Volume 2013, Issue 1 904976
Research Article
Open Access

J-Self-Adjoint Extensions for a Class of Discrete Linear Hamiltonian Systems

Guojing Ren

Corresponding Author

Guojing Ren

School of Mathematics and Quantitative Economics, Shandong University of Finance and Economics, Jinan, Shandong 250014, China sdfi.edu.cn

Search for more papers by this author
Huaqing Sun

Huaqing Sun

Department of Mathematics, Shandong University at Weihai, Weihai, Shandong 264209, China sdu.edu.cn

Search for more papers by this author
First published: 28 May 2013
Citations: 1
Academic Editor: Michiel Bertsch

Abstract

This paper is concerned with formally J-self-adjoint discrete linear Hamiltonian systems on finite or infinite intervals. The minimal and maximal subspaces are characterized, and the defect indices of the minimal subspaces are discussed. All the J-self-adjoint subspace extensions of the minimal subspace are completely characterized in terms of the square summable solutions and boundary conditions. As a consequence, characterizations of all the J-self-adjoint subspace extensions are given in the limit point and limit circle cases.

1. Introduction

Consider the following discrete linear Hamiltonian system:
where , a is a finite integer or a = −, b is a finite integer or b = +, and ba ≥ 1; Δ is the forward difference operator, that is, Δy(t) = y(t + 1) − y(t); 𝒥 is the canonical symplectic matrix, that is,
()
where In is the n × n unit matrix; the weighted function W(t) is a 2n × 2n real symmetric matrix with W(t) ≥ 0 for t, and it is of the block diagonal form,
()
where P(t) is a 2n × 2n complex symmetric matrix, that is, PT(t) = P(t). The partial right shift operator R(y)(t) = (uT(t + 1), vT(t)) T with y(t) = (uT(t), vT(t)) T and u(t), v(t) ∈ n; λ is a complex spectral parameter.

For briefness, denote = [a, b] in the case where a and b are finite integers; = [a, +) in the case where a is finite and b = +; = (−, b] in the case where a = − and b is finite; = (−, +) in the case where a = − and b = +.

Since P(t) is symmetric, it can be blocked as
()
where A, B, and C are n × n complex-valued matrices with CT = C and BT = B. Then, (1λ) can be rewritten as
To ensure the existence and uniqueness of the solution of any initial value problem for (1λ), we always assume in the present paper that
  • (A1) InA(t) is invertible in .

It can be easily verified that (1λ) contains the following complex coefficients vector difference equation of order 2m:
where pj(t) are l × l complex-valued matrices with , 0 ≤ jm; pm(t) is invertible in ; w(t) is an l × l real-valued with w(t) ≥ 0. In fact, by letting y = (uT, vT) T with , , and
()
for 1 ≤ jm, (3λ) can be converted into (1λ), as well as (2λ), with
()
It is obvious that (A1) is satisfied for (3λ).

The spectral theory of self-adjoint operators and self-adjoint extensions of symmetric operators (i.e., densely defined Hermitian operators) in Hilbert spaces has been well developed (cf. [14]). In general, under certain definiteness conditions, a formally self-adjoint differential expression can generate a minimal operator which is symmetric, and the defect index of the minimal operator is equal to the number of linearly independent square integrable solutions. All the characterizations of self-adjoint extensions of differential equation are obtained [58].

However, for difference equations, it was found in [9] that the minimal operator defined in [10] may be neither densely defined nor single-valued even if the definiteness condition is satisfied. This is an important difference between the differential and difference equations. In order to study the self-adjoint extensions of nondensely defined or multivalued Hermitian operators, some scholars tried to extend the concepts and theory for densely defined Hermitian operators to Hermitian subspaces [1115]. Recently, Shi extended the Glazman-Krein-Naimark (GKN) theory for symmetric operators to Hermitian subspaces [9]. Applying this GKN theory, the first author, with Shi and Sun, gave complete characterizations of self-adjoint extensions for second-order formally self-adjoint difference equations and general linear discrete Hamiltonian systems, separately [16, 17].

We note that when the coefficient P(t) in (1λ) is not a Hermitian matrix, that is, P*(t) ≠ P(t), system (1λ) is not formally self-adjoint, and the minimal subspace generated by (1λ) is not Hermitian. Hence the spectral theory of self-adjoint operators or self-adjoint subspaces is not applicable. To solve this problem, Glazman introduced a concept of J-symmetric operators in [3, 18] where J is an operator. The minimal operators generated by certain differential expressions are J-symmetric operators in the related Hilbert spaces [19, 20]. Monaquel and Schmidt [21] discussed the M-functions of the following discrete Hamiltonian system:
where ∇ is the backward difference operator, that is, ∇z(t) = z(t) − z(t − 1), and weighted function H(t) = diag {H1(t), H2(t)}. By letting z1(t) = v(t), z2(t) = u(t + 1), (1λ) can be converted into (4λ) with
()

In [22], the result that every J-Hermitian subspace has a J-self-adjoint subspace extension has been given. Furthermore, a result about J-self-adjoint subspace extension was obtained [22], which can be regarded as a GKN theorem for J-Hermitian subspaces.

In the present paper, enlightened by the methods used in the study of self-adjoint subspace extensions of Hermitian subspaces, we will study the J-self-adjoint subspace extensions of the minimal operator corresponding to system (1λ). A complete characterization of them in terms of boundary conditions is given by employing the GKN theorem for J-Hermitian subspaces. The rest of this paper is organized as follows. In Section 2, some basic concepts and useful results about subspaces are briefly recalled. In Section 3, a conjugation operator J is defined in the corresponding Hilbert space, and the maximal and minimal subspaces are discussed. In Section 4, the description of the minimal subspaces is given by the properties of their elements at the endpoints of the discussed intervals, the defect indices of minimal subspaces are discussed, and characterizations of the maximal subspaces are established. Section 5 pays attention to two characterizations of all the self-adjoint subspace extensions of the minimal subspace in terms of boundary conditions via linearly independent square summable solutions of (1λ). As a consequence, characterizations of all the self-adjoint subspace extensions are given in two special cases: the limit point and limit circle cases.

2. Fundamental Results on Subspaces

In this section, we recall some basic concepts and useful results about subspaces. For more results about nondensely defined J-Hermitian operators or J-Hermitian subspaces, we refer to [1719, 22] and some references cited therein. In addition, some properties of solutions of (1λ) and a result about matrices are given at the end of this section.

