1. Introduction
1.1. Preliminaries
A priori given a partial differential equation on a bounded domain, be it linear or non-linear, we might end up obtaining solutions which also happens to be critical points of certain functionals defined on some appropriate Sobolev Space.
Suppose, we consider the following boundary value problem on some bounded domain
as follows:
Where,
denotes the
boundary of
in
. A priori from the boundary condition, it only suffices to search for
weak solutions of (
1) in the Sobolev Space
, the later assertion follows from the fact that, not all functions in
is smooth (i.e.
).
Definition 1.
(Weak Solution) A function is defined to be a weak solution of (1) if,
In case when
u is a
classical solution, i.e., in other words, when,
and satisfies (
1) point-wise, then, we can indeed obtain (
2) by multiplying (
1) by
and integrating by parts.
Remark 1. (2) is valid only when, .
Define a functional,
as,
It can be deduced that,
J is in fact a
-function, and,
is a
bounded linear functional having the following expression,
Hence, we can infer that,
is a
weak solution of (
1) ⇔
.
1.2. Conditions on a -Function Defined on a Real Banach Space
Given a real Banach Space X and a -function, , our primary objective in this section shall be to obtain certain conditions on I in order to ensure that, ∃ satisfying, .
The answer is quite simple for the one-dimensional case. Consider the example when, . Correspondingly, choose any -function, . Then, for every with, , if ∃ satisfying, , it follows that, ∃ such that, .
Although, for higher dimensional cases, it’s rather difficult. For example, if we consider,
defined as,
. We can deduce that, for any
and
in
with
,
and,
. This implies that, the line,
separates the points
where,
But, an important observation that, helps us conclude that, ∄ any critical point for I.
1.3. Palais-Smale Condition
Definition 2.
(Palais-Smale Condition) Given a Banach Space X and, , we define the functional I to satisfy thePalais-Smale Condition, if every sequence in X for which is bounded and contains a convergent subsequence.
We can infer about the above problem as discussed in the Section 1.2 in case for an arbitrary Banach Spaces using a famous result by Ambrosetti and Rabinowitz.
Theorem 1. (Mountain-Pass Theorem) Given a Banach Space X and, , assume that, I satisfies the Palais-Smale Condition. Furthermore, suppose, and, satisfying, and, . Then, ∃ such that, and, .
Remark 2.
In other words, the theorem implies that, if a pair of points in the graph of I are indeed separated by a mountain , then I has a critical point.
Remark 3. One of our aims in this article is to present aproof of the Mountain-Pass Theorem (1), as well as applying the statement of the same in order to look for non-negative weak solution to the non-linear boundary value problem (1) over a bounded domain , being continuous.
1.4. Differentiation in Banach Spaces
Suppose, we choose any two real Banach Spaces X and Y, and, denotes the space of all bounded linear operators from X to Y. Let us further denote, . Moreover, let be open.
Definition 3.
Let, . A function, is defined to be Frechet Differentiable (or, just differentiable) at , if ∃ a bounded linear operator, , in other words, satisfying,
Example 1.
For any differentiable function, at some , we have, . Correspondingly, theFrechet Derivative, of f at the point has the following expression, , .
In general, in case when, , and, be differentiable at , then we define, via the canonical isomorphism, given by, .
For arbitrary real Banach Spaces
X and
Y, if
be
Frechet Differentiable at some
, then, corresponding to the
Frechet Derivative of
I at
defined as,
satisfying (
5), we write
to be the
derivative of
I at
.
Definition 4.
A function, is defined to be differentiable on A if it is differentiable at every point of A. Furthermore, if, is continuous.
Example 2.
Suppose, be a Hilbert Space and, defined as, . Then, such that,
-
Suppose, . Define, as,
We can in fact conclude that, . Furthermore,
We can comment on some important properties of the derivative as follows.
(Chain Rule) A priori given
and
Z, and,
,
being non-empty
open sets, suppose,
and
be functions satisfying,
. If,
I and
J are differentiable at
and
respectively, then,
is differentiable at
, and,
(Mean Value Theorem) A priori given a differentiable function,
and,
, let us define,
to be a
line segment in
A.. Then,
-
(Taylor’s Formula) A priori given a function,
being differentiable at some
, let us define,
Then, by (
5), the
Remainder Term,
satisfies,
Theorem 2. Suppose, be bounded and open, and, . Moreover, let be a -function satisfying,
for some constants . Also,
Then, , and, for every ,
Corollary 1.
Suppose, for any continuous function, satisfying,
for every, . Moreover, let, be the primitive of f. If, has the following definition,
Then, , and,
2. Critical Points
In this section, we assume everywhere unless mentioned otherwise, that, X is a Banach Space, and .
Definition 5.
