and
.
In particular, we have for all , which is dense in . Since is convex and lower semicontinuous in , we deduce that either for all or for all . From Theorem 3, we infer assertion (c), whence assertion (a), in the case .
To show assertion (
b) and complete the proof of assertion (
a), consider
. Since
from (
2), we infer that
, which implies that
, according to Remark 1.
Let with and , which is equivalent to .
Since
, we have that
Therefore,
and we can choose
as the test function in (
2), obtaining
On the other hand, we have
and
is convergent to
v in
. Combining the lower semicontinuity of
with Fatou’s lemma, we infer that
and the proof of assertion (
a) is complete.
Taking into account Definition 7, we deduce that
and assertion (
b) also follows. □
Proposition 4. Let with and let μ be in the dual space of .
Then, u is a minimum of the convex functionalon the linear space if and only if u is a minimum of the same functional on the convex set Proof. Similar to before, we also have
. Assume now that
u is a minimum of the convex functional
on the linear space
and let
with
and
.
Since
, we have
Similar to before, we also have that
whence
The converse is easily seen. □
Proposition 5. Let and let μ be in the dual space of . Assume that and that u is a minimum of the convex functionaldefined on . Suppose also that, for a.e. , the function is of class . Proof. Since
, we have
. Moreover, similar to before, we have
. Let now
with
and
. From Proposition 4, we infer that
and, for every
, we have
On the other hand, the convexity of
yields
Going to the upper limit as
, we infer that
In particular, the choice
yields
whence
for all
with
and
.
We can also choose as
v any element of
, which is a linear space. It follows that
and
for all
. □
For the last result of this section, we assume, more specifically with respect to (G5), that:
- (G6)
If , there exist and such that
Theorem 6. For every with , we have . In particular, for every , the function f satisfies condition .
Proof. Consider, first, the case in which
. Since
we have
and, on the other hand, similar to before,
Given
with
, let
be such that
If we set
, from Lemma 2, we infer that there exists
, with
and
, such that
whenever
,
and
. Taking into account assumption (
G6), we may also assume that
whenever
and
.
From the convexity of
, it follows that
whence
whenever
,
and
.
If now
and
, it follows that
In particular, if we set
for all
and
, we have
and
whence, by Proposition 2,
Going to the limit as , we infer that and the assertion follows in the case .
Consider now the other case, namely,
Given
, by Lemma 2, there exists
, with
, such that
whenever
,
and
. If
, we also have
whence
Then, arguing, similar to before, with
, we infer that
whenever
,
,
and
.
If now
and
, it follows that
In particular, if we set
for all
and
, we have
and
whence, by Proposition 2,
Going to the limit as , we infer that and the assertion follows also in this case. □