By and we denote the sets of the real and the complex numbers, respectively. Let X be a complex Hilbert space equipped with inner product 〈·, ·〉, T and S two linear subspaces (briefly, subspace) in X2 : = X × X, and λ. Denote
()
If TS = {0}, we write
()
which is denoted by TS in the case that T and S are orthogonal.
Denote
()
It can be easily verified that T(0) = {0} if and only if T can determine a unique linear operator from Dom  T into X whose graph is just T. For convenience, we will identify a linear operator in X with a subspace in X2 via its graph.

Definition 1 (see [11].)Let T be a subspace in X2.

  • (1)

    T is said to be a Hermitian subspace if TT*. Furthermore, T is said to be a Hermitian operator if it is an operator, that is, T(0) = {0}.

  • (2)

    T is said to be a self-adjoint subspace if T = T*. Furthermore, T is said to be a self-adjoint operator if it is an operator, that is, T(0) = {0}.

  • (3)

    Let T be a Hermitian subspace. T1 is said to be a self-adjoint subspace extension (briefly, SSE) of T if TT1 and T1 is a self-adjoint subspace.

  • (4)

    Let T be a Hermitian operator. T1 is said to be a self-adjoint operator extension (briefly, SOE) of T if TT1 and T1 is a self-adjoint operator.

Lemma 2 (see [11].)Let T be a subspace in X2. Then

  • (1)

    T* is a closed subspace in X2;

  • (2)

    and , where is the closure of T;

  • (3)

    .

In [19], an operator J defined in X is said to be a conjugation operator if for all x, yX,
()

Definition 3. Let T be a subspace in X2 and J be a conjugation operator.

  • (1)

    The J-adjoint of T is defined by

    ()

  • (2)

    T is said to be a J-Hermitian subspace if . Furthermore, T is said to be a J-Hermitian operator if it is an operator, that is, T(0) = {0}.

  • (3)

    T is said to be a J-self-adjoint subspace if . Furthermore, T is said to be a J-self-adjoint operator if it is an operator, that is, T(0) = {0}.

  • (4)

    Let T be a J-Hermitian subspace. T1 is said to be a J-self-adjoint subspace extension (briefly, J-SSE) of T if TT1 and T1 is a J-self-adjoint subspace.

  • (5)

    Let T be a J-Hermitian operator. T1 is said to be a J-self-adjoint operator extension (briefly, J-SOE) of T if TT1 and T1 is a J-self-adjoint operator.

Remark 4. (1) It can be easily verified that is a closed subspace. Consequently, a J-self-adjoint subspace T is a closed subspace since . In addition, if TS.

(2) From the definition, we have that 〈f, Jy〉 = 〈x, Jg〉 holds for all (x, f) ∈ T and , and that T is a J-Hermitian subspace if and only if 〈f, Jy〉 = 〈x, Jg〉 for all (x, f), (y, g) ∈ T.

Lemma 5 (see [22].)Let T be a subspace in X2. Then

  • (1)

    ;

  • (2)

    .

It follows from Lemmas 2 and 5 that , and is J-Hermitian if T is J-Hermitian.

Lemma 6 (see [22].)Every J-Hermitian subspace has a J-SSE.

Definition 7. Let T be a J-Hermitian subspace. Then is called to be the defect index of T.

Next, we introduce a form on X2 × X2 by
()

Lemma 8 (see [22].)Let T be a J-Hermitian subspace. Then

()

Lemma 9 (see [22].)Let T be a closed J-Hermitian subspace in X2 and satisfy d = d(T)<+. Then a subspace T1 is a J-SSE of T if and only if and there exists such that

  • (1)

    (x1, f1), (x2, f2), …, (xd, fd) are linearly independent in (modulo T);

  • (2)

    [(xj, fj):(xk, fk)] = 0, 1 ≤ j, kd;

  • (3)

    .

Lemma 9 can be regarded as a GKN theorem for J-Hermitian subspaces. A set of which is satisfying (1) and (2) in Lemma 9 is called a GKN set of T.

Definition 10. Let T be a subspace in X2.

  • (1)

    The set

    ()
    is called the resolvent set of T.

  • (2)

    The set σ(T): = ρ(T) is called the spectrum of T.

  • (3)

    The set

    ()
    is called to be the regularity field of T.

It is evident that ρ(T) ⊂ Γ(T) for any subspace T in X2.

Lemma 11 (see [22].)Let T be a J-Hermitian subspace in X2 with Γ(T) ≠ , and λ ∈ Γ(T). Then

()

The following is a well-known result on the rank of matrices.

Lemma 12. Let A be an m × l matrix and B an l × n matrix. Then

()
In particular, if AB = 0, then
()

3. Relationship between the Maximal and Minimal Subspaces

This section is divided into three subsections. In the first subsection, we define a conjugation operator in a Hilbert space. In the second subsection, we define maximal and minimal subspaces generated by (1λ) and discuss relationship between them. In the last subsection, we discuss the definiteness condition corresponding to (1λ).

3.1. Conjugation Operator

In this subsection, we define a conjugation operator in a Hilbert space and then discuss its properties.

Since b and a may be finite or infinite, we introduce the following conventions for briefness: b + 1 means + in the case of b = + and a − 1 means − in the case of a = −. Denote
()
For any 2n × 2n Hermitian matrix W(t) ≥ 0 defined in , we define
()
with the semiscalar product
()
Furthermore, denote ∥y∥ : = (〈y,y〉)1/2 for . Since the weighted function W(t) may be singular in , ∥·∥ is a seminorm. Introduce the quotient space
()
Then is a Hilbert space with the inner product 〈·, ·〉.

For a function , denote by yπ the corresponding class in . And for any , denote by a representative of yπ. It is evident that 〈yπ, zπ〉 = 〈y, z〉 for any .

For any , denote by J0y the conjugation of y; that is,
()
It can be easily verified that if and only if . Here is the conjugation of matrix W. Since each is an equivalent class, we define a operator J defined on by
()
The following result is obtained.

Lemma 13. J defined by (24) is a conjugation operator defined on if and only if W(t) is real and symmetric in .

