Abstract

Using Riemann-Liouville derivatives, we introduce fractional Sobolev spaces, characterize them, define weak fractional derivatives, and show that they coincide with the Riemann-Liouville ones. Next, we prove equivalence of some norms in the introduced spaces and derive their completeness, reflexivity, separability, and compactness of some imbeddings. An application to boundary value problems is given as well.

1. Introduction

Let . Classical Sobolev space of order one on an open bounded interval is defined by (cf. [1]) where is the space of functions that are integrable with power , is the set of smooth functions with compact support , and is the classical derivative of . The function satisfying the above condition is denoted by and called the weak derivative of of order one.

Sobolev space of order is defined by (cf. [1]) The spaces can be characterized as follows.

Theorem 1. Let . Then, if and only if there exist functions such that for and , where denotes the classical derivative of of order . In such a case there exists an absolutely continuous function such that a.e. on , which has absolutely continuous classical derivatives , the derivative , and

Remark 2. One shows that , , ( times).

Remark 3. In our paper, we will identify functions defined on and (, , resp.) that are equal a.e. on .

Each of the functions , , is denoted by and called the weak derivative of of order .

The space endowed with a norm where , has many useful properties such as completeness, reflexivity (for ), and separability. Moreover, some imbeddings of these spaces are compact (cf. [1]).

In the last years, many papers and books on fractional calculus and its applications have appeared. Most of them concern fractional differential equations, including calculus of variations and optimal control. In the classical (positive integer) case the fundamental role in this field is played by the mentioned Sobolev spaces. To our best knowledge, there are no “fractional” Sobolev spaces based on the notion of fractional derivative in Riemann-Liouville sense, which seems to be the most used in the theory of fractional differential equations. Our aim is to give some systematic basics for applications of fractional calculus to differential equations. More precisely, we extend the above definitions, with the aid of the Riemann-Liouville derivatives, to the case of noninteger positive (fractional) order , derive a fractional counterpart of Theorem 1, and prove the basic properties of the introduced spaces. When , the obtained results reduce to the classical ones. In the literature, some generalizations of Sobolev spaces to noninteger orders, on a domain , are known (cf. [2]): Gagliardo spaces , Besov spaces , and Nikolskii spaces . They have been introduced with the aid of approaches different from ours and their comparison with our spaces (in the case of ) is an open problem. Let us point that only Gagliardo spaces coincide with the classical Sobolev spaces when .

The paper is organized as follows. In the second section, we recall some basic notions and facts from the fractional calculus including a characterization of functions possessing the left (right) Riemann-Liouville derivatives. In the third section, we derive some special cases of the fractional theorem on the integration by parts. In the fourth section, we define the fractional Sobolev spaces of any order and characterize them. In the fifth section, we derive a fractional counterpart of Theorem 1, define the weak fractional derivatives of order , and show that they coincide with the Riemann-Liouville derivatives. In the sixth section, we introduce two norms in the fractional Sobolev spaces and prove their equivalence. In the seventh section, we derive completeness, reflexivity, and separability of the introduced spaces. In the eighth section, we prove compactness of some imbeddings. In the ninth section, we present some applications of the obtained results to fractional boundary value problems using a variational approach.

In the paper, we limit ourselves to the left fractional Sobolev spaces, but in an analogous way one can define right spaces and derive their appropriate properties.

2. Preliminaries

By , where , we denote the set of all functions that have a representative (a.e. on ) which is absolutely continuous together with its classical derivatives of orders . Of course, such a function possesses also the classical derivative of order , existing a.e. on and belonging to . These classical derivatives are the weak derivatives of . It is known (cf. [3, Lemma 2.4]) that if and only if there exist and such that In such a case,

By we denote the set of all functions that have representation (6) with and . It is known that for .

Let , and let . By the left Riemann-Liouville fractional integral of on the interval we mean (cf. [3]) a function given by

Remark 4. The above integral exists and is finite a.e. on .

