1 Introduction

Let \(H^2\) denote the standard Hardy space on the unit disk \(\mathbb {D}\). For \(\varphi \in L^{\infty }(\partial \mathbb {D})\) the Toeplitz operator on \(H^2\) is given by \(T_{\varphi }f=P_{+}(\varphi f)\), where \(P_{+}\) is the orthogonal projection of \(L^2(\partial \mathbb {D})\) onto \(H^2\). We will denote by \(\mathcal {M}(\varphi )\) the range of \(T_{\varphi }\) equipped with the range norm, that is, the norm that makes the operator \(T_{\varphi }\) a coisometry of \(H^2\) onto \(\mathcal {M}(\varphi )\). For a nonconstant function b in the unit ball of \(H^\infty \) the de Branges–Rovnyak space \(\mathcal {H}(b)\) is the image of \(H^2\) under the operator \((1-T_bT_{\bar{b}})^{1/2}\) with the corresponding range norm. The norm and the inner product in \(\mathcal {H}(b)\) will be denoted by \(\Vert \cdot \Vert _b\) and \(\langle \cdot ,\cdot \rangle _b\). The space \(\mathcal {H}(b)\) is a Hilbert space with the reproducing kernel

$$\begin{aligned} k_w^b(z)=\frac{1-\overline{b(w)}b(z)}{1-{\overline{w}}z}\qquad (z,w\in {\mathbb {D}}). \end{aligned}$$

In the case when b is an inner function the space \(\mathcal {H}(b)\) is the well-known model space \(K_b=H^2\ominus bH^2\).

If the function b fails to be an extreme point of the unit ball in \(H^{\infty }\), that is, when \(\log (1-|b|)\in L^1(\partial \mathbb D)\), we will say simply that b is nonextreme. In this case one can define an outer function a whose modulus on \(\partial \mathbb {D}\) equals \(\left( 1-|b|^2\right) ^{1/2}\). Then we say that the functions b and a form a pair (ba). By the Herglotz representation theorem there exists a positive measure \(\mu \) on \(\partial \mathbb {D}\) such that

$$\begin{aligned} \frac{1+b(z)}{1-b(z)}=\int _{\partial \mathbb {D}}\frac{1+ e^{-i\theta }z}{1- e^{-i\theta }z}\,d\mu (e^{i\theta })+i{{\,\mathrm{Im}\,}}\frac{1+b(0)}{1-b(0)},\quad z\in \mathbb {D}. \end{aligned}$$
(1)

Moreover the function \(\left| \frac{a}{1-b}\right| ^2\) is the Radon-Nikodym derivative of the absolutely continuous component of \(\mu \) with respect to the normalized Lebesgue measure. If the measure \(\mu \) is absolutely continuous the pair (ba) is called special.

Recall that a function \(f\in H^1\) is called rigid if and only if no other functions in \(H^1\), except for positive scalar multiples of f have the same argument as f a.e. on \(\partial \mathbb {D}\).

If (ba) is a pair, then \(\mathcal {M}(a)\) is contained contractively in \(\mathcal {H}(b)\). If a pair (ba) is special and \(f=\frac{a}{1-b}\), then \(\mathcal {M}(a)\) is dense in \(\mathcal {H}(b)\) if and only if \(f^2\) is rigid ( [20]). Spaces \(\mathcal {H}(b)\) for nonextreme b have been studied in [2, 3, 5, 15, 16, 22], and [23].

The kernels of Toeplitz operators have been studied since the late 80’s. We mention that two recent survey articles [1, 8] and the book [4] contain a number of results on this topic.

The Hayashi theorem [12] (see also [21]) states that the kernel of a Toeplitz operator \(T_{\varphi } \) is a subspace of \(H^2\) of the form \(\ker T_{\varphi }= fK_I\), where \(K_I= H^2\ominus IH^2\) is the model space corresponding to the inner function I such that \(I(0)=0\) and f is an outer function of unit \(H^2\) norm that acts as an isometric multiplier from \(K_I\) onto \(f K_{I}\). Moreover, f can be expressed as \(f=\frac{a}{1-Ib_0}\), where \((b_0,a)\) is a special pair and \(\bigl (\frac{a}{1-b_0}\bigr )^2 \) is a rigid function in \(H^1\). Then we also have \(\ker T_{\frac{\bar{I}\bar{f}}{f}}=fK_{I}\). In the recent paper [6] the authors considered the Toeplitz operator \(T_{\frac{\bar{g}}{g}}\) where \(g\in H^{\infty }\) is outer. Among other results, they described all outer functions g such that \( \ker T_{\frac{\bar{g}}{g}}=K_{I}\). In Sect. 2 we describe all such functions g for which \(\ker T_{\frac{\bar{g}}{g}}=f K_{I}\).