(Minima of a Function) Given a function, , we define a point to be a local minimum (resp. strictly local minimum) of I if, ∃ of in X satisfying,
On the other hand, is defined to be a global minimum of I if,
Therefore, we write, .
Definition 6.
(Maxima of a Function) Given a function, , we define a point to be a local maximum (resp. strictly local maximum) of I if, ∃ of in X satisfying,
On the other hand, is defined to be a global maximum of I if,
Therefore, we write, .
Remark 4.
Assume to be open, and be differentiable on A. Then,
provided, is a local minima (or, a local maxima) of I.
Definition 7.
(Critical Point) Suppose, be open, and be differentiable on A. A point is defined to be acritical pointof I if, . Subsequently, is called the critical value.
It is evident from the definition that, local minima and local maxima of a differentiable function I on some subset are indeed critical points of I. However, the same cannot be infered about the converse of this result.
For example, we take,
,
. Then,
, and,
. Thus,
is a
critical point of
I, although, is neither a
local minima nor a
local maxima of
I. It follows from the fact that,
Remark 5. Critical points satisfying the condition as above as termed as Saddle Points.
Definition 8.
(Saddle Point) Suppose, be open, and be differentiable on A. We define a critical point of I to be a saddle point of I if, for every neighbourhood of , ∃ satisfying,
Definition 9.
(Convex Function) A function is defined to be convex , if ,
Example 3. Consider, . Then, the function, defined as, is convex everywhere on .
Remark 6.
In case when, I is convex on X, the critical points of I are in fact global minima.
Proposition 1.
For a convex and differentiable function ,
Proposition 2.
Suppose, be a differentiable function satisfying the following condition,
Then, I is convex.
Proof. To prove the result, we’ll use the definition (9) of convexity.
Given that for every .
For , consider .
Combining these inequalities:
Since , . Hence, is convex.
□
Before we investigate the existence of critical points of a function, one needs to justify that, the definitions (5) and (6) of minima and maxima respectively are well-defined.
Definition 10.
A priori, given X to be a Hausdorff topological space, a function, is defined to be lower semi-continuous if , the set is closed.
Proposition 3.
-
(i)
Every continuous function is lower semi-continuous.
-
(ii)
If is open, then, is lower semi-continuous.
Proof.
-
(i)
-
We intend to show that for any continuous function , where X is a topological space, the set (say) is open for every . This implies that for any point in X, there exists a neighborhood U of such that for all x in U.
Let be arbitrarily chosen. For any point , since , by the continuity of f, ∃ a neighborhood of such that .
Thus, for any , there exists a neighborhood of contained in A, implying A is open.
Therefore, every continuous function is lower semi-continuous.
-
(ii)
-
To prove that the indicator function of an open set is lower semi-continuous, it suffices to show that for any , and for any , ∃ a neighborhood U of such that for all .
Formally, let
be the indicator function of
A, defined as:
Let be arbitrary, and let be given. Since A is open, ∃ a neighborhood U of contained in A. For any , , so . Since , for all .
Therefore, of an open set is indeed lower semi-continuous.
□
Remark 7. A function which is lower semi-continuous may not be continuous. For example, we can consider , and, for any with . Then, is indeed lower semi-continuous on (using above proposition), although it is not continuous at the points a and b in .
Theorem 3.
A priori given X to be a compact and Hausdorff topological space. Furthermore, be lower semi-continuous. Then, I is bounded below, and, such that,
Proof.
I being lower semi-continuous implies that, the set, is open. Moreover, , and, is compact. Hence, we must have, for some .
As a consequence, . Thus, it establishes that, X is bounded below.
Let us choose, . We claim that, such a do exists. If not, then, we consider the collection where, for every such that,
A priori from the compactness condition of X, satisfying, . In other words, the collection, indeed is a finite subcover of X for some .
Subsequently, , , a contradiction to the fact that, . □
For the ease of computation, we introduce the notion of sequentially lower semi-continuous.
Definition 11.
Given a Hausdorff topological space X, a function, is defined to be sequentially lower semi-continuous if, for every sequence tending to x in X,
Proposition 4. For a Hausdorff topological space X and a function , the following holds true:
- (1)
I is lower semi-continuous ⇒I is sequentially lower semi-continuous.
- (2)
Converse holds true only if X is a metric space.
Proof.
- (1)
-
Given to be lower semi-continuous. Suppose, be a sequence in X converging to x, and suppose . We wish to show that .
Since I is lower semi-continuous, for any , ∃ a neighborhood U of x such that .
Since , ∃ such that ∀, .
Thus, for every , .
Since
was arbitrary, we conclude:
Therefore, I is sequentially lower-semi continuous, and the proof is thus complete.
- (2)
-
Assume be a sequentially lower semi-continuous function, and let be arbitrary. It suffices to establish that, the set is open in X.