In view of the convergence (cf. [3, Theorem 2.7]) it is natural to put

Remark 5. If , the right side of (8) exists (and is finite) everywhere on . So, can be defined everywhere on . In such a case it is continuous on .

Remark 6. It is easy to see that if is essentially bounded on and , then the right side of (8) is defined and bounded everywhere on . So, in this case can be defined everywhere on . From [3, Theorem 3.6] it follows that in such a case is equal everywhere on to a continuous function on . It is also known (cf. [3, Theorem 2.6]) that if with , then .

Let , where . By we denote the set of all functions that have the representation with and . Of course, .

We say that possesses the left Riemann-Liouville derivative of order , , on the interval if . By this derivative we mean the derivative . Of course, .

The next theorem can be deduced from [3, Corollary 2.1, Lemma 2.5 (b), Lemma 2.6 (b)] but, to our best knowledge, it has not been formulated by other authors. In [4] we give a direct proof of it in the case of , in [5]—when . In the case of it reduces to the theorem on the integral representation of type (6) of functions belonging to .

Theorem 7. If , , and , then has the left Riemann-Liouville derivative of order on the interval if and only if ; that is, has the representation (11). In such a case,

By () we denote the set of all functions possessing representation (11) with , .

Remark 8. It is easy to see that the above theorem implies the following one (for any ): has the left Riemann-Liouville derivative if and only if ; that is, has the representation (11) with .

Let . By the right Riemann-Liouville fractional integral of on the interval we mean a function Similarly, as in the case of left integral, we put

Remark 9. Clearly, has the properties analogous to those described in Remarks 5 and 6.

By we denote the set of all functions that have the representation with and . It is easy to see that .

We say that possesses the right Riemann-Liouville derivative of order , , on the interval if . By this derivative we mean the function . Of course, .

We also have the following.

Theorem 10. If , , and , then has the right Riemann-Liouville derivative of order on the interval if and only if ; that is, has the representation (15). In such a case,

By , , we denote the set of all functions possessing representation (15) with , .

Remark 11. As in the case of left Riemann-Liouville derivative, the following theorem holds true for any : has the right Riemann-Liouville derivative if and only if ; that is, has the representation (15) with .

In the next section, we will use the following two theorems (cf. [6, 7]).

Theorem 12. If , then if and only if or

Theorem 13. (a) If , , , and is of the form (11), then where is given by (b) If , , , , , and are of the form (11), then where

3. Some Special Cases of Integration by Parts

Below, is the set of continuous functions and is the set of functions such that , (by we mean the function ). Of course, these classical derivatives coincide with the weak ones .

Lemma 14. If , , , and then , Consequently, .

Proof. If and
then (cf. Remark 6 and [3, formulas (2.21) and (2.58)]) for all . This means (cf. Remark 6) that and . Equalities (28) imply also the equalities So (cf. Theorem 7), .

Remark 15. Analogous lemma can be obtained for the left integral and derivative (without the term ).

We have the following special case of the theorem on integration by parts.

Theorem 16. If , , then for any and satisfying boundary conditions

Proof. Let be of the form (11) with and . We have (Lemma 14 and Remark 9) Existence of the first integral and the first equality follows from Lemma 14; in the third equality we used an integral form of a classical fractional theorem on the integration by parts (cf. [3, formula (2.20)]), in the fourth equality we used continuity of and in the fifth equality we used the equalities The proof is completed.

The next theorem will be used in the last section of the paper. Proof of this theorem is contained in [4] and in [5] its extension to the case of fractional derivatives of higher order is derived (the method of the proof is the same in both papers).

Theorem 17. If , , , , and , then for and .

4. Fractional Sobolev Spaces

Let and let . By left Sobolev space of order we will mean the set given by A function given above will be called the weak left fractional derivative of order of ; let us denote it by . Uniqueness of this weak derivative follows from [1, Lemma IV.2 and Propositions IV.18, IV.21].

Let , , . By the left Sobolev space of order we mean the set given by

Remark 18. Since for , we see that the weak left fractional derivative of coincides with the classical weak derivative of . Consequently, for .