If (ba) is a special pair, \(f=\frac{a}{1-b}\) and \(b=Ib_0\), where I as above, then \(f K_{I}\subset \ker T_{\frac{\bar{I}\bar{f}}{f}}\). In the next two sections we study the space \(\ker T_{\frac{\bar{I}\bar{f}}{f}}\ominus f K_{I}\) and show that it is isometrically isomorphic to the orthogonal complement of \(\mathcal {M}(a)\) in the de Branges–Rovnyak space \(\mathcal {H}(b_0)\). We also give an example of a function f for which the space \(\ker T_{\frac{\bar{I}\bar{f}}{f}}\ominus f K_{I}\) is one dimensional. In the last section we discuss the orthogonal complement of \(\mathcal {M}(a)\) in \({\mathcal {H}}(b)\) and get a generalization of results obtained in [15] and [5] for the case when pairs are rational.

2 The kernel of \(T_{\frac{{\bar{g}}}{g}}\)

It is known that if g is an outer function in \(H^2\), then the kernel of \(T_{\frac{\bar{g}}{g}}\) is trivial if and only if \(g^2\) is rigid (see e.g. [18]).

The finite dimensional kernels of Toeplitz operators were described by Nakazi [17]. Nakazi’s theorem says that \(\dim \ker T_{\varphi }=n\) if and only if there exists an outer function \(f\in H^2\) such that \(f^2\) is rigid and \(\ker T_{\varphi }=\{fp:\, p\in \mathcal {P}_{n-1}\},\) where \(\mathcal {P}_{n-1}\) denotes the set of all polynomials of degree at most \(n-1\).

Consider the following example.

Example

For \(\alpha >-\frac{1}{2}\) set \(g(z)=(1-z)^\alpha \), \(z\in \mathbb {D}\). Then the kernel of \(T_{\frac{\bar{g}}{g}}\) is trivial for \(\alpha \in (-\frac{1}{2},\frac{1}{2}]\) and dimension of the kernel of \(T_{\frac{{\overline{g}}}{g}}\) is n for \(\alpha \in (n-\frac{1}{2},n+\frac{1}{2}], n=1,2,\ldots \), and

$$\begin{aligned} \ker T_{\frac{\overline{(1-z)^\alpha }}{(1-z)^\alpha }}=(1-z)^{\alpha -n}K_{z^n}. \end{aligned}$$

In the general case the kernels of Toeplitz operators are characterized by Hayashi’s theorem. To state this theorem we need some notation. We note that an outer function f having unit norm in \(H^2\) \((\Vert f\Vert _2=1)\) can be written as

$$\begin{aligned} f=\frac{a}{1-b}, \end{aligned}$$

where a is an outer function, b is a function from the unit ball of \(H^{\infty }\) such that \(|a|^2+|b|^2=1\) a.e. on \(\partial \mathbb {D}\). Following Sarason [p. 156] Sarason3 we call (ba) the pair associated with f. Note also that b is a nonextreme point of the closed unit ball of \(H^{\infty }\) and is given by

$$\begin{aligned} \frac{1+b(z)}{1-b(z)}=\frac{1}{2\pi }\int _{0}^{2\pi }\frac{1+ e^{-i\theta }z}{1- e^{-i\theta }z}|f(e^{i\theta })|^2d\theta ,\quad z\in \mathbb {D}. \end{aligned}$$
(2)

Let S denote the unilateral shift operator on \(H^2\), i.e., \(S=T_z\). A closed subspace M of \(H^2\) is said to be nearly \(S^*\)-invariant if for every \(f\in M\) vanishing at 0, we also have \(S^*f\in M\). It is known that the kernels of Toeplitz operators are nearly \(S^*\)-invariant.

