Abstract

The purpose of this paper is to introduce new spaces and that consist of all sequences whose Riesz transforms of order one are in the spaces and , respectively. We also show that and are linearly isomorphic to the spaces and , respectively. The and duals of the spaces and are computed. Furthermore, the classes and of infinite matrices are characterized for any given sequence space and determine the necessary and sufficient conditions on a matrix to satisfy , for all .

1. Introduction and Preliminaries

Let be the space of all real or complex valued sequences. Then, each linear subspace of is called a sequence space. For example, the notations , , , , , and are used for the sequence spaces of all bounded, convergent, and null sequences, absolutely convergent series, convergent series, and bounded series, respectively. Let and be two sequence spaces and an infinite matrix of real or complex numbers , where . Then, defines a matrix mapping from to and is denoted by if for every sequence the sequence , the -transform of , is in where By , we denote the class of matrices such that . Thus, if and only if the series on the right side of (1.1) converges for each and every , and we have for all . The matrix domain of an infinite matrix in a sequence space is defined by If we take , then is called, convergence domain of , and we write the limit of as . Further is called regular if for each convergent sequence .

Let be a sequence space. Then is called solid if and only if , [1]. The approach constructing a new sequence space by means of the matrix domain of a particular limitation method has recently been employed by Altay and Başar [2], Başarir [3], Aydın and Başar [4], Kirişçi and Başar [5], Şengânül and Başar [6], Polat and Başar [7], and Malkowsky et al. [8]. Finally, the new technique for deducing certain topological properties, such as AB-, KB-, and AD-properties, and solidity and monotonicity, and determining the -, - and -duals of the domain of a triangle matrix in a sequence space is given by Altay and Başar [9].

Furthermore, quite recently, Kirişçi and Başar [10] introduced the new sequence space derived from the space of almost convergent sequences by means of the domain of the generalized difference matrix .

Define the sets and by

If , then is said to be almost convergent to the generalized limit . When , we write .

Lorentz [11] introduced this concept and obtained the necessary and sufficient conditions for an infinite matrix to contain in its convergence domain. These conditions on an infinite matrix consist of the standard Silverman Toeplitz conditions for regularity plus the condition . Such matrices are called strongly regular. One of the best known strongly regular matrices is , the CesΓ ro matrix of order one which is a lower triangular matrix defined by

for all .

A matrix is called the generalized CesΓ ro matrix if it is obtained from by shifting rows. Let . Then, is defined by

for all .

Let us suppose that is the set of all such matrices obtained by using all possible functions . Now, right here, let us give a new definition for the set of almost convergent sequences that was introduced by Butković et al. [12]:

Lemma 1.1. The set of all almost convergent sequences is equal to the set .

Other one of the best known regular matrices is , the Riesz matrix which is a lower triangular matrix defined by

for all , where is real sequence with and .

Let be a subset of . The natural density of is defined by

where the vertical bars indicate the number of elements in the enclosed set. The sequence is said to be statistically convergent to the number if, for every , (see [13]). In this case, we write . We will also write and to denote the sets of all statistically convergent sequences and statistically null sequences. The statistically convergent sequences were studied by several authors (see [13, 14] and others).

Let us consider the following functionals defined on :

In [15], the -core of a real bounded sequence is defined as the closed interval and also the inequalities (-core of -core of ) (-core of -core of ), for all , have been studied. Here the Knopp core, in short -core, of is the interval . In particular, when , since , -core of is reduced to the Banach core, in short -core, of defined by the interval .

The concepts of -core and -core have been studied by many authors [16, 17].

Recently, Fridy and Orhan [13] have introduced the notions of statistical boundedness, statistical limit superior (or briefly ), and statistical limit inferior (or briefly ), defined the statistical core (or briefly -core) of a statistically bounded sequence as the closed interval , and also determined the necessary and sufficient conditions for a matrix to yield -core-core for all .

Let us write

Quite recently, -core of a sequence has been introduced by the closed intervals and also the inequalities

have been studied for all in [18].

Definition 1.2. Let . Then, -core of is defined by the closed interval , where Therefore, it is easy to see that -core of is if and only if .

As known, the method to obtain a new sequence space by using the convergence field of an infinite matrix is an old method in the theory of sequence spaces. However, the study of the convergence field of an infinite matrix in the space of almost convergent sequences is new.