We have the following characterization of .

Theorem 19. If , , and , then

Proof. Case of follows from (37) and from the fact that . So, let us consider the case of . We will apply the induction with respect to .
Let . If , then from Theorem 7 it follows that has the derivative . Theorem 16 implies that for any . So, with
Now, let us assume that , that is, , and there exists a function such that for any . To show that it is sufficient to check (cf. Theorem 7 and definition of ) that possesses the left Riemann-Liouville derivative of order , belonging to , that is, that is absolutely continuous on and its classical derivative of the first order (existing a.e. on ) belongs to .
If , then (cf. Lemma 14) and . From the differential form of the classical fractional theorem on the integration by parts (cf. [3, formula (2.64)]) it follows that This means (cf. (41) and (42)) that for any . So, . Consequently, is absolutely continuous and its classical derivative is equal a.e. on to .
Now, let and assume that for any . We will show that this equality holds true for any .
Indeed, let . If , then and . So, where and . If , from the above formula it follows (cf. (40)) that If , from Theorem 13 (b) and (40) it follows that This means that and, consequently, So, in both cases (, ) there exist and such that To show that it is sufficient to check that Indeed, we have (below, we use the following elementary formula for , ) for a.e. Thus, .
Conversely, let . To show that we have to check that Theorem 13(b) implies the equality where with . So (cf. Theorem 12), . From Theorem 13 (b) it also follows that (existence of the weak fractional derivative and equality follow from the relation and (40)).

From the first part of the above proof (case of ) and from the uniqueness of the weak fractional derivative the following theorem follows (cf. also [4, fractional fundamental lemma]).

Theorem 20. If , then the weak left fractional derivative of a function coincides with its left Riemann-Liouville fractional derivative a.e. on .

Remark 21. If and , then and, consequently, . If , then is the set of all functions belonging to that satisfy the condition .

5. Weak Fractional Derivatives

Now, we will prove an extension of Theorem 1.

Theorem 22. Let , , , and . Then, if and only if there exist functions such that for any . In such a case there exist the left Riemann-Liouville derivatives of and

Proof. Case of follows from Theorem 1. So, let us consider the case of . We will apply the induction with respect to .
When , it is sufficient to use the definition of with and Theorem 20.
So, let and assume that the theorem is true for any . We will show that it is true for any .
Indeed, if with a fixed , then Relation (59) implies (by the induction assumption) the existence of functions such that (of course, ) for any and . Moreover, From the existence of and from (60) it follows that . In particular, there exists derivative . If we put we see that and for any (the second equality follows from Theorem 16).
Now, let us assume that there exist functions such that for any . Since , the above condition for and the induction assumption mean that
Moreover, condition (64) for , Lemma 14, and the integral form of the classical fractional theorem on the integration by parts (cf. [3, formula (2.20)]) imply that for any . This means that . So, if we fix and use once again Lemma 14 and the integral form of the classical fractional theorem on the integration by parts, we obtain Thus, from Theorem 1 it follows that Using (65) and (68) we assert that .

Functions will be called the weak left fractional derivatives of of orders , respectively. Their uniqueness follows from [1, Lemma IV.2 and Propositions IV.18,IV.21]. From the above theorem it follows that they coincide with the appropriate Riemann-Liouville derivatives.

We have the following counterpart of Remark 2.

Theorem 23. If , , and , then

Proof. If , then and from the definition of the Riemann-Liouville derivative and Theorem 22 it follows that The proof is completed.

6. Norms in

Let us fix where and consider in the space a norm given by (here denotes the classical norm in ).

We have the following theorem.

