Volume 2022, Issue 1 4288928
Research Article
Open Access

Reducing Subspaces for Toeplitz Operator on the Weighted Hardy Space over the Bidisk

Changguo Wei

Changguo Wei

School of Mathematical Sciences, Ocean University of China, Qingdao, Shandong, China ouc.edu.cn

Search for more papers by this author
Xin Ding

Xin Ding

School of Mathematical Sciences, Ocean University of China, Qingdao, Shandong, China ouc.edu.cn

Search for more papers by this author
Yanyue Shi

Corresponding Author

Yanyue Shi

School of Mathematical Sciences, Ocean University of China, Qingdao, Shandong, China ouc.edu.cn

Search for more papers by this author
First published: 26 March 2022
Citations: 1
Academic Editor: Sivaram K. Narayan

Abstract

In this paper, we completely characterize the reducing subspaces for on weighted Hardy space under three assumptions on ω, where , k, l2,  kl, and a ∈ (0,1). It is shown that the coefficient a ∈ (0,1) does not affect the reducing subspaces for . We also prove that, for every δ > 0, weighted Dirichlet space is a weighted Hardy space which satisfies these assumptions. As an application, we describe the reducing subspaces for on and get the structure of commutant algebra .

1. Introduction

Let ℳ be a closed subspace of a complex Hilbert space ℋ; ℳ is said to be an invariant subspace for AB(ℋ) if Aℳ ⊂ ℳ. The subspace ℳ is said to be a reducing subspace for A if both ℳ and ℳ are invariant under A. Let P be the orthogonal projection from ℋ onto ℳ; then, ℳ is a reducing subspace for A if and only if , where .

For a bounded analytic function φ over disk , let Mφ denote the multiplication operator with the symbol φ on Bergman space over the disk, given by Mφf = φf with . Sun and Wang [1] described the commutators and reducing subspaces for multiplication operator whose symbol B2 is the product of two Blaschke factors. Zhu [2] also independently characterized the reducing subspaces for multiplication operator . Let B be a finite Blaschke product. The reducing subspaces for MB have been completely described with the joint efforts of many scholars (see [3, 4] and their references).

On some familiar function spaces such as weighted Bergman space and Dirichlet-type space over the polydisk, the reducing subspaces for Toeplitz operator Tφ have been described when φ is a polynomial. Some details can be seen in [512] and their references.

Gu [13] firstly described the reducing subspaces for operator-weighted shifts on weighted Hardy space . In fact, many known function spaces are all weighted Hardy spaces, such as weighted Bergman space and weighted Dirichlet space. So, Gu provided a uniform way to study the reducing subspaces for multiplication operators on these function spaces. Moreover, Gu [14] characterized the reducing subspaces for with a ∈ (0,1] on under five assumptions on ω. He showed that the coefficient a does not affect the structure of reducing subspaces. Recently, Ren and Shi [15] characterized the reducing subspaces for on .

In this paper, we keep on considering reducing subspaces for Tφ with on . As in [14], we also pay attention to the case that a ∈ (0,1). A complete characterization is given under three assumptions on ω, which also shows that coefficient a does not affect the structure of reducing subspaces.

In order to state the results more precisely, we recall some notations. Let be the set of all positive integers and + be the set of all nonnegative integers. Given a sequence of positive numbers ω = {ωj : j+} such that . Denote . The weighted Hardy space over the bidisk with weights ω is defined by
(1)
where . Set ; then, for every ,
(2)
Denote Qn(p) ≡ 0 if Qn(p) = 0,  ∀ p. Otherwise, we write Qn(p)≢0. Set
(3)
and si = gcd{ki, li}(i = 1,2). For every m ∈ Δ, let
(4)
When a = 1, the reducing subspaces for are described in [15] under seven assumptions as follows:
  • (P1) limp⟶+(ωm+p(k + l)/ωn+p(k + l)) = 1.

  • (P2) If there exists {pj}⊆ such that limj⟶+pj = + and Qn(pj) = 0, then Qn(p) ≡ 0.

