Volume 2013, Issue 1 687393
Research Article
Open Access

On the Fine Spectrum of the Operator Defined by the Lambda Matrix over the Spaces of Null and Convergent Sequences

Medine Yeşilkayagil

Medine Yeşilkayagil

Department of Mathematics, Uşak University, 1 Eylül Campus, 64200 Uşak, Turkey usak.edu.tr

Search for more papers by this author
Feyzi Başar

Corresponding Author

Feyzi Başar

Department of Mathematics, Fatih University, The Hadimköy Campus, Büyükçekmece, 34500 İstanbul, Turkey fatih.edu.tr

Search for more papers by this author
First published: 28 March 2013
Citations: 13
Academic Editor: Pavel Kurasov

Abstract

The main purpose of this paper is to determine the fine spectrum with respect to Goldberg′s classification of the operator defined by the lambda matrix over the sequence spaces c0 and c. As a new development, we give the approximate point spectrum, defect spectrum, and compression spectrum of the matrix operator Λ on the sequence spaces c0 and c. Finally, we present a Mercerian theorem. Since the matrix Λ is reduced to a regular matrix depending on the choice of the sequence (λk) having certain properties and its spectrum is firstly investigated, our work is new and the results are comprehensive.

1. Introduction

Let X and Y be Banach spaces, and let T : XY also be a bounded linear operator. By R(T), we denote the range of T; that is,
()
By B(X), we also denote the set of all bounded linear operators on X into itself. If X is any Banach space and TB(X) then the adjoint T* of T is a bounded linear operator on the dual X* of X defined by (T*f)(x) = f(Tx) for all fX* and xX.
Let X ≠ {θ} be a nontrivial complex normed space and T : D(T) → X a linear operator defined on a subspace D(T)⊆X. We do not assume that D(T) is dense in X or that T has closed graph {(x, Tx) : xD(T)}⊆X × X. By the statement “T is invertible,” it is meant that there exists a bounded linear operator S : R(T) → X for which ST = I on D(T) and , such that S = T−1 is necessarily uniquely determined and linear; the boundedness of S means that T must be bounded below, in the sense that there is M > 0 for which ∥Tx∥ ≥ Mx∥ for all xD(T). Associated with each complex number, α is the perturbed operator
()
defined on the same domain D(T) as T. The spectrum σ(T, X) consists of those α, the complex field, for which Tα is not invertible, and the resolvent is the mapping from the complement σ(T, X) of the spectrum into the algebra of bounded linear operators on X defined by .

2. The Subdivisions of Spectrum

In this section, we define the parts of spectrum called point spectrum, continuous spectrum, residual spectrum, approximate point spectrum, defect spectrum, and compression spectrum. There are many different ways to subdivide the spectrum of a bounded linear operator. Some of them are motivated by applications to physics, in particular, quantum mechanics.

2.1. The Point Spectrum, Continuous Spectrum, and Residual Spectrum

The name resolvent is appropriate since helps to solve the equation Tαx = y. Thus, provided that exists. More importantly, the investigation of properties of will be basic for an understanding of the operator T itself. Naturally, many properties of Tα and depend on α, and the spectral theory is concerned with those properties. For instance, we are interested in the set of all α′s in the complex plane such that exists. Boundedness of is another property that will be essential. We will also ask for what α′s the domain of is dense in X, to name just a few aspects. A regular value α of T is a complex number such that exists and is bounded and whose domain is dense in X. For our investigation of T, Tα, and , we need some basic concepts in the spectral theory which are given, as follows (see [1, pages 370-371]).

The resolvent set ρ(T, X) of T is the set of all regular values α of T. Furthermore, the spectrum σ(T, X) is partitioned into the following three disjoint sets.

The point (discrete) spectrum σp(T, X) is the set such that does not exist. An ασp(T, X) is called an eigenvalue of T.

The continuous spectrum σc(T, X) is the set such that exists and is unbounded, and the domain of is dense in X.

The residual spectrum σr(T, X) is the set such that exists (and may be bounded or not) but the domain of is not dense in X.

Therefore, these three subspectra form a disjoint subdivision
()
To avoid trivial misunderstandings, let us say that some of the sets defined above may be empty. This is an existence problem which we will have to discuss. Indeed, it is well known that σc(T, X) = σr(T, X) = and the spectrum σ(T, X) consists of only the set σp(T, X) in the finite-dimensional case.

2.2. The Approximate Point Spectrum, Defect Spectrum, and Compression Spectrum

In this subsection, following Appell et al. [2], we define three more subdivisions of the spectrum called approximate point spectrum, defect spectrum, and compression spectrum.