Theorem 24. If and , then the norm is equivalent to a norm given by

Proof. We will use the induction with respect to .
Assume that and . Then, for given by with and , we have where (cf. [3, formula (2.72)]). Thus (of course, and ), where
Consequently, where .
Now, we will prove that there exists a constant such that Indeed, let and consider a coordinate function of with a fixed . The mean value theorem implies the existence of such that From the absolute continuity of it follows that for any . Consequently, for . In particular, So, where . Thus, and, consequently, where .
When , then (cf. Remark 21) is the set of all functions belonging to that satisfy the condition . Consequently, in the same way as in the case of (putting ) we obtain the inequality with some . The inequality with some is obvious (it is sufficient to put and use the fact that ).
Now, let us assume that the assertion holds true for some . We will prove that it is true for .
Let . We know that there exist constants and such that for any . If , then and therefore (from the induction assumption) Since (because of ), therefore (cf. case of ) where is such that for . Thus, where (we used the equality ).
On the other hand, since , therefore (from the induction assumption) So, Let us recall that , so , , and . Thus, where is such that Consequently, where .

7. Basic Properties of

Now, we are in a position to prove some basic properties of the introduced spaces.

Theorem 25. The space is complete with respect to each of the norms and , for any and .

Proof. Let be such that . Of course, it is sufficient to show that with the norm is complete. Let be a Cauchy sequence with respect to this norm. So, the sequences , , are Cauchy sequences in and is the Cauchy sequence in . Let and be limits of the above sequences in and , respectively. Then the function belongs to and is the limit of in with respect to (to assert that it is sufficient to consider the cases and —in the second case for any and, consequently, ).

In the proofs of the next two theorems we use the method presented in [1].

Theorem 26. The space is reflexive with respect to each of the norms and , for any and .

Proof. Let us consider with the norm and define a mapping Since is the isometry, is the closed linear subspace of the reflexive space . So (cf. [8, Corollary 1 in Part V.73]), it is reflexive and, consequently, is reflexive with respect to .
From the equivalence of the norms it follows that with the norm is also reflexive (it is sufficient to consider the identity mapping being the linear homeomorphism and use [8, Remark in Part V.7.3]).

Theorem 27. The space is separable with respect to each of the norms and , for any and .

Proof. Let us consider with the norm and mapping defined in the proof of Theorem 26. Of course, is separable as a subset of separable space . Since is the isometry, is also separable with respect to the norm . Equivalence of the norms and implies separability of with respect to .

8. Imbeddings

We have the following extension of Theorem 12.

Theorem 28. (a) If , then for and ; consequently, for and .
(b) If , , , and , then for and ; consequently, for and .
(c) If , , , , and , then for and ; consequently, for , , , and or , , and .
(d) If , then for ; consequently, for .

Proof. (a) Let us fix . Theorems 12 and 13 (a) imply that provided that and (cf. (20), (21)); that is, and . Consequently, for such and ,
(b) Let us fix . In this case and, consequently, for any . So (cf. Theorems 12 and 13 (a)) provided that and (cf. (20), (21)); that is, . Consequently, for such and , (we used here the inclusion ).
(c) Let us fix . Theorems 12 and 13 (b) imply that for any (cf. (22), (24)).
If , then (cf. Remark 21) . So, provided that (cf. Theorem 13 (b)) . If , then (cf. Remark 21) is the set of all functions belonging to that satisfy the condition . Consequently (cf. Theorem 13 (b)), , that is, for any .
(d) These facts are obvious.

The first part of the point (c) of the above theorem implies the following.

Corollary 29. If , , , , and , then for .

In the next section we will use the following important result obtained in [9, Lemma 1.1].

Theorem 30. If and , then the operator is completely continuous, that is, it maps the bounded sets onto relatively compact ones.

Now, we are in a position to prove theorems on compactness of some imbeddings.

Theorem 31. The imbedding for , , and , given in Theorem 28 (a), is compact.

Proof. Let , , , and , where be a bounded sequence in . We will show that it contains a subsequence which is convergent in .
Since is bounded (cf. Theorem 24), the sequences and are bounded in and , respectively. So, one can choose a subsequence of positive integers such that is convergent to some in and (cf. Theorem 30) such that is convergent to some in . Of course, is convergent to in . Moreover, the sequence converges in to . This means (cf. (20), (21)) that the sequence converges in to given by where . The proof is completed.

The above theorem implies the following.