  • (P3) If Qn(p) ≡ 0, then Qn+l(p)≢0 and Qn+k(p)≢0.

  • (P4) If Qn(p) ≡ 0, then

  • (P5) Let n ∈ Ω1 and m ∈ Ω4. If Qn(p)≢0 and λn = λm, then Qm(p)≢0.

  • (P6) If nm and Qn(p) ≡ Qm(p), then the following statements hold:

    • (i)

      If Qn+l(p) ≡ Qm+l(p), then Qn+l(p)≢0 and Qn(p)≢0.

    • (ii)

      If Qn+k(p) ≡ Qm+k(p), then Qn+k(p)≢0 and Qn(p)≢0.

  • (P7) Let m, n ∈ Δ and nm. If ωm+k = ωn+k and ωm+h(k + l) = ωn+h(k + l) for h+, then znLm.

Given . For every β ∈ Δ, set
(5)
where qzβ+p(k + l). The equalities and ATφ(zα) = TφA(zα) imply that
(6)
where α ∉ {(p, q) : pk1, qk2} ∪ {(p, q) : pl1, ql2} and c, c ∈  . It seems that the results in [15] can be generalized to the case that as soon as replacing (P4) by the following.

(P4′) If Qn(p) ≡ 0, then

However, (P1) and (P4′) imply that , which contradicts with 0 < a < 1. We find that {p : Qn(p) = 0} is a finite set at most on some classical spaces, such as weighted Dirichlet space. So, this paper is based on the following three assumptions:
  • (A1) limp⟶+(ωm+p(k + l)/ωn+p(k + l)) = 1.

  • (A2) If {p : Qn(p) = 0} is nonempty, then it is a finite set.

  • (A3) Let m,  n ∈ Δ and nm. If ωm+k = ωn+k and ωm+h(k + l) = ωn+h(k + l) for h+, then .

Our main results are as follows: if ω satisfies (A1) and (A2), then the reducing subspace for with a ∈ (0,1) generated by zm is Lm (Theorem 1); if ω satisfies (A1), (A2), and (A3), then Lm is minimal (Theorem 2); for every δ > 0, is a weighted Hardy space satisfying (A1), (A2), and (A3). In addition, we get the structure of reducing subspaces for on (Theorem 3) and the algebraic structure of (Theorem 5).

This paper is arranged as follows: in Section 2, we pay attention to the properties of Qn and give the expression of the element in commutant algebra; in Section 3, we characterize the structure of reducing subspaces for ; in Section 4, we check (A1), (A2), and (A3) on weighted Dirichlet space (δ > 0). This part contains a large number of skilled calculations.

2. The Expression of the Element in Commutant Algebra

Throughout this paper, let k = (k1, k2), l = (l1, l2) ∈ 2,  kl, and . Let where 0 < a < 1. On weighted Hardy space , Toeplitz operator Tφ with nonanalytic symbol φ is defined to be
(7)
where and are multiplication operators and is the adjoint of . Let be the commutant algebra of the von Neumann algebra generated by Tφ. In this section, we mainly consider the properties of under assumptions (A1) and (A2). Clearly, we can write A as follows:
(8)
where cp∖{0},
(9)
(10)

We will prove that the number of elements in Δα,β is no more than 1 under proper assumptions. To show the details, we recall several notes.

Denote by αβ if α1β1 and α2β2. Otherwise, denote by αβ. Clearly, where
(11)
A direct computation gives
(12)
Let
(13)
It is not difficult to check that
(14)
where
(15)
Denote by
(16)

Lemma 1. If Δα,β is nonempty, βk + l, then Δα,β is a finite set.

Proof. Since ωj > 0 for j+, we have λα, λβ. For every , it is easy to check that

(17)

So, Δα,β∖{0}⊆{p : Qβ(p) = λα}. If Δα,β is infinite, there exists such that pj ∈ Δα,β and . Then, Qβ(pj) = λα,  ∀ j. (A1) shows that , which contradicts with (A2).

For s+, if Δα+s(k + l),β is nonempty, βk + l, by the above lemma, we can set

(18)

Lemma 2. Let α, βk + l and s+. Then, Mα,β(s + 1) = Mα,β(s) + 1.