Given a bounded linear operator T in a Banach space X, we call a sequence (xk) in X a Weyl sequence for T if ∥xk∥ = 1 and ∥Txk∥ → 0, as k. Then, the approximate point spectrum σap(T, X) of T is defined by
()
Moreover, the subspectrum
()
is called the defect spectrum of T.
The two subspectra given by (4) and (5) form a (not necessarily disjoint) subdivision
()
of the spectrum. There is another subspectrum,
()
which is often called compression spectrum in the literature. The compression spectrum gives rise to another (not necessarily disjoint) decomposition
()
of the spectrum. Clearly, σp(T, X)⊆σap(T, X) and σco (T, X)⊆σδ(T, X). Moreover, comparing these subspectra with those in (3), we note that
()

Sometimes it is useful to relate the spectrum of a bounded linear operator to that of its adjoint. Building on classical existence and uniqueness, results for linear operator equations in Banach spaces and their adjoints are also useful.

Proposition 1 (see [2], Proposition 1.3, page 28.)Spectrum and subspectrum of an operator TB(X) and its adjoint T*B(X*) are related by the following relations:

  • (a)

    σ(T*, X*) = σ(T, X),

  • (b)

    σc(T*, X*)⊆σap(T, X),

  • (c)

    σap(T*, X*) = σδ(T, X),

  • (d)

    σδ(T*, X*) = σap(T, X),

  • (e)

    σp(T*, X*) = σco(T, X),

  • (f)

    σco (T*, X*)⊇σp(T, X),

  • (g)

    σ(T, X) = σap(T, X) ∪σp(T*, X*) = σp(T,X) ∪ σap(T*,X*).

The relations (c)–(f) show that the approximate point spectrum is in a certain sense dual to the defect spectrum, and the point spectrum is dual to the compression spectrum. The equality (g) implies, in particular, that σ(T, X) = σap(T, X) if X is a Hilbert space and T is normal. Roughly speaking, this shows that normal (in particular, self-adjoint) operators on the Hilbert spaces are most similar to matrices in finite dimensional spaces (see [2]).

2.3. Goldberg′s Classification of Spectrum

If X is a Banach space and TB(X), then there are three possibilities for R(T):
  • (A)

    R(T) = X,

  • (B)

    ,

  • (C)

    ,

and
  • (1)

    T−1 exists and is continuous,

  • (2)

    T−1 exists but is discontinuous,

  • (3)

    T−1 does not exist.

If these possibilities are combined in all possible ways, nine different states are created. These are labelled by: A1, A2, A3, B1, B2, B3, C1, C2, and C3. If an operator is in state C2, for example, then and T−1 exists but is discontinuous (see [3]). Figure 1 due to Wenger [4] may be useful for the readers.

Details are in the caption following the image
State diagram for B(X) and B(X*) for a non-reflective Banach space X.

If α is a complex number such that TαA1 or TαB1, then αρ(T, X). All scalar values of α not in ρ(T, X) comprise the spectrum of T. The further classification of σ(T, X) gives rise to the fine spectrum of T. That is, σ(T, X) can be divided into the subsets A2σ(T, X) = , A3σ(T, X), B2σ(T, X), B3σ(T, X), C1σ(T, X), C2σ(T, X), and C3σ(T, X). For example, if Tα is in a given state, C2 (say), then we write αC2σ(T, X).

By the definitions given above, we can illustrate subdivision (3) in Table 1.

Table 1. Subdivision of spectrum of a linear operator.
1 2 3
exists exists
and is bounded and is unbounded does not exist
A R(αIT) = X αρ(T, X) ασp(T, X)
ασap(T, X)
  
ασc(T, X) ασp(T, X)
B αρ(T, X) ασap(T, X) ασap(T, X)
ασδ(T, X) ασδ(T, X)
  
ασr(T, X) ασr(T, X) ασp(T, X)
C ασδ(T, X) ασap(T, X) ασap(T, X)
ασδ(T, X) ασδ(T, X)
ασco(T, X) ασco(T, X) ασco(T, X)

Observe that the case in the first row and the second column cannot occur in a Banach space X, by the closed graph theorem. If we are not in the third column, that is, if α is not an eigenvalue of T, we may always consider the resolvent operator (on a possibly “thin” domain of definition) as “algebraic inverse” of αIT.

By a sequence space, we understand a linear subspace of the space ω = of all complex sequences which contain ϕ, the set of all finitely nonzero sequences, where = {0,1, 2, …}. We write , c, c0, and bv for the spaces of all bounded, convergent, null, and bounded variation sequences which are the Banach spaces with the sup-norm ∥x = sup k | xk| and , respectively, while ϕ is not a Banach space with respect to any norm. Also by p, we denote the space of all p-absolutely summable sequences which is a Banach space with the norm , where 1 ≤ p < .