Proof. The sufficiency is evident. Next, we consider the necessity. Assume that J defined by (24) is a conjugation operator in . Then for any , it follows from 〈Jxπ, Jyπ〉 = 〈yπ, xπ〉 that

()
By the arbitrariness of xπ, yπ, one has that W(t) = WT(t). This, together with W(t) = W*(t), yields that W(t) is real. The proof is complete.

For any x, yl(), we denote
()
where 𝒥 is the canonical symplectic matrix given in Section 1. In the case of b = +, if lim tb (x, y)(t) exists and is finite, then its limit is denoted by (x, y)(+). In the case of a = −, if lim ta (x, y)(t) exists and is finite, then its limit is denoted by (x, y)(−).
Denote
()
where τ and δ are called the natural difference operators corresponding to system (1λ). The following result can be easily verified, and so we omit the proof.

Lemma 14. Assume that (A1) holds. Let x, yl().

  • (1)

    δ(J0y) = J0τ(y),   τ(J0y) = J0δ(y).

  • (2)

    For any s, k,

    ()

  • (3)

    For any λ, c0, and any two solutions x(t) and y(t) of (1λ), it follows that

    ()

Moreover, let Y(t, λ) be a fundamental solution of (1λ), then

()

3.2. Relationship between the Maximal and Minimal Subspaces

In this subsection, we first introduce the maximal and minimal subspaces corresponding to (1λ) and then show that the minimal subspace is J-Hermitian, and its J-adjoint subspace is just the maximal subspace.

Denote
()
and define
()
It can be easily verified that H(τ) and H00(τ) are both linear subspaces in . Here, H(τ) and H00(τ) are called the maximal and preminimal subspaces corresponding to τ or (1λ) in , and is called the minimal subspace corresponding to τ or (1λ) in .
Since the end points a and b may be finite or infinite, we need to divide into two subintervals in order to characterize the maximal and minimal subspaces in a unified form. Choose a < c0 < b and fix it. Denote
()
and denote by 〈·, ·〉 a, 〈·, ·〉 b and ∥·∥a, ∥·∥b the inner products and norms of , , respectively. Let and be defined by (31) with replaced by a and b, respectively. Furthermore, let Ha(τ) and Ha,00(τ) be the left maximal and preminimal subspaces defined by (32) with replaced by a, respectively, and Hb(τ) and Hb,00(τ) the right maximal and preminimal subspaces defined by (32) with replaced by b, respectively. The subspaces and are called the left and right minimal subspaces corresponding to system (1λ) in a and b, respectively. Similarly, we can define H(δ), H00(δ), and H0(δ); Hb(δ), Hb,00(δ), and Hb,0(δ); Ha(δ), Ha,00(δ) and Ha,0(δ).

The following result is directly derived from (1) of Lemma 14.

Lemma 15. Assume that (A1) holds. Then (yπ, gπ) ∈ H(τ) if and only if (Jyπ, Jgπ) ∈ H(δ).

In order to study properties of the above subspaces, we first make some preparation.

Let Y(t) be the fundamental solution matrix of (10) with Y(c0) = I2n. For any finite subinterval with c0, denote
()
It is evident that Φ is a 2n × 2n positive semidefinite matrix and dependent on . By the same method used in [23, Lemma 3.2], it follows that there exists a finite subinterval 0 with c00 such that
()
for any finite subinterval    with 0. In the present paper, we will always denote 0 : = [s0, t0] and define
()
whenever is finite or infinite. In the case that is finite, 0 can be taken as .
In the case that is finite, we define
()
It is evident that ϕ is a bounded linear map and its range is a closed subset in 2n.
In the case that is infinite, that is, = [a, +) or = (−, b] or = (−, +), where a, b are finite integers, we introduce the following subspaces of , respectively:
()
()
()
It can be easily shown that is dense in . In this case, we define
()
By the method used in [23, Lemma 3.3], one has the following properties of ϕ.

Lemma 16. Assume that (A1) holds.

  • (1)

    Ran  ϕ = Ran  Φ.

  • (2)

    In the case that is finite,

    ()

  • in the case that is infinite, let l = rank  Φ. Then there exist linearly independent elements , 1 ≤ jl, such that

    ()

  • (3)

    Ker  ϕ ⊂ Ran (H00(τ)).

The following is the main result of this section.

Theorem 17. Assume that (A1) holds. Then , , and .

Proof. Since the method of the proofs is similar, we only show the first assertion. By , it suffices to show .

We first show that . Let (yπ, gπ) ∈ H(τ). Then for any (xπ, fπ) ∈ H00(τ), there exists xxπ with such that τ(x)(t) = W(t)R(f)(t) in . So, it follows from (2) of Lemma 14 that

()
This implies that .

Next, we show . Fix any . It suffices to show that there exists y0yπ such that τ(y0)(t) = W(t)R(g)(t) in . Let z be a solution of τ(z)(t) = W(t)R(g)(t) on . For any (xπ, fπ) ∈ H00(τ), there exits xxπ with such that τ(x)(t) = W(t)R(f)(t) in . Thus, it follows from (2) of Lemma 14 that

()
In addition, it is clear that
()
Combining (45) and (46), one has that for all (xπ, fπ) ∈ H00(τ),
()
By (2) and (3) of Lemma 16, we get that for any hπ ∈ Ker  ϕ and any ξ2n
()

The following discussion is divided into two parts.

Case 1. is finite. It is evident that . Then, from (2) of Lemma 16, there exists ξ0 ∈ Ran  Φ such that (J0(yzYξ0)) π ∈ Ker  ϕ. This, together with (48), implies that (J0(yzYξ0)) π = 0. This is equivalent to (yzYξ0) π = 0. Let y0(t): = z(t) + Y(t)ξ0. Then y0yπ and satisfies

()
Hence, (yπ, gπ) ∈ H(τ). Since is arbitrary, we have .

Case 2. is infinite. We only consider the case that = [a, +). For the other two cases, it can be proved with similar arguments.

Let rank  Φ = l. With a similar argument as Case 2 of the proof of [23, Theorem 3.1], it can be shown that there exist linearly independent elements , and ξ0 ∈ Ran  Φ such that (yzYξ0) π ∈ Ker  ϕ,

()
Combining (48)–(50), one has that for any
()
This implies that and consequently y0 : = z + Yξ0 is a representative of yπ such that τ(y0)(t) = W(t)R(g)(t). So (yπ, gπ) ∈ H(τ). By the arbitrariness of (yπ, gπ) one has .

The entire proof is complete.