Proof. Let

(19)
where fzβ+t(k + l), t = Mα,β(s), and ct ≠ 0. implies that
(20)
where , , ρ > 0, η ≥ 0, and η = 0 if and only if α + (s − 1)(k + l)⪰0. We prove this lemma by induction.

If s = 0, then η = 0 in equation (20). Since (F + G)⊥zβ+(t + 1)(k + l) and Azαzβ+(t + 1)(k + l), we have

(21)

By equation (20), we get Mα,β(1) = t + 1 = Mα,β(0) + 1.

Suppose the statement holds when s = N − 1(N ≥ 1). If s = N, η > 0 in equation (20). By hypothesis, Mα,β(N − 1) = Mα,β(N) − 1 = t − 1, which implies that Azα+(N − 1)(k + l)zβ+(t + 1)(k + l). Together with (F + G)⊥zβ+(t + 1)(k + l) and Azα+N(k + l)zβ+(t + 1)(k + l), equation (20) shows that

(22)

So, Mα,β(N + 1) = t + 1 = Mα,β(N) + 1.

Lemma 3. Suppose αk + l and . If Δα,β is nonempty, then CardΔα,β = 1. Moreover, if Δα,β = {p0}, then Δα+h(k + l),β = {p0 + h} for any h+.

Proof. Without loss of generality, we may assume βk + l. By Lemma 1, let p0 = Mα,β(0). For every p+, as in Lemma 1, we can prove that {h+ : 〈Azα+h(k + l), zβ+p(k + l)〉 ≠ 0} is a finite set at most. Set

(23)

Then, , and is finite. Let ; we prove the following statements hold by induction:

  • (i)

    For every h+, mα,β(h0 + h + q) > p0 + h, ∀q.

  • (ii)

    For every h+, mα,β(h0 + h + 1) = p0 + h + 1.

Step 1. It is clear that (i) holds for h = 0 from the definition of h0. And we can set

(24)
where , and . By , we have
(25)
where . Since , we have mα,β(h0 + i) ≥ p0 + 1,  i = 1,2. So, equation (25) deduces that
(26)
where . It means that mα,β(h0) ≥ p0. On the contrary, the definition of h0 shows that mα,β(h0) ≤ p0. So,
(27)
and d1 ≠ 0 from equation (26). Hence, equation (24) implies that mα,β(h0 + 1) = p0 + 1. So, (ii) holds for h = 0.

Step 2. Suppose (i) and (ii) hold for hN; then, mα,β(h0 + N + q) > p0 + N, ∀q. So, mα,β(h0 + N + 1 + q) > p0 + N. We can set

(28)
where . By , we get
(29)
where . Since (i) holds for h = N and (F2 + G2) is orthogonal to , we obtain . This means that d2 = 0. So, (i) holds for h = N + 1 from equation (28). In particular, mα,β(h0 + N + 2) ≥ p0 + N + 2. Replacing h0 and p0 by h0 + N + 1 and p0 + N + 1 in equation (24), respectively, we can prove that mα,β(h0 + N + 2) = p0 + N + 2. So, (ii) holds for h = N + 1.

Equation (27) and (ii) show that mα,β(h0 + h) = p0 + h, ∀h+. Together with Lemma 2, we get

(30)

So,

(31)

If h0 ≠ 0, then from (A1), which contradicts with (A2). Therefore, equation (30) implies that mα,β(h) = p0 + h = Mα,β(h), ∀h+. So, we get the desired results.

Lemma 4. Let α ∈ Ω1 and . Then, .

Proof. Suppose there exists some β ∈ Ω2 ∪ Ω3 ∪ Ω4 such that 〈Azα, zβ〉 ≠ 0. Lemma 3 shows that Azα = cβzβ + h(z), where cβ ≠ 0 and . Now, we discuss in two cases.