Let μ and ν be two sequence spaces, and let A = (ank) be an infinite matrix of complex numbers ank, where k, n. Then, we say that A defines a matrix transformation from μ into ν, and we denote it by writing A : μν if for every sequence x = (xk) ∈ μ, the sequence Ax = {(Ax) n}, the A-transform of x, is in ν, where
()
By (μ : ν), we denote the class of all matrices A such that A : μν. Thus, A ∈ (μ : ν) if and only if the series on the right side of (10) converges for each n and each xμ, and we have Ax = {(Ax) n} nν for all xμ.
Throughout this paper, let λ = (λk) be a strictly increasing sequence of positive reals tending to infinity; that is,
()
Following Mursaleen and Noman [20], we define the matrix Λ = (λnk) of weighted mean relative to the sequence λ by
()
for all k, n. It is easy to show that the matrix Λ is regular and is reduced, in the special case λk = k + 1 for all k, to the matrix C1 of Cesàro mean of order one. Introducing the concept of Λ-strong convergence, several results on Λ-strong convergence of numerical sequences and Fourier series were given by Móricz [21]. Since we have
()
in the special case qk = λkλk−1 for all k, the matrix Λ is also reduced to the Riesz means Rq = (rnk) with respect to the sequence q = (qk). We say that a sequence x = (xk) ∈ ω is λ-convergent if Λxc. In particular, we say that x is a λ-null sequence if Λxc0 and we say that x is λ-bounded if Λx.

Lemma 2 (see [22], Theorem 1.3.6, page 6.)The matrix A = (ank) gives rise to a bounded linear operator TB(c) from c to itself if and only if

  • (1)

    the rows of A are in 1 and their 1 norms are bounded;

  • (2)

    the columns of A are in c;

  • (3)

    the sequence of row sums of A is in c.

The operator norm of T is the supremum of the 1 norms of the rows.

Corollary 3. Λ : cc is a bounded linear operator with the norm ∥Λ∥(c:c) = 1.

Lemma 4 (see [22], Example 8.4.5.A, page 129.)The matrix A = (ank) gives rise to a bounded linear operator TB(c0) from c0 to itself if and only if

  • (1)

    the rows of A are in 1 and their 1 norms are bounded,

  • (2)

    the columns of A are in c0.

The operator norm of T is the supremum of the 1 norms of the rows.

Corollary 5. Λ : c0c0 is a bounded linear operator with the norm .

We give a short survey concerned with the spectrum of the linear operators defined by some triangle matrices over certain sequence spaces. Wenger [4] examined the fine spectrum of the integer power of the Cesàro operator in c and Rhoades [5] generalized this result to the weighted mean methods. The fine spectrum of the operator on the sequence space p was studied by González [23], where 1 < p < . The spectrum of the Cesàro operator on the sequence spaces c0 and bv were also investigated by Reade [6], Akhmedov and Başar [7], and Okutoyi [8], respectively. The fine spectrum of the Rhaly operators on the sequence spaces c0 and c were examined by Yıldırım [9]. Furthermore, Coşkun [10] has studied the spectrum and fine spectrum for p-Cesàro operator acting on the space c0. Besides, de Malafosse [11] and Altay and Başar [12], respectively, studied the spectrum and the fine spectrum of the difference operator on the sequence spaces sr and c0, c, where sr denotes the Banach space of all sequences x = (xk) normed by , (r > 0). Altay and Karakuş [24] determined the fine spectrum of the Zweier matrix which is a band matrix as an operator over the sequence spaces 1 and bv. In 2010, Srivastava and Kumar [16] determined the spectra and the fine spectra of the double sequential band matrix Δν on 1, where Δν is defined by (Δν) nn = νn and (Δν) n+1,n = −νn for all n, under certain conditions on the sequence ν = (νk) and they have just generalized these results by the double sequential band matrix Δuv defined by Δuvx = (unxn + vn−1xn−1) n for all n (see [18]). Altun [25] studied the fine spectra of the Toeplitz operators, which are represented by upper and lower triangular n-band infinite matrices, over the sequence spaces c0 and c. Later, Karakaya and Altun determined the fine spectra of upper triangular double-band matrices over the sequence spaces c0 and c, in [26]. Quite recently, Akhmedov and El-Shabrawy [15] obtained the fine spectrum of the double sequential band matrix Δa,b, defined as a double-band matrix with the convergent sequences and having certain properties, over the sequence space c. The fine spectrum with respect to Goldberg’s classification of the operator B(r, s, t) defined by a triple band matrix over the sequence spaces p and bvp with 1 < p < has recently been studied by Furkan et al. [14]. Quite recently, Karaisa and Başar [19] have determined the fine spectrum of the upper triangular triple band matrix B(r, s, t) over the sequence space p, where 0 < p < . At this stage, Table 2 may be useful.