The following result is directly derived from Lemmas 5 and 15, and Theorem 17.

Theorem 18. Assume that (A1) holds. Then , , and .

3.3. Definiteness Condition

In this subsection, we introduce the definiteness condition for (1λ), and give some important results on it. Since the proofs are similar to those given in [23], we omit the proofs.

The definiteness condition for (1λ) or H(τ) is given by the following.
  • (A2) There exists a finite subinterval 1 such that for any λ and for any nontrivial solution y(t) of (1λ), the following always holds:

    ()

In particular, the definiteness condition for (3λ) can be described as there exists a finite subinterval 1 such that for any λ and for any nontrivial solution z(t) of (3λ), the following always holds:
()

Lemma 19. Assume that (A1) holds. Then (A2) holds if and only if there exists a finite subinterval 1 such that one of the following holds:

  • (1)

    ;

  • (2)

    for some λ, every nontrivial solution y(t) of (1λ) satisfies

    ()

By Lemma 19, if (52) (or (53)) holds for some λ, then it holds for every λ. In addition, if (A2) holds on some finite interval 1, then it holds on 0 = [s0, t0].

The following is another sufficient and necessary condition for the definiteness condition.

Lemma 20. Assume that (A1) holds. Then (A2) holds if and only if for any (yπ, gπ) ∈ H(τ), there exists a unique yyπ such that τ(y)(t) = W(t)R(g)(t) for t.

Remark 21. (1) It can be easily verified that the definiteness condition for H(τ) holds if and only if that for H(δ) holds.

(2) In the following of the present paper, we always assume that (A2) holds. In this case, we can write (y, gπ) ∈ H(τ) instead of (yπ, gπ) ∈ H(τ) in the rest of the present paper.

(3) Denote by (Aa,2) and (Ab,2), the definiteness conditions for (1λ) in a and b, and

()
By the corresponding intervals, respectively. It is evident that one of (Aa,2) and (Ab,2) implies (A2).

But (A2) cannot imply that there exists c0 such that both (Aa,2) and (Ab,2) hold.

(4) Several sufficient conditions for the definiteness condition can be given. The reader is referred to [23, Section 4].

For convenience, denote

()

Lemma 22. Assume that (A1) holds. For any λ, dim  λ = dim  Mλ if and only if (A2) holds.

4. Characterizations of Minimal and Maximal Subspaces and Defect Indices of Minimal Subspaces

This section is divided into three subsections. In the first subsection, we give all the characterizations of the minimal subspaces generated by (1λ) in , a, and b. In the second subsection, we study the defect indices of the minimal subspaces. In the third subsection, characterizations of the maximal subspaces are established.

4.1. Characterizations of the Minimal Subspaces

In this subsection, we study characterizations of the minimal subspaces generated by (1λ) in , a, and b.

The following result is a direct consequence of Theorem 17.

Theorem 23. Assume that (A1) holds. Then H0(τ), Hb,0(τ), and Ha,0(τ) are closed J-Hermitian subspace in , , and , respectively.

Now, we introduce boundary forms on , , and by
()

Lemma 24. Assume that (A1) holds.

  • (1)

    If (A2) holds, then for any (x, fπ), (y, gπ) ∈ H(τ),

    ()

  • (2)

    If (Aa,2) holds, then for any (x, fπ), (y, gπ) ∈ Ha(τ),

    ()

  • (3)

    If (Ab,2) holds, then for any (x, fπ), (y, gπ) ∈ Hb(τ),

    ()

Proof. Since the proofs of (1)–(3) are similar, we only show that assertion (1) holds.

For any (x, fπ), (y, gπ) ∈ H(τ), we have from (2) of Lemma 14 that

()
for any s < k. This yields that lim tb (x, y)(t) exists and is finite for any (x, fπ), (y, gπ) ∈ H(τ). Similarly, it can be shown that lim ta (x, y)(a) exists and is finite for any (x, fπ), (y, gπ) ∈ H(τ). Hence, assertion (1) holds. The proof is complete.

Lemma 25. Assume that (A1) and (A2) hold. Then for any given finite subset 1 = [s, k] with 01 and for any given α, β2n, there exists such that the following boundary value problem:

()
has a solution .

Proof. Set

()
Let ϕj,   1 ≤ j ≤ 2n, be the linearly independent solutions of system (10). Then we have
()
In fact, the linear algebraic system
()
where C = (c1, c2, …, c2n) T2n, can be written as
()
which yields
()
Since is a solution of system (10), it follows from (A2) that . Then C = 0; that is, (65) has only a zero solution. Consequently, (64) holds.

Let α, β be any given vectors in 2n. By (64), the linear algebraic system

()
has a unique solution C12n. Set ψ1 = (ϕ1, ϕ2, …, ϕ2n)C1. It follows from (68) that
()
Let u(t) be a solution of the following initial value problem:
()
Since τ(ϕi)(t) = 0 for t and 1 ≤ i ≤ 2n, we get by (70) and (2) of Lemma 14 that
()
Since ϕ1, ϕ2, …, ϕ2n are linearly independent in l(), we get from (69) and (71) that u(k + 1) = β. So, u(t) is a solution of the following boundary value problem:
()

On the other hand, the linear algebraic system

()
has a unique solution C22n by (64). Set ψ2 = (ϕ1, ϕ2, …, ϕ2n)C2. Then, by (73)
()
Let v(t) be a solution of the following initial value problem:
()
Since τ(ϕi)(t) = 0 for t and 1 ≤ i ≤ 2n, we get by (2) of Lemma 14 and (75) that
()
which, together with (74), implies that v(s) = α. So, v(t) is a solution of the following boundary value problem:
()
Set ψ = ψ1ψ2 and ϕ = u + v. Then ϕ is a solution of the boundary value problem (62). The proof is complete.

Remark 26. Lemma 25 is called a patch lemma. Based on Lemma 25, any two elements of H(τ) (Hb(τ), Ha(τ), resp.) can be patched up to construct another new element of H(τ) (Hb(τ), Ha(τ), resp.). In particular,

  • (1)

    if (Ab,2) holds, we can take , , and β = 0, 1 ≤ i ≤ 2n. Then there exist satisfying

    ()

  • (2)

    if (Aa,2) holds, we take , α = 0, and β = ei. Then there exist satisfying

    ()

  • (3)

    if both (Ab,2) and (Aa,2) hold, then there exist satisfying

    ()

The above auxiliary elements , , and will be very useful in the sequent discussions.