Let be any point in the set, say, , i.e., .
Since I is sequentially lower semi-continuous, for any sequence in X converging to , we have .
Since is in the set , we have .
As a result, ∃ such that ∀, [ ∵ the limit inferior of a sequence is the greatest lower bound of the set of subsequential limits ].
Therefore, for any sequence in X converging to , ∃ a neighborhood U of such that for all , which implies that is open in X.
Since was arbitrary, this holds for all . Therefore, we can conclude that, I is lower semi-continuous.
□
Remark 8. A sequentially lower semi-continuous function on a non metrizable space may not be lower semi-continuous. Consider the example where, is equipped with the cofinite topology.
Define the function as follows:
Consider any sequence in X converging to x. Since every neighborhood of x in the cofinite topology contains all but finitely many points of X, equals x for all sufficiently large n. Therefore, , which makes sequentially lower semi-continuous.
Let’s examine the point . The set is the set of irrational numbers in X, which is not open in the cofinite topology because it contains infinitely many points. Hence, is not lower semi-continuous.
Proposition 5.
Given a Hausdorff topological space X and, , suppose that, for every ,
Then, I is bounded below, and, ∃ satisfying, .
Proof. To prove this statement, we’ll use the fact that compactness of the level sets of I implies certain properties about I.
Suppose I is not bounded below. Then, for each , ∃ such that . However, this implies that the set is nonempty and contains the sequence , but it cannot be compact as the sequence has no convergent subsequence due to the unboundedness of I. This contradicts the assumption that all such sets are compact. Hence, I must be bounded below.
Since I is bounded below, let . Then, for each , ∃ such that . Consider the sequence . Since is compact, there exists a subsequence converging to some in X. By continuity of I, . But since , we have . Conversely, since c is the greatest lower bound of I, we have . Therefore, , and is the desired point. □
Remark 9. Consider, . Define a function, as, . We can observe that, I is indeed smooth and bounded below, but I does not achieve its infimum. Applying Theorem (3) and Proposition (4) in order to obtain infimum, a compactness condition either for the space or, for the function must be required.
In case for infinite dimensional Banach Space X, the compactness condition is not achieved under the norm topology. However a certain amount of compactness is achieved to ensure the attainment of the infimum can be obtained in weaker topology on X.
In the next section, we shall learn more about a weaker topology called weak topology on X as compared to the norm topology defined on it.
5. Applications to the Critical Point Theory
Theorem 15.
Suppose, , and, . Also, we consider to be a bounded domain. Then which is in fact a weak solution of the problem,
Proof. We define,
as,
Clearly, we can verify that,
I satisfies
. Next we check the conditions for the
Mountain Pass Theorem. We obtain,
SInce,
, hence,
for some
. Therefore,
For
, we have,
Since,
, thus choosing
R small enough such that,
Let us take any
. Thus,
for any
. Now, from the fact that,
Choosing t large enough such that, and, , and, , and applying the Mountain Pass Theorem, we can assert that, and, .
It can also be derived that, , as, , and the proof is thus complete.
□
On the other hand,
Brezisand
Nirenberg [
28] has developed signifcant results in observing the model problem:
When and, be any real constant. We can consider the following cases.
Theorem 16. In case when, , the problem (41) indeed has a solution for every , where, denotes the first eigenvalue of .
Moreover, the problem does not admit any solution if and, Ω is starshaped.
Theorem 17. For , the problem (41) turns out to be much more delicate. In this scenario, a complete solution exists only if Ω is a ball. Subsequently, we shall have that, (41) yields a solution iff , being the first eigenvalue of .
Remark 15. In case when , Brezis and Nirenberg [28] discusses the concept of commenting on the results related to existence of solutions to (41) using the notion of general Bifurcation Theory. For example, as mentioned by Rabinowitz [29], the problem (41) possesses a component of solutions , which meets and which is unbounded in . Furthermore, if and , applying the result in Theorem (16), we can conclude that, the projection of on the λ-axis does in fact contain the interval (with appropriate modifications being done when ).
As in another scenario when, and Ω is star-shaped, then the problem (41) has no solution for , being some positive constant depending on Ω and p. This was explicitly derived by Rabinowitz [30] for the case when, . One can in fact use similar argument in the general case by applying Pohozaev’s Identity.
In greater generality as compared to the Dirichlet boundary value problem (
41),
Brezis and
Nirenberg [
28] also have dealt with the following problem in detail:
Where,
and,
,
being a constant. Considering different cases for
n, we can conclude about the existence of solution (if any) for (
42) in the manner as described below.
Theorem 18. For , the problem (42) indeed admits of a solution for every .