Table 2. Spectrum and fine spectrum of some triangle matrices in certain sequence spaces.
  σ(A, λ) σp(A, λ) σc(A, λ) σr(A, λ) Refer to
[4]
σ(W, c) [5]
σ(C1, c0) [6]
σ(C1, c0) σp(C1, c0) σc(C1, c0) σr(C1, c0) [7]
σ(C1, bv) [8]
σ(R, c0) σp(R, c0) σc(R, c0) σr(R, c0) [9]
σ(R, c) σp(R, c) σc(R, c) σr(R, c) [9]
[10]
σ(Δ, sr) [11]
σ(Δ, c0) [11]
σ(Δ, c) [11]
σ(1), c) σp(1), c) σc(1), c) σr(1), c) [12]
σ(1), c0) σp(1), c0) σc(1), c0) σr(1), c0) [12]
σ(B(r, s), p) σp(B(r, s), p) σc(B(r, s), p) σr(B(r, s), p) [13]
σ(B(r, s), bvp) σp(B(r, s), bvp) σc(B(r, s), bvp) σr(B(r, s), bvp) [13]
σ(B(r, s, t), p) σp(B(r, s, t), p) σc(B(r, s, t), p) σr(B(r, s, t), p) [14]
σ(B(r, s, t), bvp) σp(B(r, s, t), bvp) σc(B(r, s, t), bvp) σr(B(r, s, t), bvp) [14]
σa,b, c) σpa,b, c) σca,b, c) σra,b, c) [15]
σν, 1) σpν, 1) σcν, 1) σrν, 1) [16]
[17]
σuv, 1) σpuv, 1) σcuv, 1) σruv, 1) [18]
σ(B(r, s, t), p) σp(B(r, s, t), p) σc(B(r, s, t), p) σr(B(r, s, t), p) [19]

In this work, our purpose is to determine the fine spectrum of the operator Λ over the sequence spaces c0 and c with respect to Goldberg’s classification. Additionally, we give the approximate point spectrum, defect spectrum, and compression spectrum of the matrix operator Λ over the spaces c0 and c. Finally, we state and prove a Mercerian theorem.

3. The Fine Spectrum of the Operator Λ on the Sequence Space c0

In this section, we examine the spectrum, the point spectrum, the continuous spectrum, the residual spectrum, the fine spectrum, the approximate point spectrum, the defect spectrum, and the compression spectrum of the operator Λ on the sequence space c0. For simplicity in the notation, we write throughout that cn = (λnλn−1)/λn for all n and we use this abbreviation with other letters.

Theorem 6. σ(Λ, c0)⊆{α : |2α − 1| ≤ 1}.

Proof. Let |2α − 1| > 1. Since Λ − αI is triangle, (Λ−αI)−1 exists, and solving the matrix equation (Λ − αI)x = y for x in terms of y gives the matrix (Λ−αI)−1 = B = (bnk), where

()
for all k, n. Thus, we observe that
()
The inequality |2α − 1| > 1 is equivalent to γ > −1, where −(1/α) = γ + iβ. For all α,
()
holds for all j. So, 1/|1 − (cj/α)| ≤ 1/(1 + γcj).

Firstly we take −1 < γ < 0. Since 0 < cj ≤ 1, we have 1 + γ ≤ 1 + γcj < 1. Therefore 1/(1 + γcj) < 1/(1 + γ) and 1 < 1/(1 + γ) < for 0 < 1 + γ < 1.

()

Secondly we get 0 ≤ γ. Since 1 < 1 + γcj ≤ 1 + γ, 1/(1 + γcj) < 1. So,

()
Therefore, we have
()
that is, (Λ−αI)−1 ∈ (c0 : c0). But for |2α − 1| ≤ 1,
()
that is, (Λ−αI)−1 is not in B(c0). This completes the proof.

Theorem 7. Define μ and η by μ = limsup jcj and η = liminf jcj. Then,

()

Proof. Let |α − 1/(2 − μ)| < (1 − μ)/(2 − μ) and αcj for any j. Then,

()
So we have,
()
Note that |1 + (1 − α−1)(λjλj−1)/λj−1| ≤ 1 if and only if
()
where −α−1 = γ + iβ. So, one can see that
()
which is equivalent to the inequality
()
For inequality (26) to be true for all sufficiently large j, it is sufficient to have
()
We can write (λjλj−1)/λj−1 = λj(λjλj−1)/(λjλj−1) and λj/λj−1 = 1/(1 − cj). Therefore,
()
()
since the function g defined by g(x) = x/(1 − x) is monotone increasing in x for 0 < x < 1.

For (27) to be true for all sufficiently large j, it is sufficient to have μ satisfying

()
which is equivalent to
()
Therefore, for all nN, for some fixed N,
()
which diverges in the light of (29).

If α = cj for any j, then clearly α lies in the spectrum of Λ. This completes the proof.

Theorem 8. σ(Λ, c0)⊆{α : |α − 1/(2 − η)| ≤ (1 − η)/(2 − η)} ∪ S.

Proof. Let α be fixed and satisfy the inequality

()
and αcj for any j. We will show that αρ(Λ, c0). From Theorem 6, we need consider only those values of α satisfying |2α − 1| > 1; that is, γ > −1. Under the assumption on α, we wish to verify that
()
for all sufficiently large j. It will be sufficient to show that
()
that is,
()
which is equivalent to (33).