Nearly \(S^*\)-invariant spaces are characterized by Hitt’s theorem [14].

Hitt’s Theorem

The closed subspace M of \(H^2\) is nearly \(S^*\)-invariant if and only if there exists a function f of unit norm and a model space \(K_I=H^2\ominus IH^2\) such that \(M= T_fK_I\), where I is an inner function vanishing at the origin, and \(T_f\) acts isometrically on \(K_I\).

It has been proved by D. Sarason [18] that \(T_f\) acts isometrically on \(K_I\) if and only if I divides b (the first function in the pair associated with f). Consequently, the function f in Hitt’s theorem can be written as

$$\begin{aligned} f=\frac{a}{1-Ib_0}. \end{aligned}$$

The function \( \frac{1+b_0(z)}{1-b_0(z)}\) has a positive real part and is the Herglotz integral of a positive measure on \(\partial \mathbb {D}\) up to an additive imaginary constant,

$$\begin{aligned} \frac{1+b_0(z)}{1-b_0(z)}=\int _{\partial \mathbb {D}}\frac{1+ e^{-i\theta }z}{1- e^{-i\theta }z}\,d\mu (e^{i\theta })+ic. \end{aligned}$$
(3)

Clearly \(b_0\) is also a nonextreme point of the closed unit ball of \(H^{\infty }\) and \(|a|^2+|b_0|^2=1\) a.e. on \(\partial \mathbb {D}\). We remark that in view of (2) the pair (ba) associated with an outer function \(f\in H^2\) is special, while the pair \((b_0,a)\) need not to be special. We also put

$$\begin{aligned} f_0=\frac{a}{1-b_0} \end{aligned}$$

and note that \(f_0\in H^2\) (see, e.g. [4, Theorem 23.1]).

Under the above notations Hayashi’s theorem reads as follows:

Hayashi’s Theorem

The nearly \(S^*\)-invariant space \(M=T_fK_I\) is the kernel of a Toeplitz operator if and only if the pair \((b_0,a)\) is special and \(f_0^2\) is a rigid function.

Moreover, it follows from Sarason’s proof of Hayashi’s theorem that if \(M=T_fK_I\) is the kernel of a Toeplitz operator then it is the kernel of \(T_{\frac{\bar{I}{\bar{f}}}{f}}\).

Recently E. Fricain, A. Hartmann and W. T. Ross [6] considered the Toeplitz operators \(T_{\frac{\bar{g}}{g}} \) where \(g\in H^{\infty }\) is outer. If \(\ker T_{\frac{\bar{g}}{g}}\) is non-trivial, then by Hayashi’s theorem there exist the outer function f and the inner function I, \(I(0)=0\), such that

$$\begin{aligned} \ker T_{\frac{\bar{g}}{g}}=fK_{I}. \end{aligned}$$

In the above mentioned paper [6] the authors described all outer functions \(g\in H^{\infty }\) for which

$$\begin{aligned} \ker T_{\frac{\bar{g}}{g}}=K_I, \end{aligned}$$

where I is an inner function not necessarily satisfying \(I(0)=0\).

We prove the following

Theorem 1

Assume that \(g\in H^2\) is outer and \(M= T_fK_{I}\) is the nearly \(S^*\)-invariant space, where I is an inner function such that \(I(0)=0\), \((b_0I,a)\) is the pair associated with the outer function f, \((b_0,a)\) is special, and \(f_0^2\) is rigid. Then \(\ker T_{\frac{{\bar{g}}}{g}}=M\) if and only if

$$\begin{aligned} g= i\frac{I_1+I_2}{I_1-I_2}(1+ I)f, \end{aligned}$$

where \(I_1\) and \(I_2\) are inner and \(I_1-I_2\) is outer.

Recall that the Smirnov class \({\mathcal {N}}^{+}\) consists of those holomorphic functions in \(\mathbb {D}\) that are quotients of functions in \(H^{\infty }\) in which the denominators are outer functions.

In the proof of Theorem 1, similarly to [6], we use the following result due to H. Helson [13].