2. The Sequence Spaces and

In this section we introduce the new spaces and as the sets of all sequences such that their -transforms are in the spaces and , respectively, that is

With the notation of (1.2), we can write and . Define the sequence , which will be frequently used, as the -transform of a sequence , that is, If , which is CesΓ ro matrix, order 1, then the space and correspond to the spaces and (see [18]).

Suppose that . Then we have the following proposition.

Proposition 2.1. dir.

Proof. The proof is similar to the proof of Lemma 1.1 so we omit the details, (see [12]).

Consider the function , and define The function is a norm and is BK-space. The proof of this is as follows.

Theorem 2.2. The sets and are linear spaces with the coordinate wise addition and scalar multiplication that is the BK- space with the .

Proof. The first part of the theorem can be easily proved. We prove the second part of the theorem. Since (1.2) holds and and are the BK-spaces [1] with respect to their natural norm, also the matrix is normal and gives the fact that the spaces and are BK-spaces.

Theorem 2.3. The sequence spaces and are linearly isomorphic to the spaces and , respectively.

Proof. Since the fact β€œthe spaces and are linearly isomorphic” can also be proved in a similar way, we consider only the spaces and . In order to prove the fact that , we should show the existence of a linear bijection between the spaces and . Consider the transformation defined, with the notation of (2.2), from to by . The linearity of is clear. Further, it is trivial that whenever and hence is injective.
Let , and define the sequence by Then, we have which shows that . Consequently, we see that is surjective. Hence, is a linear bijection that therefore shows that the spaces and are linearly isomorphic, as desired. This completes the proof.

Theorem 2.4. The spaces and are not solid sequence spaces.

Proof. If we take and , then we see that and . Let be , that is, . Then, since and it is not hard to see by taking into account the definition Riesz matrix that . This shows that the multiplication of the spaces and is not a subset of and therefore the space is not solid. The proof for the space is similar to the proof of the space , so we omit it.

Theorem 2.5. Let the spaces and be given. Then,(1)the inclusion holds and the space is not a subset of the space ,(2) if and , then strictly holds.

Proof. (1) Clearly, the inclusion holds. Let us consider the sequence given by Since , the sequence is not a bounded sequence. But clearly . This shows that to us, the space is not a subset of the space .
(2) If and , then for all we have . Therefore, since , we see that .

In Theorem 2.6, we will use some similar techniques that are due to MΓ³ricz and Rhoades [19].

Theorem 2.6. Define the sequences and by for all . Then, for each and(i)the sequence is nondecreasing,(ii)the sequence is nonincreasing.

Proof. It is trivial that for each .
Since the part (ii) can be proved in a similar way, we prove only part (i) This step completes the proof.

Theorem 2.7. if and only if .

Proof. Suppose that . For each , choose to satisfy . We may write in a dyadic representation of the form , where each is 0 or 1, , and . Then, since , and hence Thus, If is the lower triangular matrix with nonzero entries , then, is a regular matrix so that . From the equality (2.12), we see that .
Conversely, assume that . Then, since implies we have If we take , then the proof of sufficiency is obtained. This step completes the proof.

3. Some Duals of the Spaces and

In this section, by using techniques in [9], we have stated and proved the theorems determining the - and -duals of the spaces and . For the sequence spaces and , define the set by With the notation of (3.1), the -, -, and -duals of a sequence space , which are, respectively, denoted by , , and , are defined by

The following two lemmas are introduced in [20] which we need in proving Theorems 3.3 and 3.4.

Lemma 3.1. if and only if

Lemma 3.2. if and only if

Theorem 3.3. The -duals of the spaces and are the set , where

Proof. Define the matrix via the sequence by for all . Here, . By using (2.2), we derive that From (3.7), we see that whenever if and only if whenever . Then, we derive by Lemma 3.1 that which yields the desired result .

Theorem 3.4. Define the set by Then, .

Proof. Consider equality (3.7), again. Thus, we deduce that whenever if and only if whenever . It is obvious that the columns of that matrix , defined by (3.6), are in the space . Therefore, we derive the consequence from Lemma 3.2 that .

In this section, we characterize the matrix mappings from into any given sequence space via the concept of the dual summability methods of the new type introduced by Başar [21].

Note that some researchers, such as, Başar [21], Başar and Γ‡olak [22], Kuttner [23], and Lorentz and Zeller [24], worked on the dual summability methods. Now, following Başar [21], we give a short survey about dual summability methods of the new type.