Define the function f by f(t) = 1 + 2(1 + γ)t + [(1 + γ) 2 + β2]t2. f has a minumum at t0 = −(1 + γ)/[(1+γ)2 + β2]. The above inequality is equivalent to η(γ2 + β2) + 2γ > η − 2 and is also equivalent to

()
Therefore, for those values of η satisfying (37), f is monotone increasing. Let ϵ > 0 and small. Then f(η/(1 − η) − ϵ) = f(η/(1 − η)) − (2ϵ)g(ϵ), where g(ϵ) = 1 + γ + [(1 + γ) 2 + β2][η/(1 − η) − ϵ/2]. Note that g(ϵ) > 0 for small ϵ, since f is monotone increasing for t > η/[2(1 − η)], we will now show that f(η/(1 − η)) > 1. From (37),
()
which is equivalent to
()
But 1/(1 − η) = 1 + η/(1 − η), so we have f(η/(1 − η)) = . Now choose ϵ > 0 and so small that f(η/(1 − η) − ϵ) = f(η/(1 − η)) − 2ϵg(ϵ) = m2 > 1. Then, by the definition of η, there exists an N such that n > N implies (λn+1λn)/λn > η/(1 − η) − ϵ, so that f((λn+1λn)/λn) > f(η/(1 − η) − ϵ) = m2. Using (23),
()
for all nN. Therefore {|bnk|} is monotone decreasing in n for each k, n > N, so that B has bounded columns. It remains to show that B has finite norm.

For ϵ being used, from (29), we can enlarge N, if necessary, to ensure that (λnλn−1)/λn−1 < μ/(1 − μ) + 1 for nN. From (23),

()
where H is a constant independent of n. Further
()
Hence, B has a finite norm.

Corollary 9. Let δ = lim jcj exist. Then,

()

If TB(c0) with the matrix A, then it is known that the adjoint operator is dened by the transpose At of the matrix A. It should be noted that the dual space of c0 is isometrically isomorphic to the Banach space 1 of absolutely summable sequences normed by .

Theorem 10. Let δ be defined as in Corollary 9. Then, .

Proof. Suppose that Λ*x = αx for xθ in . Then, by solving the system of linear equations

()
we can write xn = (λnλn−1)/λn−1(1 − α−1)[1+(1 − α−1)(λjλj−1)/λj−1]. Let |α − 1/(2 − δ)| < (1 − δ)/(2 − δ) or αS and [1 + (1 − α−1)(λjλj−1)/λj−1]. One can see that |1 + (1 − α−1)(λjλj−1)/λj−1| < 1 for all sufficiently large j if and only if
()
Then, we have, from the discussion in Theorem 7 and the hypothesis on α,
()
for all sufficiently large n, so is convergent. Since |(1 − α−1)(λnλn−1)/λn−1x0| is bounded, it follows that is convergent, so that Λ*x = αx has nonzero solutions. Therefore, the proof is completed.

Theorem 11. Let δ be defined as in Corollary 9. Then

()

Proof. Let ck be any diagonal entry satisfying 0 < ckδ/(2 − δ). Let j be the smallest integer such that cj = ck. By setting xn = 0 for n > j + 1, x0 = 0, the system (Λ*cjI)x = θ reduces to a homogeneous linear system of j equations in j + 1 unknowns, so that nontrivial solutions exist. Therefore Λ − cjI ∈ 3.

Λ − αI is not one to one for α = 0,1 and so Λ − αI ∈ 3. This step concludes the proof.

Lemma 12 (see [3], page 59.)T has a dense range if and only if T* is one to one.

Theorem 13. .

Proof. For , the operator Λ − αI is triangle, so has an inverse. But Λ*αI is not one to one by Theorem 10. Therefore by Lemma 12, , and this step concludes the proof.

Theorem 14. Let δ be defined as in Corollary 9 and cnδ for all sufficiently large n. Then,

()

Proof. Fix α ≠ 1, δ/(2 − δ), and satisfying |α − 1/(2 − δ)| = (1 − δ)/(2 − δ). Since the operator Λ − αI is triangle, it has an inverse. Consider the adjoint operator Λ*αI. As in Theorem 11, x0 is arbitrary and

()
for all positive n. From the hypothesis, there exists a positive integer N such that nN implies cnδ. This fact, together with the condition on α, implies that |1 + (1 − α−1)(λnλn−1)/λn−1| ≥ 1 for nN. Thus, |xn | = C(λnλn−1)/λn−1 for nN, where C is a positive constant independent of n. We can write
()
Therefore (xn) ∈ 1x0 = 0; that is, Λ*αI is one to one. From Lemma 12, the range of Λ − αI is dense in c0. This completes the proof.

Lemma 15 (see [3], page 60.)T has a bounded inverse if and only if T* is onto.

Theorem 16. Let δ be defined as in Corollary 9 and less than 1. If α satisfies |α − 1/(2 − δ)| < (1 − δ)/(2 − δ) and αS, then αC1σ(Λ, c0).

Proof . First of all Λ − αI is a triangle; hence 1 − 1. Therefore Λ − αI ∈ 1 ∪ 2. To verify that Λ − αIC1σ(Λ, c0) it is sufficient to show that Λ*αI is onto by Lemma 15.