Theorem 27. Assume that (A1) holds.

  • (1)

    If (A2) holds, then

    ()
    In particular, if = [a, b], then
    ()

  • (2)

    If (Ab,2) holds, then

    ()

  • (3)

    If (Aa,2) holds, then

    ()

Proof. We first show that assertion (1) holds. By Lemmas 8 and 24, and Theorem 17, one has

()
For convenience, denote
()
Clearly, H0(τ) ⊂ H0(τ). We now show that H0(τ) ⊂ H0(τ). Fix any (x, fπ) ∈ H0(τ). It follows from (85) that for all (y, gπ) ∈ H(τ),
()
For any given (y, gπ) ∈ H(τ), by Remark 26 there exists (z, hπ) ∈ H(τ) such that
()
Thus, it follows from (87) that (x, y)(b + 1) = (x, z)(b + 1) = (x, z)(a) = 0, and consequently (x, y)(a) = 0 for all (y, gπ) ∈ H(τ).

In the case that = [a, b], it is clear that

()
So it remains to show that H0(τ) = H00(τ). It suffices to show that x(a) = x(b + 1) = 0 for any (x, fπ) ∈ H0(τ). Fix any (x, fπ) ∈ H0(τ), and let 0 = , α = ei, 1 ≤ i ≤ 2n, and β = 0. Then by Lemma 25, there exist with yi(a) = ei and yi(b + 1) = 0. Inserting these yi into (87) one has that x(a) = 0. Similarly, one can show that x(b + 1) = 0. Thus H0(τ) = H00(τ). Therefore, assertion (1) has been shown.

With similar arguments, one can show that assertion (2) and (3) hold by using (78) and (79), separately. This completes the proof.

4.2. Defect Indices of Minimal Subspaces

In this subsection, we first give a valued range of the defect indices of Hb,0(τ) and Ha,0(τ) and then discuss the relationship among the defect indices of Hb,0(τ), Ha,0(τ), and H0(τ).

For briefness, denote
()
For any λ, let b,λ, Mb,λ, a,λ, and Ma,λ be defined as (56) with replaced by b and a, respectively.

The following results are obtained.

Theorem 28. Assume that (A1) holds.

  • (1)

    If (Ab,2) holds and Γ(Hb,0(τ)) ≠ , then db = dim  b,λ for any λ ∈ Γ(Hb,0(τ)), and ndb ≤ 2n.

  • (2)

    If (Aa,2) holds and Γ(Ha,0(τ)) ≠ , then da = dim  a,λ for any λ ∈ Γ(Ha,0(τ)), and nda ≤ 2n.

Proof. Since the method of the proofs is the same, we only give the proof of assertion (1).

For any λ ∈ Γ(Hb,0(τ)), it follows from Lemma 11 and Theorem 18 that

()
On the other hand, by using Lemma 5 and Theorems 17 and 18, one has that
()
It is clear that
()
Combining (91)–(93), one has db = dim  Mb,λ. This, together with Lemma 22, implies that db = dim  b,λ. In addition, it has been shown in [21] that n ≤ dim  b,λ ≤ 2n for any λ. So assertion (1) is true. The proof is complete.

Next, we discuss relationship among defect indices of H0(τ), Ha,0(τ), and Hb,0(τ). For convenience, denote
()
for any . It is evident that , , and
()
On the other hand, for any and , we define y by
()
Then (95) still holds, and consequently . Furthermore, it is clear that
()

The following result can be easily verified. So we omit its proof.

Lemma 29. Assume that (A1), (Aa,2), and (Ab,2) hold.

  • (1)

    If (y, gπ) ∈ H(τ), then and .

  • (2)

    If and with ya(c0) = yb(c0), then (y, gπ) ∈ H(τ).

Let be the restriction of subspace H0(τ), defined by
()

Lemma 30. Assume that (A1), (Aa,2), and (Ab,2) hold. Then

  • (1)

    if and only if and ;

  • (2)

    if and only if and .

Proof. (1) We first consider the necessity. Fix any . Let ya, yb, ga, and gb be defined as (94). Then ya(c0) = yb(c0) = 0. By (1) of Lemma 29 one has that . In addition, for any , it follows from Remark 26 that there exits (z, hπ) ∈ H(τ) with

()
So one has
()
which implies that by the arbitrariness of and (3) of Theorem 27. With a similar argument, one can show .

Next, we consider the sufficiency. Fix any and . Let y and g be defined by (96). By (2) and (3) of Theorem 27 one has that ya(c0) = yb(c0) = 0. So y(c0) = 0. It follows from (2) of Lemma 29 that (y, gπ) ∈ H(τ). For any (x, fπ) ∈ H(τ), it follows from (1) of Lemma 29 that and . Thus one has by (2) and (3) of Theorem 27 that

()
This implies that (y, gπ) ∈ H0(τ) by (1) of Theorem 27, and consequently, .

(2) We first consider the necessity. Fix any . Let ya, yb, ga, and gb be defined as (94). Then and .

Set xb(t) = fb(t) = 0 for , where is defined by (19) with replaced by b. It is clear that . For any , let x and f be defined by (96). Then by the above result (1) one has that . It follows that

()
By the arbitrariness of , one has that by Theorem 17. With a similar argument one can show .

Next, we consider the sufficiency. Fix any and . Let y and g be defined by (96). For any , it follows from the above result (1) that and . So one has by Theorem 17 that

()
This yields that . The whole proof is complete.

Lemma 31. Assume that (A1), (Aa,2), and (Ab,2) hold. Then and .

Proof. The first assertion holds because , and the second assertion can be proved by (1) of Lemma 30 and (95). The proof is complete.

In the following of the present paper, we assume that
  • (A3) Γ(H0(τ)) ≠ .

By Definition 10, one has that if Γ(H0(τ)) = , then ρ(H0(τ)) = , and consequently σ(H0(τ)) = . We do not consider this case in the present paper.

Theorem 32. Assume that (A1), (Aa,2), (Ab,2), and (A3) hold. Then

()

Proof. The proof is divided into two steps.