Let us suppose that the infinite matrices and map the sequences and , which are connected by the relation (2.2) to the sequences and , respectively, that is, It is clear here that the method is applied to the -transform of the sequence while the method is directly applied to the entries of the sequence . So, the methods and are essentially different.

Let us assume that the matrix product exists, which is a much weaker assumption than the conditions on the matrix belonging to any matrix class, in general. The methods and in (4.1), (4.2) are called dual summability methods of the new type if reduces to (or reduces to ) under the application of formal summation by parts. This leads us to the fact that exists and is equal to and formally holds if one side exists. This statement is equivalent to the following relation between the entries of the matrices and : for all .

Now, we give the following theorem concerning the dual matrices of the new type.

Theorem 4.1. Let and be the dual matrices of the new type and any given sequence space. Then, if and only if and for every fixed .

Proof. Suppose that and are dual matrices of the new type, that is to say (4.2) holds and is an any given sequence space. Since the spaces and are linearly isomorphic, now let and . Then, exists and , which yields that for each . Hence, exists for each , and thus letting in the equality for all , we have by (4.2) that , which gives the result .
Conversely, let for each and hold, and take any . Then, exists. Therefore, we obtain from the equality as that , and this shows that . This completes the proof.

Theorem 4.2. Suppose that the entries of the infinite matrices and are connected with the relation and is any given sequence space. Then, if and if only .

Proof. Let , and consider the following equality with (4.6): which yields as that whenever if and if only whenever . This step completes the proof.

Now, right here, we give the following propositions that are obtained from Lemmas 3.2 and 3.1 and Theorems 4.1 and 4.2.

Proposition 4.3. Let be an infinite matrix of real or complex numbers. Then,

Proposition 4.4. Let be an infinite matrix of real or complex numbers. Then,

Proposition 4.5. Let be an infinite matrix of real or complex numbers. Then,

Proposition 4.6. Let be an infinite matrix of real or complex numbers. Then,

5. Core Theorems

In this section, we give some core theorems related to the space . We need the following lemma due to Das [25] for the proof of next theorem.

Lemma 5.1. Let and . Then, there is a such that and

Theorem 5.2. for all if and only if and

Proof. Necessity: Suppose first that for all . If , then we have . By this hypothesis, we get If , then . So, we have , which implies that .
Now, let us consider the sequence of infinite matrices defined by Then, it is easy to see that the conditions of Lemma 5.1 are satisfied for the matrix sequence . Thus, by using the hypothesis, we can write This gives the necessity of (5.2).
Sufficiency: Conversely, let and (5.2) hold for all . For any real number we write and ; then and . Therefore, for any given , there is a such that for all . Now, we can write Thus, by applying and using the hypothesis, we have . This completes the proof since is arbitrary and .

In particular for all since is reduced to CesΓ ro matrix, see [18].

Theorem 5.3. for all if and only if and

Theorem 5.4. if and only if and for every with natural density zero.

Proof. Necessity: Let . Then, immediately follows from the fact that . Now, define a sequence for as where is any subset of with . Then, and , and so we have . On the other hand, since , the matrix defined by for all , must belong to the class . Hence, the necessity of (5.8) follows from Proposition 4.5.
Sufficiency: Conversely, suppose that and (5.8) holds. Let and . Write for any given so that . Since and , we have On the other hand, since condition (5.8) implies that Hence, ; that is, , which completes the proof.

Similarly, for all since is reduced to CesΓ ro matrix, see [18].

Theorem 5.5. if and only if and for every with natural density zero.

Theorem 5.6. for all if and only if and (5.2) holds.

Proof. Necessity: Let for all . Then, for all , where . Hence, since for all (see [13]), we have (5.2) from Theorem 5.2. Furthermore, one can also easily see that , that is, If , then . Thus, the last inequality implies that , that is, .
Sufficiency: Conversely, assume that and (5.2) hold. If , then is finite. Let be a subset of defined by for a given . Then it is obvious that and if .
For any real number we write and whence , and . Now, we can write By applying the operator and using the hypothesis, we obtained that . Since is arbitrary, we conclude that for all , that is, for all and the proof is complete. Now if for all , then is reduced to CesΓ ro matrix and we have and (5.7) holds, see [18].

Acknowledgment

The authors would like to thank the referees for much constructive criticism and attention for details.