Suppose y = (Λ*αI)x, where x, y1. Then, x0 = 1/(1 − α)y0λ0/[(λ1λ0)(1 − α)]y1 and

()
Choose x1 = 0 and solve (51) for x in terms of y to get
()
()
For example, substituting (52) into (53), with n = 2, yields
()
so that x2 = (λ2λ1)/[α(λ1λ0)]y1 − (1/α)y2. For n = 3,
()

Continuing this process, the entries of the matrix B = (bnk) such that By = x are calculated as

()
and bnk = 0 otherwise.

To show that BB(1), it is sufficient to establish that is finite independent of k. . We may write 1 − (cj/α) = (λj−1/λj)[1 + (1 − α−1)(λjλj−1)/λj−1]. Also, sup n|(λnλn−1)/λn−1| ≤ M < . Therefore,

()
and, for k > 1,
()

Since k > 1, the series in inequality (24) is absolutely convergent from Theorem 7. Therefore, is finite.

Because of (Λ−αI)−1 is bounded, it is continuous, and αC1σ(Λ, c0). This completes the proof.

Theorem 17. Let δ be defined as in Corollary 9 and δ < 1. If α = δ or α = cn for all n and δ/(2 − δ) < α < 1, then αC1σ(Λ, c0).

Proof. First assume that Λ has distinct diagonal entries and fix j ≥ 1. Then the system (Λ − cjI)x = θ implies that xn = 0 for n = 0,1, …, j − 1, and for nj

()
The system (59) yields the following recursion relation:
()
which can be solved for xn to yield
()

Since 0 < cj < 1, the argument of Theorem 7 applies and (24) is true. Therefore xc0 implies x = θ and Λ − cjI is 1 − 1, so that Λ − cjI ∈ 1 ∪ 2.

Clearly Λ − cjIC. It remains to show that Λ*cjI is onto.

Suppose that (Λ*cjI)x = y, where x, y1. By choosing xj+1 = 0 we can solve for x0, x1, …, xj in terms of y0, y1, …, yj+1. As in Theorem 16, the remaining equations can be written in the form x = By, where the nonzero entries of B = (bnk) are as follows

()
From (62), we have
()
For m > 1,
()
Using 1 − (cj/α) = (λj−1/λj)[1 + (1 − α−1)(λjλj−1)/λj−1], one can convert (63) and (64) the similar expressions to (57) and (58), and therefore is finite.

Suppose that Λ does not have distinct diagonal entries. The restriction on α guarantees that no zero diagonal entries are being considered. Let cj ≠ 0 be any diagonal entry which occurs more than once, and let k, r denote, respectively, the smallest and largest integers for which cj = ck = cr. From (61) it follows that xn = 0 for nr. Also, xn = 0 for 0 ≤ n < k. Therefore the system (Λ − cjI)x = θ becomes

()
Case 1. Let r = k + 1. Then (65) reduces to the single equation
()
which implies that xk = 0, since cj = cr = ck+1 and cj ≠ 0. Therefore x = θ.

Case 2. Let r > k + 1. From (65) one can obtain the recursion formula xn = λn+1(cjcn+1)xn+1/(λncj) with k < n < r. Since xr = 0 it then follows that xn = 0 for k < n < r. Using (65) with n = k + 1 yields xk = 0 and so again x = θ.

To show that Λ*cjI is onto, suppose (Λ*cjI)x = y, where x, y1. By choosing xj+1 = 0 one can solve for x0, x1, …, xj in terms of y0, y1, …, yj+1. As in Theorem 16, the remaining equations can be written in the form x = By, where the nonzero entries of B are as in (62) with the other entries of B clearly zero.

Since kjr, there are two cases to consider.

Case 1. If j = r, then the proof proceeds exactly as in the argument following (62).

Case 2. If j < r, then from (62), bj+m,j+k = bj+m,j+1 = 0 at least for mrj + 2. If there are other values of n with j < n < r for which cncj, then additional entries of B will be zero. These zero entries do not affect the validity of the argument showing that (63) converges.

If δ = 0, then 0 does not lie inside the disc, and so it is not considered in this theorem.

Let α = δ > 0. If λnnδ for each n ≥ 1, all i sufficiently large, then the argument of Theorem 16 applies and Λ − δIC1. If λnn = δ for some n, then the proof of Theorem 17 applies with cj replaced by δ and again, Λ − δIC1.

Therefore, in all cases, Λ − cjI ∈ 1 ∪ 2.

Theorem 18. If ασp(Λ, c0), αC3σ(Λ, c0).

Proof . For ασp(Λ, c0), Λ − αI ∈ 3 and Λ*αI is not one to one. Therefore by Lemma 12. This concludes the proof.

Theorem 19. The statement A3σ(Λ, c0) = C2σ(Λ, c0) = holds.

Proof. Let δ be defined as in Corollary 9 and cnδ for all sufficiently large n, then A3σ(Λ, c0) = and C2σ(Λ, c0) = follow from Corollary 9, Theorems 14, and 1618.