Step  1. We show that

()

It follows from Γ(H0(τ)) ≠ and Lemma 31 that Γ(Hb,0(τ))∩Γ(Ha,0(τ)) ≠ . This, together with (Aa,2), (Ab,2), and Theorem 28, implies that for any λ ∈ Γ(Hb,0(τ))∩Γ(Ha,0(τ)), (1λ) has just db linearly independent solutions , 1 ≤ jdb, and (1λ) has just da linearly independent solutions , 1 ≤ jda; that is, for 1 ≤ jdb and for 1 ≤ jda.

Set for . It is clear that for 1 ≤ jda. Let xj, 1 ≤ jda, be defined by (96). Then it follows from (2) of Lemma 30 that . Similarly, set for . Then one has for 1 ≤ jdb. It is evident that are linearly independent.

On the other hand, for any , it follows from (2) of Lemma 31 that and ; that is, is a solution of (1λ) in a, and is a solution of (1λ) in b. Therefore, there exist unique cj  (1 ≤ jda) and dj  (1 ≤ jdb) such that

()
Noting the constructions of xj and yj by (96), one has that
()
This, together with Lemma 11, implies that (105) holds.

Step  2. We show that

()

It is evident that , 1 ≤ i ≤ 2n, where are defined by (80). We claim that

()
In fact, for each (y, gπ) ∈ H0(τ), set C0 = (c1, c2, …, c2n) T = y(c0). Let
()
Then
()
It can be easily verified that this decomposition is unique. Hence, (109) holds.

For any λ ∈ Γ(H0(τ)), we claim that

()
Since (109) holds, it suffices to show that
()
Suppose that and c1, c2, …, c2n satisfy
()
that is,
()
Since and , it follows that . Since λ ∈ Γ(H0(τ)), it yields
()
which, together with (109), implies that cj = 0 for 1 ≤ j ≤ 2n, and consequently gπ = yπ = 0. This yields that (113) holds.

Since H0(τ) and are closed J-Hermitian subspaces, it follows that Ran (H0(τ) − λ) and are closed subspaces in , respectively. Hence, there exists a closed subspace Q in such that

()
In addition, again by the fact that λ ∈ Γ(Hb,0(τ))∩Γ(Ha,0(τ)), it follows that are linearly independent in . It follows from (113) that dim  Q = 2n. Consequently,
()
This yields that (108) holds. It follows from (105) and (108) that (104) holds. The proof is complete.

Remark 33. Theorem 32 (formula (104)) generalizes the classical result for 2nth order ordinary differential equations that go back to the classical work by Akhiezer and Glazman [1, Theorem 3 in Appendix 2]. To the case of symmetric Hamiltonian systems, formula (104) was extended in [15].

So it follows from Theorems 28 and 32 that d = 0 if and only if da = db = n, and d = 2n if and only if da = db = 2n. The following definition is obtained.

Definition 34. Assume that (A1), (A2), and (Ab,2) hold. Then (1λ) is said to be in the limit db case at t = b. In the special case that db = n, (1λ) is said to be in the limit point case (l.p.c.) at t = b, and in the other special case that db = 2n, (1λ) is said to be in the limit circle case (l.c.c.) at t = b.

The same definition can be given at t = a provided that (Aa,2) holds.

4.3. Characterizations of Ha(τ) and Hb(τ)

In this subsection, we characterize the maximal subspaces Ha(τ) and Hb(τ). We first consider characterization of Hb(τ). Assume that (Ab,2) holds and Γ(Hb,0(τ)) ≠ . Let λ ∈ Γ(Hb,0(τ)). It follows from the proof of Lemma 11 that
()
where and , and are linearly independent (mod U1). For convenience, denote
()
Clearly, , 1 ≤ i ≤ 2db. So is finite for all 1 ≤ i, j ≤ 2db. Then the following result can be directly derived from (119) and (120).

Lemma 35. Assume that (A1) and (Ab,2) hold, and Γ(Hb,0(τ)) ≠ . Then every y ∈ Dom  Hb(τ) can be expressed as

()
where and ai.

By Lemma 35, , defined by (78), can be uniquely expressed as
()
where and aij. Denote
()

Lemma 36. Assume that (A1) and (Ab,2) hold, and Γ(Hb,0(τ)) ≠ . Then rank  E = 2n and rank   F1 = 2db − 2n. Furthermore, we can rearrange the order of such that

()

Proof. It follows from (122) that

()
which, together with (78), implies that
()
By Lemma 12, one has rank  E = 2n.

On the other hand, it follows from (122) that

()
for 1 ≤ i ≤ 2n, 1 ≤ s ≤ 2db, which, together with (78) and (2) of Theorem 27, implies that
()
Noting that rank  E = 2n, one has
()
We now want to show rank  F1 ≥ 2db − 2n. By (3) of Lemma 14 one has
()
where . Since rank  X0 = db, we have that
()
which, together with (129), yields that
()
Because rank  F2 = 2db − 2ndb, one can rearrange the order of such that the first 2db − 2n rows of F2 are linearly independent; that is, (124) holds. The proof is complete.

Without loss of generality, we assume that (124) holds in the rest of this paper. Now, we can give a characterization of Hb(τ).

Theorem 37. Assume that (A1) and (Ab,2) hold, λ ∈ Γ(Hb,0(τ)) ≠ , and are linearly independent solutions of (1λ) in such that (124) holds. Then

()
is invertible, and any y ∈ Dom  Hb(τ) can be uniquely expressed as
()
where , are defined by (78), and ci, dj.

Proof. Let E = (E1, E2), where E1 and E2 are 2n × (2db − 2n) and 2n × 2n matrices, respectively. It follows from (124) that there exists an invertible matrix L such that

()
So, it follows from (128) that E1 + E2F3 = 0, which is equivalent to E1 = −E2F3. Since rank  E = 2n, it follows that E2 is invertible. Multiplying (125) by from the right-hand side, we get
()
This implies that each of can be uniquely expressed as a linear combination of , , and . Therefore, (134) follows from Lemma 35.

Since for 1 ≤ jdb, can be uniquely expressed as

()
where , and bjl, cjs. This, together with (78) and (2) of Theorem 27, implies that for 1 ≤ i ≤ 2db − 2n,   1 ≤ jdb
()
that is,
()
Therefore, Gb is invertible from (124). This completes the proof.

With a similar argument, one can obtain the following characterization of Ha(τ).