Helson’s Theorem

The functions \(f\in {\mathcal {N}}^{+}\) that are real almost everywhere on \(\partial \mathbb {D}\) can be written as

$$\begin{aligned} f=i\,\frac{I_1+I_2}{I_1-I_2}, \end{aligned}$$

where \(I_1\) and \(I_2\) are inner and \(I_1-I_2\) is outer.

We also apply a description of kernels in terms of \(S^*\)-invariant subspaces \(K_I^p(|f|^p)\) of weighted Hardy spaces (in the case when \(p=2\)) considered by A. Hartmann and K. Seip in their paper [10] (see also [7]). For an outer function f in \(H^2\) the weighted Hardy space is defined as follows

$$\begin{aligned} H^2(|f|^2)=\{g\in {\mathcal {N}}^{+}:\, \Vert g\Vert ^2_{2,f}=\frac{1}{2\pi }\int _0^{2\pi }|g(e^{it})|^2|f(e^{it})|^2dt<\infty \} \end{aligned}$$

and, for an inner function I, \(K^2_I(|f|^2)= K_{I}(|f|^2)\) is given by

$$\begin{aligned} K_{I}(|f|^2)=\{g=I{\overline{\psi }}\in H^2(|f|^2):\, \psi \in H_0^2(|f|^2)\}, \end{aligned}$$

where \(H_0^2(|f|^2)=zH^2(|f|^2)\).

Then \(K_{I}(|f|^2)\) is \(S^*\)-invariant and \(fK_{I}(|f|^2)=\ker T_{\frac{{\bar{I}} {\bar{f}}}{f}}\) (see [10]).

Proof of Theorem 1

Assume that \(\ker T_{\frac{{\bar{g}}}{g}}=fK_{I}\). Then

$$\begin{aligned} fK_I=\ker T_{\frac{{\bar{I}} {\bar{f}}}{f}}=\ker T_{\frac{{\bar{g}}}{g}}. \end{aligned}$$

Since \(f\in \ker T_{\frac{{\bar{I}} {\bar{f}}}{f}}\), the last equalities imply that

$$\begin{aligned} \frac{{\overline{g}}f}{g}= {\overline{I}}_0{\bar{h}}, \end{aligned}$$

where \( I_0\) is an inner function such that \(I_0(0)=0\), and \(h\in H^2\) is outer. This means that \(|f(z)|=|h(z)|\) a.e. on \(|z|=1\) and consequently \(h(z)=cf(z)\), where c is a unimodular constant. Replacing \(cI_0\) by \(I_0\), we get

$$\begin{aligned} \frac{\overline{g}}{g}={\overline{I}}_0\frac{{\overline{f}}}{f}. \end{aligned}$$
(4)

It then follows

$$\begin{aligned} fK_{I}=\ker T_{\frac{{\bar{g}}}{g}}= \ker T_{\frac{ {\overline{I}}_0{\overline{f}}}{f}}= fK_{I_0}(|f|^2), \end{aligned}$$

which implies \(I=I_0\) up to a unimodular constant. Indeed, these equalities imply that an analytic function h can be written in the form \(h = f I_0{\overline{\psi }}_0\), where \(\psi _0\in H^2_0(|f|^2) \), if and only if \(h= f I{\overline{\psi }}\), where \(\psi \in H^2_0\). Since \( |\psi _0|= |\psi |\) a.e. on \(|z|=1\) and \(\psi _0\in {\mathcal {N}}^{+}\), we see that also \(\psi _0\in H_0^2\). Hence \(K_{I}=K_{I_0}\).

Consequently, equality (4) can be written as

$$\begin{aligned} \frac{{\bar{g}}}{g}=\frac{\overline{f(1+ I)}}{f(1+ I)} \quad \text {a.e. on } \partial \mathbb {D}, \end{aligned}$$

which means that the function \(\frac{g}{f(1+ I)}\) is real a.e. on \(\partial \mathbb {D}\). Since this function is in the Smirnov class \({\mathcal {N}}^{+}\), our claim follows from Helson’s theorem. To prove the other implication it is enough to observe that if

$$\begin{aligned} g=i\,\frac{I_1+I_2}{I_1-I_2}(1+ I)f, \end{aligned}$$

then

$$\begin{aligned} \frac{{\bar{g}}}{g}=\frac{{\overline{I}}{{\bar{f}}}}{f}. \end{aligned}$$