Define the set E by
()
where η is as in Theorem 7.

We will consider δ = η, that is, for which the main diagonal entries converge, where δ as in Corollary 9.

Theorem 20. The following results hold:

  • (a)

    σap(Λ, c0) = {α : |α − (2 − δ) −1| = (1 − δ)/(2 − δ)} ∪ E,

  • (b)

    σδ(Λ, c0) = σ(Λ, c0),

  • (c)

    σco(Λ, c0) = {α : |α − (2 − δ) −1| < (1 − δ)/(2 − δ)} ∪ S.

Proof . (a) Since the relation

()
holds by Theorems 16 and 17 and from Table 1, σap(Λ, c0) = σ(Λ, c0)∖C1σ(Λ, c0). Therefore, we have σap(Λ, c0) = {α : |α − (2−δ)−1| = (1 − δ)/(2 − δ)} ∪ E.

(b) Since σδ(Λ, c0) = σ(Λ, c0)∖A3σ(Λ, c0) from Table 1 and A3σ(Λ, c0) = by Theorem 19, we have σδ(Λ, c0) = σ(Λ, c0).

(c) Since the equality σco (Λ, c0) = C1σ(Λ, c0) ∪ C2σ(Λ, c0) ∪ C3σ(Λ, c0) holds from Table 1, we have σco (Λ, c0) = {α : |α − (2 − δ) −1| < (1 − δ)/(2 − δ)} ∪ S by Theorems 1619.

The next corollary can be obtained from Proposition 1.

Corollary 21. The following results hold:

  • (a)

    σap*, 1) = σ(Λ, c0),

  • (b)

    σδ*, 1) = {α : |α − (2−δ)−1| = (1 − δ)(2 − δ)} ∪ E,

  • (c)

    σp*, 1) = {α : |α − (2−δ)−1| < (1 − δ)/(2 − δ)} ∪ S.

4. The Fine Spectrum of the Operator Λ on the Sequence Space c

In this section, we investigate the fine spectrum of the operator Λ over the sequence space c.

Theorem 22. σ(Λ, c)⊆{α : |2α − 1| ≤ 1}.

Proof. This is obtained in a similar way to that used in the proof of Theorem 6.

Theorem 23. Suppose that μ, η and S be defined as in Theorem 7. Then,

()

Proof. This is similar to the proof of Theorems 7 and 8. To avoid the repetition of the similar statements, we omit the detail.

Corollary 24. Let δ be defined as in Corollary 9. Then,

()

If T : cc is a bounded linear operator with the matrix A, then T* : c*c* acting on 1 has a matrix representation of the form
()
where χ is the limit of the sequence of row sums of A minus the sum of the limit of the columns of A and b is the column vector whose kth entry is the limit of the kth column of A for each k. For Λ : cc, the matrix Λ*B(1) is of the form
()

Theorem 25. Let δ be defined as in Corollary 9. Then,

()

Proof. Suppose that Λ*x = αx for xθ in c*1. Then, by solving the system of linear equations

()
we get by assumption (1 − α)x0 = 0 with α = 1 that x = (x0, x1, 0,0, …) ∈ c. If α ≠ 1, we have x0 = 0 and (xn) ∈ 1 if and only if |1 + (1 − α−1)(λjλj−1)/λj−1| < 1, by Theorem 11. This completes the proof.

Theorem 26. Let δ be defined as in Corollary 9. Then,

()

Proof. The proof is identical to the proof of Theorem 11.

Theorem 27. σr(Λ, c) = σp*, c*)∖σp(Λ, c).

Proof. For ασp*, c*)∖σp(Λ, c), the operator Λ − αI is triangle, so has an inverse. But Λ*αI is not one to one by Theorem 26. Therefore by Lemma 12, and this concludes the proof.

Since Theorems 2831 can be proved in a similar way to that used in the proof of Theorems 14 and 1618; respectively, to avoid the repetition of the similar statements we omit the detailed proof and give them without proof.

Theorem 28. Let δ be defined as in Corollary 9 and cnδ for all sufficiently large n. Then,

()

Theorem 29. Let δ be defined as in Corollary 9 and less than 1. If α satisfies |α − 1/(2 − δ)| < (1 − δ)/(2 − δ) and αS, then αC1σ(Λ, c).

Theorem 30. Let δ be defined as in Corollary 9 and δ < 1. If α = δ or α = cn for all n and δ/(2 − δ) < α < 1, then αC1σ(Λ, c).

Theorem 31. If α ∈  σp(Λ, c), αC3σ(Λ, c).

Theorem 32. The following statement holds:A3σ(Λ, c) = C2σ(Λ, c) = .

Proof. Let δ be defined as in Corollary 9 and cnδ for all sufficiently large n, then A3σ(Λ, c) = and C2σ(Λ, c) = follow from Corollary 24 and Theorems 2831.