Theorem 38. Assume that (A1) and (Aa,2) hold, Γ(Ha,0(τ)) ≠ . Then, for some λ ∈ Γ(Ha,0(τ)), system (1λ) has 2da − 2n linearly independent solutions in such that Ga is invertible, where

()
and each y ∈ Dom  Ha(τ) can be uniquely expressed as
()
where , ci, dj, and are defined as (79).

5. Characterizations of J-SSEs of H0(τ)

In this section, we give a complete characterization of all the J-SSEs of minimal subspace H0(τ) in terms of the square summable solutions of system (1λ). As a consequence, characterizations of all the J-self-adjoint subspace extensions are obtained in the two special cases: the limit point and limit circle cases. The following discussion is divided into two parts based on the form of .

5.1. Both the Endpoints Are Infinite

Let = (−, +), and assume that (A1), (Aa,2), and (Ab,2) hold, and Γ(H0(τ)) ≠ in this subsection. It follows from Theorem 32 that
()
In addition, for some λ ∈ Γ(H0(τ), let given in Theorems 37 and be given in Theorems 38.

Theorem 39. Assume that (A1), (Aa,2), (Ab,2), and (A3) hold. Then a subspace is a J-SSE of H0(τ) if and only if there exist two matrices and such that

  • (1)

    rank  (M, N) = d,

  • (2)

    NGbNTMGaMT = 0, and

    ()

where Gb and Ga are the same as those in Theorems 37 and 38.

Proof. We first show the sufficiency.

Suppose that there exist two matrices and such that conditions (1) and (2) hold and T is defined by (143). We now prove that T is a J-self-adjoint subspace extension of H0(τ) by Lemma 9.

Denote

()
and set
()
Clearly, and for 1 ≤ id. By Remark 26, there exist such that
()
where and are specified by (Aa,2) and (Ab,2), respectively.

Since H(τ) and H0(τ) are liner subspaces, β1, β2, …, βd are linearly independent in H(τ) (modulo H0(τ)) if and only if ω1, ω2, …, ωd are linearly independent in Dom  H(τ) (modulo Dom  H0(τ)). So it suffices to show that ω1, ω2, …, ωd are linearly independent in Dom  H(τ) (modulo Dom  H0(τ)). Suppose that there exists C = (c1, c2, …, cd) ∈ d such that

()
It follows from (145), (146), and (2) and (3) of Theorem 27 that
()
Since Gb and Ga are invertible, we get from (148) that CM = CN = 0. Then C = 0 by condition (1). So, ω1, ω2, …, ωd are linearly independent in Dom  H(τ) (modulo Dom  H0(τ)), and consequently β1, β2, …, βd are linearly independent in H(τ) (modulo H0(τ)).

Next, we show that [βi : βj] = 0 for 1 ≤ i, jd. It follows from (145) and (146) that

()
which, together with Lemma 24 and condition (2), implies that
()
Consequently, [βi : βj] = 0 for 1 ≤ i, jd. Therefore, satisfy the conditions (1) and (2) of Lemma 9.

Note that for each y ∈ Dom  H(τ), it follows that

()
where (145) and (146) have been used. Therefore, it follows from Lemma 24 that T can be expressed as
()
Hence, T is a J-SSE of H0(τ) by Lemma 9. The sufficiency is proved.

We now show the necessity. Suppose that T is a J-SSE of H0(τ). By Lemma 9 and Theorem 17, there exists a set of such that (152) holds. Write . Then and for 1 ≤ jd. By Theorems 37 and 38, each and can be uniquely expressed as

()
where , , and dij, cij, njs, mjs. Set
()

First, we want to show that M and N satisfy condition (1). Otherwise, suppose that rank  (M, N) < d. Then there exists C = (c1, c2, …, cd) ∈ d with C ≠ 0 such that

()
Then CM = 0 and CN = 0. Let . We then have
()
For each y ∈ Dom  H(τ), yb can be uniquely expressed as (134) by Theorem 37 and ya can be uniquely expressed as (141) by Theorem 38. So it follows from (156), (78), (79), and (2) and (3) of Theorem 27 that for any y ∈ Dom  H(τ), (ω, y)(+) = (ω, y)(−) = 0, which yields that ω ∈ Dom  H0(τ) by (1) of Theorem 27. Then ω1, ω2, …, ωd are linearly dependent in Dom  H(τ) (modulo Dom  H0(τ)), and consequently β1, β2, …, βd are linearly dependent in H(τ) (modulo H0(τ)). This is a contradiction. Hence, rank  (M, N) = d.

Next, we prove that M and N satisfy condition (2). It can be easily verified that

()
Hence, by Lemma 14 and [βi : βj] = 0 for 1 ≤ i, jd, M and N satisfy condition (2).

In addition, it follows from (78), (79), (153), and (2) and (3) of Theorem 27 that

()
Hence, T in (152) can be expressed as (143). The necessity is proved.

The entire proof is complete.

To end this subsection, we give characterizations of J-SSEs of H0(τ) in four special cases of defect indices: da = db = n; da = n, db = 2n; da = 2n, db = n; da = db = 2n.

In the case that da = db = n, that is, d = 0 by Theorem 32, the following result is derived from Lemma 11 and Theorem 17.

Theorem 40. Assume that (A1), (Aa,2), (Ab,2), and (A3) hold. If d = 0, then H0(τ) is a J-self-adjoint subspace.

In the case that da = n, db = 2n, it follows from Theorem 32 that d = n. Let , (1 ≤ i ≤ 2n), be 2n linearly independent solutions in of (1λ) satisfying . Then, by (3) of Lemma 14 one has that
()
The following result can be directly derived from Theorem 39.

Theorem 41. Assume that (A1), (A2), (Aa,2), (Ab,2), and (A3) hold. If (1λ) is in l.p.c. at t = − and in l.c.c. at t = +. Let be 2n linearly independent solutions in of (1λ) satisfying . Then a subspace is a J-SSE of H0(τ) if and only if there exists a matrix Nn×2n such that

()

In the case that da = 2n, db = n, a similar result can be easily given. So we omit the details in this case.

In the case that da = db = 2n, it follows from Theorem 32 that d = 2n. Let χi  (1 ≤ i ≤ 2n) be 2n linearly independent solutions in of (1λ) satisfying
()
Then, by Lemma 14, Gb and Ga, defined by (133) and (140), satisfy
()
The following result is a direct consequence of Theorem 39.