Case one: β ∈ Ω2 ∪ Ω4. By , we have A(azα+l) = acβzβ+l + cβ(ωβ/ωβk)zβk + H1(z), where . Hence, Δα+l,βk = {p : 〈Azα+l, zβk+p(k + l)〉 ≠ 0} = {0,1}. This is a contradiction to Lemma 3.

Case two: β ∈ Ω3. By ATφ(zα) = TφA(zα), we have A(azα+k) = cβzβ+k + acβ(ωβ/ωβl)zβl + H2(z), where . Hence, Δα+k,βl = {p : 〈Azα+k, zβl+p(k + l)〉 ≠ 0} = {0,1}. This is also a contradiction.

So, ; then, .

3. Reducing Subspaces for Tφ on the Weighted Hardy Space

In this section, we mainly consider the structure of [zm] where m ∈ Δ and [zm] is the reducing subspace for Tφ on generated by zm. Clearly, is a reducing subspace for Tφ with . Let and
(32)
where si = gcd{ki, li}(i = 1,2). Then, and . Let
(33)

The following theorem is similar as Theorem 3.2 in [15]. However, the addition of the coefficient a changes the expressions of Tφ and . For the reader’s convenience, we show the details of the proof.

Theorem 1. Let be a weighted Hardy space satisfying assumptions (A1) and (A2). Then, Lm = [zm] for every m ∈ Δ.

Proof. To show Lm⊆[zm], we only need to prove . Obviously, . Since {cn} and {dn} are strictly monotonically increasing series (see Theorem 3.1 in [15]),

(34)

To show , we will prove the following statements hold for every n by induction.

  • (T1)

  • (T2)

  • (T3)

We mainly use the following two important equalities:

(35)
(36)

Step 1. We prove (T1), (T2), and (T3) hold when n = 1. By calculating, we know

(37)

Thus, and .

If d1 = 0, then (T1) holds. If d1 ≥ 1, equation (35) implies that since and . Similarly, it is easy to check that . Thus, . Together with , we obtain . (T1) holds for n = 1.

Let P be the orthogonal projection from into [zm]. Since , we have

(38)

So,

(39)

By the definition of cn, we know m + c1kl ∈ Ω1 and m + (c1 + 1)k ∈ Ω4. Lemma 4 implies that . Therefore, from equation (39). So, (T2) holds for n = 1. The proof of (T3) is similar to that of (T2). We can prove , which implies (T3) holds for n = 1.

Step 2. Suppose (T1), (T2), and (T3) hold for 1 ≤ np (p), and then we prove that (T1), (T2), and (T3) also hold for n = p + 1. Since ,

(40)
so and .

By equation (36), and imply that

(41)

Also, for 1 ≤ jdp + p. Therefore, .

By equation (35), and imply that

(42)

Using the same technique, we finally get . By hypothesis to (T1), we obtain . So, (T1) holds for n = p + 1.

By (T1), we have . Notice that m + cp+1k − (p + 1)l ∈ Ω1 and m + (cp+1 + 1)kpl ∈ Ω4. Put u = cp+1 and v = −p in equation (35). As in Step 1, it is easy to prove that

(43)

That is, (T2) holds for n = p + 1. Similarly, by and equation (36), we can prove that (T3) holds for n = p + 1. So, we finish the proof.

Lemma 5. Let m, n ∈ Ω1 and nm. If 〈Azm, zn〉 ≠ 0, then ωm+k = ωn+k and ωm+h(k + l) = ωn+h(k + l) for every h+.

Proof. Since 〈Azm, zn〉 ≠ 0, Lemma 3 shows that Δm+p(k + l),n = {p},  ∀ p+. Set

(44)
where complex number ap ≠ 0 and . We claim that
(45)

For every N, deduces that

(46)
for some ρ, η > 0. Since A(zm+N(k + l)), A(zm+(N − 1)(k + l)), and are all orthogonal to zn+(N + 1)(k + l), we have
(47)
for some . Comparing with equation (44) for n = N + 1, there is aN = aN+1. So, it is easy to get the claim by mathematical induction.

Furthermore, equation (46) implies that

(48)

Let p⟶+; then,

(49)

That is, ωm = ωn and ωm+p(k + l) = ωn+p(k + l) for every p.