Theorem 33. The following results hold:

  • (a)

    σap(Λ, c) = {α : |α − (2 − δ) −1| = (1 − δ)/(2 − δ)} ∪ E,

  • (b)

    σδ(Λ, c) = σ(Λ, c),

  • (c)

    σco(Λ, c) = {α : |α − (2 − δ) −1| < (1 − δ)/(2 − δ)} ∪ S.

Proof. (a) Since the relation C1σ(Λ, c) = {{α : |α − (2 − δ) −1| < (1 − δ)/(2 − δ)}∖S}⋃ {α = λnn : δ/(2 − δ) < α < 1} holds by Theorems 29 and 30 and from Table 1, σap(Λ, c) = σ(Λ, c)∖C1σ(Λ, c). Therefore, we have σap(Λ, c) = {α : |α − (2 − δ) −1| = (1 − δ)/(2 − δ)} ∪ E.

(b) Since σδ(Λ, c) = σ(Λ, c)∖A3σ(Λ, c) from Table 1 and A3σ(Λ, c) = by Theorem 32, we have σδ(Λ, c) = σ(Λ, c).

(c) Since the equality σco (Λ, c) = C1σ(Λ, c) ∪ C2σ(Λ, c) ∪ C3σ(Λ, c) holds from Table 1, we have σco (Λ, c) = {α : |α − (2 − δ) −1| < (1 − δ)/(2 − δ)} ∪ S by Theorems 2932.

The next corollary can be obtained from Proposition 1.

Corollary 34. The following results hold:

  • (a)

    σap*, 1) = σ(Λ, c),

  • (b)

    σδ*, 1) = {α : |α − (2 − δ) −1| = (1 − δ)/(2 − δ)} ∪ E,

  • (c)

    σp*, 1) = {α : |α − (2 − δ) −1| < (1 − δ)/(2 − δ)} ∪ S.

Let A be an infinite matrix and let the set cA denote the convergence domain of that matrix A, a theorem which proves that cA = c is called a Mercerian theorem, after Mercer, who proved a significant theorem of this type [28, page 186].

Now, we may give our final theorem.

Theorem 35. Suppose that |α + 1 | >|α − 1|. Then the convergence field of A = αI + (1 − α)Λ is c.

Proof. By Theorem 22, Λ − [α/(α − 1)]I has an inverse in B(c). That is to say that

()
Since A is a triangle and is in B(c), A−1 is also conservative which implies that cA = c [22, page 12].

5. Conclusion

Although the matrix Λ is used for obtaining some new sequence spaces by its domain from the classical sequence spaces, it is not considered for determining the spectrum or fine spectrum acting as a linear operator on any of the classical sequence spaces c0, c, or p. Following Altay and Başar [12] and Karakaya and Altun [26], we determine the fine spectrum with respect to Goldberg’s classification of the operator defined by the triangle matrix Λ over the sequence spaces c0 and c which reduces to a new regular triangle matrix depending on choosing the strictly increasing sequence λ = (λk) of positive real numbers tending to infinity. Additionally, we give the approximate point spectrum, the defect spectrum, and the compression spectrum of the matrix operator Λ over the spaces c0 and c. Since the present paper is devoted to the fine spectrum of the operator defined by the lambda matrix over the sequence spaces c0 and c with new subdivision of spectrum, this makes it significant. We should note that the main results of the present paper are given as an extended abstract without proof by Yeşilkayagil and Başar [29].

The generalized weighted means G(u, v) = (gnk) is defined by
()
for all k, n, where un depends only on n and vk only on k such that un, vk ≠ 0. It is immediate that in the case un = 1/λn and vk = λkλk−1, the generalized weighted means G(u, v) corresponds to the matrix Λ. Although the Riesz means Rq, the generalized weighted means G(u, v), and the matrix Λ were used for different purposes, their fine spectrum over the classical sequence spaces was not studied. As a beginning, the present work has an advantage.

Finally, we record from now on that our next paper will be devoted to the investigation of the fine spectrum of the matrix operator Λ on the spaces p and bvp in the cases 0 < p < 1 and 1 ≤ p < , where bvp denotes the space of all sequences whose Δ-transforms are in the space p and was studied in the case 0 < p < 1 by Altay and Başar [30] and in the case 1 ≤ p by Başar and Altay [31].

Acknowledgments

The authors would like to express their pleasure to Professor Bilâl Altay, Department of Mathematical Education, Faculty of Education, İnönü University, Malatya, Turkey, for many helpful suggestions and interesting comments on the main results of the earlier version of the paper. Additionally, the authors are very grateful to the referee for making some useful remarks which improved the presentation of the paper. The main results of this paper has been presented in part at the conference First International Conference on Analysis and Applied Mathematics (ICAAM 2012) to be held on October 18–21, 2012, in Gümüşhane, Turkey, at the University of Gümüşhane and published in the conference proceedings with AIP, as an extended abstract.

      The full text of this article hosted at iucr.org is unavailable due to technical difficulties.