Theorem 42. Assume that (A1), (Aa,2), (Ab,2), and (A3) hold. If (1λ) is in l.c.c. at both t = + and t = −, then for any given λ ∈ Γ(H0(τ)), let χi  (1 ≤ i ≤ 2n) be 2n linearly independent solutions in of (1λ) satisfying (χ1(c0), χ2(c0), …, ϕ2n(c0)) = I2n. Then a subspace is a J-SSE of H0(τ) if and only if there exist two matrices M2n×2n and N2n×2n such that

()
()

Remark 43. As we have seen, there is no boundary condition at the endpoints at which system (1λ) is in l.p.c., and the matrix Ga or Gb can be replaced by 𝒥 in the case that system (1λ) is in l.c.c. at t = a or t = b.

5.2. At Least One of the Two Endpoints Is Finite

In this subsection, we characterize the J-SSEs of H0(τ) in the special case that at least one of the two endpoints a and b is finite. We first consider the case that a is finite, and b is finite or infinite.

We point out that in this case, characterizations of all the J-SSEs of H0(τ) can also be given by the proof of Theorem 39, provided that assumptions (Ab,2), (Aa,2), and (A3) hold. But, if there does not exist a c0 such that both (Aa,2) and (Ab,2) are satisfied, then Theorem 39 fails. We will remark again that the division of is not necessary in the case that one of the two endpoints is finite, and characterizations of all the J-SSEs of H0(τ) can still be given provided that (A3) and (A3) hold.

In the case that a is finite, can be regarded as b with a = c0, and (A2) is equivalent to (Ab,2). So all the characterizations for Hb,0(τ) and Hb(τ) given in Sections 3 and 4 are available to H0(τ) and H(τ), respectively, with c0 replaced by a. Assume that (A2) holds. Then for any given λ ∈ Γ(H0(τ)), as discussed in Section 4, let χ1, …, χ2d−2n be 2d − 2n linearly independent solutions in of (1λ) such that G is invertible, where G is defined by (133) with is replaced by χj, 1 ≤ j ≤ 2d − 2n. Then all the results of Theorem 37 hold with Gb and replaced by G and χj, respectively.

Theorem 44. Assume that the left endpoint a is finite, and (A1)–(A3) hold. Then a subspace is a J-SSE of H0(τ) if and only if there exist two matrices Md×2n and Nd×(2d−2n) such that

  • (1)

    rank  (M, N) = d,

  • (2)

    M𝒥MTNGNT = 0, and

    ()

Proof. The main idea of the proof is similar to that of Theorem 39.

We first show the sufficiency. Denote

()
and set
()
Clearly, wi ∈ Dom  H(τ) for 1 ≤ id. By Remark 26, there exist βi : = (ωi, ρπ) ∈ H(τ) (1 ≤ id) such that
()
where t0 is specified by (A2). It can be verified by the method used in the proof of Theorem 39 that the set satisfies the conditions (1) and (2) in Lemma 9. Note that for each y ∈ Dom  H(τ), it follows that
()
Therefore, it follows from Lemma 24 that T can be expressed as
()
Hence, T is a J-SSE of H0(τ) by Lemma 9. The sufficiency is proved.

We now show the necessity. Suppose that T is a J-SSE of H0(τ). By Lemma 9 and Theorem 17, there exists a set of for {H0(τ), H(τ)} such that conditions (1) and (2) in Lemma 9 hold, and T can be expressed as (152). Write βi : = (ωi, ρπ). Then ωj ∈ Dom  H(τ) for 1 ≤ jd. By Theorem 37, each ωi can be uniquely expressed as

()
where and aij, nis. Set
()

With a similar argument to that used in the proof of Theorem 39, we can prove that M and N satisfy conditions (1) and (2). In addition, it is clear that T in (170) can be expressed as (165). The necessity is proved, and then the entire proof is complete.

At the end of this subsection, we give the characterizations of J-SSEs of H0(τ) in two special cases of defect indices.

In the special case that d = n, Theorem 44 can be described in the following simpler form.

Theorem 45. Assume that the left endpoint a is finite, and (A1)–(A3) hold. If (1λ) is in l.p.c. at t = b, then a subspace is a J-SSE of H0(τ) if and only if there exists a matrix Mn×2n satisfying the self-adjoint condition

()
such that T can be defined by
()

In the other special case that d = 2n, the following result is a direct consequence of Theorem 44.

Theorem 46. Assume that the left endpoint a is finite, and (A1)–(A3) hold. If (1λ) is in l.c.c. at t = b, let χi  (1 ≤ i ≤ 2n) be 2n linearly independent solutions in of (1λ) satisfying (χ1(c0), χ2(c0), …, χ2n(c0)) = I2n. Then a subspace is a J-SSE of H0(τ) if and only if there exist two 2n × 2n matrices M and N such that

  • (1)

    rank  (M, N) = 2n,

  • (2)

    M𝒥MT = N𝒥NT, and

    ()

For the case that b is finite and a = −, it can be considered by a similar method. Here we only give the following basic result.

Theorem 47. Assume that the right endpoint b is finite, and (A1)–(A3) hold. Let χj, 1 ≤ j ≤ 2d − 2n, be linearly independent solutions of (1λ) in such that G is invertible, where G is defined as Ga in Theorem 38 with replaced by χj. Then a subspace is a J-SSE of H0(τ) if and only if there exist two matrices Mn×2n and Nd×(2d−2n) such that

  • (1)

    rank  (M, N) = d,

  • (2)

    MGMTN𝒥NT = 0, and

    ()

In the case that both the two endpoints a and b are finite, that is, = [a, b], it is clear that d = 2n by (A2). The characterization of J-SSEs given in Theorem 44 can be simplified as follows.

Theorem 48. Let = [a, b]. Assume that (A1)–(A3) hold. Then a subspace is a J-SSE of H0(τ) if and only if there exist two 2n × 2n matrices M and N such that

  • (1)

    rank  (M, N) = 2n,

  • (2)

    N𝒥NT = M𝒥MT, and

    ()

Proof. Let , 1 ≤ j ≤ 2n, be defined as those in Theorem 46. Then it follows that

()
where N1 = N(χ1, …, χ2n) T(b + 1)𝒥. It is evident (χ1, …, χ2n)(b + 1) = I2n. So by Lemma 14, one has that
()
Hence, the assertion follows from Theorem 46. This completes the proof.

Acknowledgment

This research was supported by the NNSF of China (Grants 11071143, 11101241, and 11226160).

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