When p = 0, Azm = a0zn + f0(z). By , we have

(50)

By the expression of Azm+k+l and Azm, we get (ωm+k/ωm) = (ωn+k/ωn). Therefore, ωm+k = ωn+k.

Theorem 2. Let be a weighted Hardy space satisfying assumptions (A1), (A2), and (A3). Then, for every m ∈ Δ, Lm = [zm] is a minimal reducing subspace for Tφ.

Proof. Assume a nonzero subspace ℳ⊆Lm is a reducing subspace for Tφ. Let P be the orthogonal projection from into ℳ; then, . By Lemma 4, set

(51)

Lemma 5 and (A3) show that if nm, then znLm. So, Pzm = amzm for some am. If am = 0, then Lm⊥ℳ, which contradicts with ℳ ≠ {0}. Hence, am ≠ 0 and Lm = [zm]⊆ℳ.

4. Results on the Weighted Dirichlet Space

The weighted Dirichlet space is defined as follows:
(52)
So, we can get a more specific formula about Qn(p):
(53)

It is clear that the weight ω of satisfies (A1). The following lemmas show that ω satisfies (A2) and (A3). Then, the results in Section 3 hold.

Lemma 6. For , {p : Qn(p) = 0} is a finite set at most.

Proof. Suppose {p : Qn(p) = 0} is an infinite set; then, there exists such that Qn(pj) = 0 and . By a direct computation, we get Qn(p) = 0 if and only if

(54)

Let

(55)
where m = (m1, m2) and mi ≥ −max{ki, li} with i = 1,2. Then, there exists r > 0 such that vmC[−r, r]. Denote by
(56)

Thus, for tj = (1/pj),

(57)

Set

(58)
(59)

A careful computation gives

(60)
(61)

It follows that . By L’Hospital’s rule, we have

(62)

That is,

(63)
(64)

Moreover, equation (60) shows that . By L’Hospital’s rule again, we get

(65)

That is,

(66)

Similarly,

(67)

Together with equations (61), (64), and (66), we get . That is, y1 + y2 = x1 + x2. Notice that x1 + y1 = x2 + y2; we obtain x1 = y2 and x2 = y1. Combining with equation (63), we have a = 1, which contradicts with a ∈ (0,1).

Lemma 7. Let m,  n ∈ Δ and nm. If ωm+k = ωn+k and ωm+h(k + l) = ωn+h(k + l) for every h+, then

(68)

Proof. By the definition of ωn and the equalities ωm = ωn and ωm+k = ωn+k, we have

(69)

Therefore, k1(m2n2) + k2(m1n1) = 0. Thus, m1 = n1 if and only if m2 = n2. Similarly, by ωm+k = ωn+k and ωm+k+l = ωn+k+l, we can prove that l1(m2n2) + l2(m1n1) = 0. Since nm, we get n1m1, n2m2, and k1l2 = k2l1. Furthermore, by k1((m2 + 1) − (n2 + 1)) = k2((n1 + 1) − (m1 + 1)) and equation (69), we get the desired results.

The above lemma shows that ω satisfies (A3). In fact, if znLm, then there exist u,  v such that n1 = (l1/l2)(m2 + 1) − 1 = m1 + uk1 + vl1 and n2 = (l2/l1)(m1 + 1) − 1 = m2 + uk2 + vl2. It is easy to see that uk1 + vl1 = uk2 + vl2 = 0. Then, n1 = m1 and n2 = m2, which contradict with nm.

Since ω satisfies (A1), (A2), and (A3), Theorem 2 implies the following result.

Corollary 1. Let k,  l2,  kl. Then, for every m ∈ Δ, Lm = [zm] is a minimal reducing subspace for Tφ on .

Combining Lemmas 4, 5, and 7, we characterize Azm for and m ∈ Δ.

Corollary 2. Let and m ∈ Δ. Then,

  • (i)

    If k1l2k2l1, then Azm = czm, where c.