Theorem 19. In case when and consequently, , we can assert the following about existence of solution to the DIrichlet problem (42):
-
(i)
If , ∃ solution to the problem for every .
-
(ii)
For , solution does exists only for sufficiently large values of λ.
Brezis Theorem allows us to conclude that, for any function which is bounded below, and satisfy , satisfying, .
As for another application of Theorem (12), we next justify the existence of a critical point for I in case it is bounded below on a finite dimensional subspace of X.
Theorem 20.
(Saddle Point Theorem (Rabinowitz) [1]) Assume X to be a Banach Space and, satisfies . Also suppose, be a finite dimensional subspace of X and, . Furthermore, we consider that, such that,
Then, satisfying, . Moreover, the critical value, , which in fact, can be characterized as,
Proof of the above Rabinowitz Theorem requires an application of the Topological Degree in .
Definition 15.
(Topological Degree) Suppose, be a bounded and open set. Given and, , the topological degree (or, Brouwer Degree), is defined to be an integer satisfying the following properties:
-
(I)
-
(II)
such that, .
-
(III)
if, .
-
(IV)
(Addition-Excision Property) If are open with and, , then,
-
(V)
If and, be continuous, and moreover, , then, is independent of t.
-
(VI)
whenever, .
-
(VII)
(Product Property) If ’s are bounded open sets in for every , and, and are such that, , . Then,
Remark 16.
Given to be bounded and open, and, , we can relate the theory of the Brouwer Degree with the existence and multiplicity of solutions of the equation,
Assuming to be non-singular whenever (45) holds true. Then, the Inverse Function Theorem yields, (45) can have only a finite number of solutions in Ω. In this so called "nice" case, the corresponding Brouwer Degree of φ with respect to Ω and p, denoted by has the following expression,
where, denotes the determinant of a square matrix A.
Remark 17. The notion of the Brouwer Degree can also be extended from "regular" to "singular" values of -functions, and then to continuous functions on [ref. [27].
Remark 18.
The definition of topological degree as provided above as well as its properties can in fact be extended to an infinite dimensional space, in which case it is kown as Leray-Schauder Degree [ref. [27].
Proof of Theorem (20):
Proof. For any
and
, we define the following sets,
and,
First, we intend to prove that,
for
c as mentioned in (
44). Assume if possible that,
in this case. We choose
as,
and, consider
such that,
Such that,
condition is satisfied under these assumptions. Let,
be a
I-decreasing homotopy (ref. Corollary
[
25]) satisfying the following conditions,
if, .
, where, .
We denote,
. If
, then by (
47) we derive,
which contradicts the definition of
c.
Let us establish that, . Consider such that, and, . We thus have, and, . Hence, if, .
If (
48) does hold true, then, for
, using (
43) and (
46),
This helps us assert that, .
Therefore, it only suffices to show (
48). Wednote
and
to be the projections of
X onto
V and
E respectively. Furthermore, we identify
V with
for some
n. For
,
. Hence, by properties
and
in the definition (15) of the
Topological Degree, we obtain,
Applying property
in definition (15),
satisfying,
. Consequently, for each
,
such that,
On the other hand, from (
43), we can conclude,
Using (
44), it follows that,
, and the proof is thus complete. □
Remark 19. Heuristically speaking, in the above Theorem (20), c is the minimax of I over all surfaces modelled on , sharing the same boundary. Unlike the Mountain-Pass Theorem, in applications of the Saddle Point Theorem, in general, no critical points of I are known initially. Important to note that, the condition (43) are satisfied if I is convex on E, concave on V, and appropriately coercive.
Another version of the
Rabinowitz Saddle Point Theorem (Theorem (20)) can be found in [
26].
Theorem 21.
Given a real Banach Space X having the following direct sum decomposition, , where V and E are closed subspaces with . Suppose, satisfy condition, and,
Define, with R is chosen so large that, and ⇒.
and, ∃ satisfying, and, .
Remark 20. For other versions of proof of Theorem (20) and, other important applications of the Mountain Pass Theorem, one can refer to [1,11,15].
Remark 21. Theorem (20) essentially states that under certain conditions, a functional (a function of functions) will have at least one critical point that is a saddle point. This critical point is where the functional doesn’t increase or decrease, representing a sort of equilibrium.
To put it in a more formal geometric context, consider a functional defined on an infinite-dimensional space. The space can be thought of as a “landscape” of all possible functions. The functional assigns a “height” (or value) to each function in this landscape. Rabinowitz’s theorem guarantees that there’s at least one function in this landscape that has a saddle point: it’s not the highest or lowest point, but it’s a point of balance between different “directions” in the function space.
This geometric interpretation is quite abstract because we’re dealing with spaces that are not easy to visualize. However, the concept of a saddle point as a point of equilibrium remains a powerful image to understand the essence of the theorem.