Corollary 32. The imbedding for , , or , and is compact.

Theorem 33. The imbedding for , , , , , and , given in Theorem 28 (b), is compact.

Proof. Let us consider a bounded sequence in where with and . We will show that it contains a subsequence which is convergent in .
Since is bounded, the sequences and are bounded in and , respectively. So, one can choose a subsequence of positive integers such that are convergent to some in and (cf. Theorem 30) such that the sequence is convergent in to some where . Consequently, is convergent in to and in , because So, the sequence , where converges to the function
in . Thus, the sequence converges in to .
The proof is completed.

The above theorem implies the following.

Corollary 34. The imbedding for , , , , or , , , , and is compact.

Theorem 35. The imbedding for , , , , , and , given in Corollary 29, is compact.

Proof. Let us consider a bounded sequence in where with and . We will show that it contains a subsequence which is convergent in .
Since is bounded, the sequences and are bounded in and , respectively. So, one can choose a subsequence of positive integers such that are convergent to some in and (cf. Theorem 30) such that are convergent to some in and, consequently, in . So (cf. Theorem 13 (b)), in , where is given by This means (cf. (22), (23), and (24)) that the sequence converges in to given by
where
The proof is completed.

The above theorem implies the following.

Corollary 36. The imbedding for , , , and is compact.

Remark 37. It is easy to see that imbeddings for and , given in Theorem 28 (d), are not compact.

Remark 38. From Corollary 32 it follows that the imbedding is compact for any . Corollary 34 implies the compactness of the imbedding for any , , and . The following problem is open: is it possible to strengthen Theorem 31 or Theorem 33 to deduce the compactness of the imbedding for any ?

9. Application to Boundary Value Problems

In this section, we will demonstrate an application of the obtained results to fractional boundary value problems.

Namely, let us fix and consider the following problem: where . By a solution to this problem we mean a function such that and exist, and satisfying the above equation and boundary conditions. Under the assumption on the boundary conditions make sense ([10, Property 4]).

Let , where to be in the following bilinear form It is easy to see that is the closed subspace of the Hilbert space (, ); it is sufficient to observe that it is closed in with respect to and to use Theorem 24. Of course, is a scalar product in and the norm generated by is simply the norm restricted to . Clearly, is continuous and coercive; that is, there exists a constant such that for ; in fact for . So, Lax-Milgram theorem [1] or simply Riesz-Frechet theorem implies that there exists such that Condition (135) means that From a counterpart of Theorem 20 for the weak right fractional derivative it follows that and Since , . Moreover, applying Theorem 17 to the left side of (137) we obtain for . So, This means that . Indeed, it is sufficient to consider functions , , of the form where , with nonzero th coordinate function (of course, and ).

Remark 39. In the same way one can prove the existence of a solution to system with boundary conditions or in the space . In [4], system (142) with nonhomogeneous boundary conditions where , is investigated using the Stampacchia theorem (cf. [1]). Nonlinear system of the form where is the gradient (in ) of a potential , with the above boundary conditions can be studied using the direct method of calculus of variations (cf. [11] for the case of (144)).
It is worth noting that if , , and , then . It follows from the integral representation of and from the fact that is Hölder continuous on and when and (cf. [10, Property 4]). So, it is natural to put in such a case . It seems that the most accurate functions for investigating the above systems with boundary conditions involving condition (or more general ), in general case of , are the functions possessing the fractional derivatives in Caputo sense; on such functions one assumes that they are absolutely continuous on and, consequently, condition makes sense. To our best knowledge, fractional Sobolev spaces via Caputo derivatives have not been investigated up to now and are an open problem.

Remark 40. From the condition (136) it follows that one can search approximate solutions to (132) using numerical methods, for example, the gradient or projection of gradient methods applied to the functional defined on and , respectively.

Conflict of Interests

The authors declare that there is no conflict of interests regarding the publication of this paper.

Acknowledgment

The project was financed with funds of National Science Centre, granted on the basis of decision DEC-2011/01/B/ST7/03426.