  • (ii)

    If k1l2 = k2l1, then , where m = ((l1/l2)(m2 + 1) − 1, (l2/l1)(m1 + 1) − 1), a, b. In particular, if , then b = 0.

Recall that two reducing subspaces ℳ1 and ℳ2 for Tφ are called unitarily equivalent if there exists a linear operator U on such that is unitary from ℳ1 onto ℳ2, , and U commutes with both Tφ and . Together with Corollary 2, we get the following theorem, which gives a complete description of the reducing subspaces for Tφ.

Theorem 3. Let k, l2,  kl. If ℳ is a reducing subspace for Tφ on , then ℳ is the orthogonal sum of some minimal reducing subspaces. Moreover, ℳ is a minimal reducing subspace for Tφ if and only if ℳ has the form as follows:

  • (i)

    If k1l2l1k2, then ℳ = Lm for some m ∈ Δ.

  • (ii)

    If k1l2 = l1k2, then there exist m ∈ Δ and s, t such that ℳ = ℳst where ℳst are defined by

    (70)

  • with m = ((l1/l2)(m2 + 1) − 1, (l2/l1)(m1 + 1) − 1). In particular, if , then t = 0.

Proof. For n ∈ Δ, if zn⊥ℳ, Corollary 1 shows that [zn] = Ln⊥ℳ. There is m ∈ Δ such that Pzm ≠ 0. If k1l2l1k2, Corollary 2 (i) implies that Pzm = czm with c ≠ 0. And then, Lm⊆ℳ. If k1l2 = l1k2, Corollary 2 (ii) implies that there are s, t such that , where m = ((l1/l2)(m2 + 1) − 1, (l2/l1)(m1 + 1) − 1). If , then t = 0. If , we claim that . Therefore, ℳst⊆ℳ. In fact, we only need to prove there is a linear operator U on such that is unitary from Lm onto ℳs,t, UTφ = TφU, and . That is, Lm and ℳst are unitarily equivalent.

Let and

(71)
for any i, j such that . By k1l2 = k2l1, a computation gives . Thus, ‖U(zm+ik+jl)‖ = ‖zm+ik+jl‖. It remains to prove UTφ = TφU. Referring to Theorem 3.7 of [6], we can get
(72)
where m is defined as in (ii). Then, ml implies that ml. In this case,
(73)

Similarly, ml implies that ml; then,

(74)

Since and , we obtain UTφ = TφU. The rest of the proof is obvious.

Let m, m be as in the above theorem, and let and be a linear operator defined by

(75)

It is not difficult to get the following theorem.

Theorem 4. Let k, l2,  kl. Suppose m, m ∈ Δ; then, the following statements hold:

  • (i)

    If k1l2k2l1, then Lm and are unitarily equivalent if and only if m = m.

  • (ii)

    If k1l2 = k2l1, then Lm and are unitarily equivalent if and only if m = m or m = ((l1/l2)(m2 + 1) − 1, (l2/l1)(m1 + 1) − 1). In particular, if , then Lm and are unitarily equivalent if and only if m = m.

By the above theorem and Corollary 8.2.6 in [4], we can characterize . Since the proof of the following theorem is the same as that in [6, 15], we omit the details.

Theorem 5. Let k, l2,  kl. Then, is a type I von Neumann algebra. Furthermore, the following statements hold:

  • (i)

    If k1l2k2l1, then is abelian and is ∗−isomorphic to , where j = |k1l2k2l1|.

  • (ii)

    If k1l2 = k2l1 and s = (s1, s2) with si = gcd{ki, li} (i = 1,2), then and is never abelian. Moreover, if s1 = s2 = r, then is ∗−isomorphic to

    (76)

If s1s2, then is ∗−isomorphic to the direct sum of countably many M2() ⊕ .

Conflicts of Interest

The authors declare that they have no conflicts of interest.

Acknowledgments

This study was supported by Fundamental Research Funds for the Central Universities (Grant no. 201964007), the Shandong Provincial Natural Science Foundation (Grant no. ZR2020MA009), and the National Natural Science Foundation of China (Grant no. 11701537).

    Data Availability

    No data were used to support this study.

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