\(\square \)

3 The complement of \(fK_I\) in \(\ker T_{\frac{\bar{I}\bar{f}}{f}}\)

It was noticed in [4, Corollary 30.21] that if f is an outer function of the unit norm, (ba) is the pair associated with f, and I is an inner function vanishing at the origin that divides b, then

$$\begin{aligned} fK_{I}\subset \ker T_{\frac{\bar{I}{\bar{f}}}{f}} \end{aligned}$$

and, according to Hayashi’s theorem, the equality holds if and only if the pair \((b_0,a)\) is special and \(f_0^2\) is rigid.

Recall that \(\mathcal {M}(a)\) is dense in \(\mathcal {H}(b_0)\) if and only if the pair \((b_0,a)\) is special and \(f_0^2\) is a rigid function.

Theorem 2

Assume that \((Ib_0, a)\), where I is inner, and \(I(0)=0\), is the pair associated with an outer function f. If the pair \((b_0,a)\) is not special or the function \(f_0^2\) is not rigid, then for a positive integer k,

$$\begin{aligned} \dim \bigl (\ker T_{\frac{{\bar{I}} {\bar{f}}}{f}}\ominus fK_{I}\bigr )=k \end{aligned}$$

if and only if the codimension of \(\overline{\mathcal {M}(a)}\) in the de Branges–Rovnyak space \(\mathcal {H}(b_0)\) is k.

In the proof of this theorem we use some ideas from Sarason’s proof of Hayashi’s theorem. If a positive measure \(\mu \) on the unit circle \(\partial \mathbb {D}\) is as in (1) and \(H^2(\mu )\) is the closure of the polynomials in \(L^2(\mu )\), then an operator \(V_b\) given by

$$\begin{aligned} (V_bq)(z)=(1-b(z))\int _{\partial \mathbb {D}}\frac{q(e^{i\theta })}{1-e^{-i\theta }z}\,d\mu (e^{i\theta }) \end{aligned}$$
(5)

is an isometry of \(H^2(\mu )\) onto \(\mathcal {H}(b)\). Furthermore, if (ba) is a pair and \(f=\frac{a}{1-b}\), then the operator \(T_{1-b}T_{\bar{f}}\) is an isometry of \(H^2\) into \(\mathcal {H}(b)\). Its range is all of \(\mathcal {H}(b)\) if and only if the pair (ba) is special ( [21, III-6,7] and [4, Theorem 24.26]). We note that in the last case \(d\mu (e^{i\theta })=\frac{1}{2\pi }|f(e^{i\theta })|^2d\theta \).

Proof of Theorem 2

Since the pair (ba) is special, the operator \(T_{1-b}T_{{\bar{f}}}\) is an isometry of \(H^2\) onto \(\mathcal {H}(b)\). Moreover, since I divides b, \(T_{f}\) acts as an isometry on \(K_{I}\) and \(T_{1-b}T_{\bar{f}}(fK_I)=K_I\) ( [20]). Hence

$$\begin{aligned} \mathcal {H}(b)&=T_{1-b}T_{{\bar{f}}}(H^2)= T_{1-b}T_{\bar{f}}\bigl (\overline{T_{\frac{If}{{\bar{f}}}}(H^2)}\oplus \big (T_{\frac{If}{{\bar{f}}}}(H^2)\bigr )^\bot \bigr )\\&= \overline{I\mathcal {M}(a)}^b\oplus T_{1-b}T_{{\bar{f}}}(\ker T_{\frac{\bar{I}{\bar{f}}}{f}})\\ {}&= \overline{I\mathcal {M}(a)}^b\oplus T_{1-b}T_{{\bar{f}}} (fK_I) \oplus T_{1-b}T_{{\bar{f}}}(\ker T_{\frac{{\bar{I}}{\bar{f}}}{f}}\ominus fK_I)\\&= \overline{I\mathcal {M}(a)}^b\oplus K_I\oplus T_{1-b}T_{\bar{f}}(\ker T_{\frac{{\bar{I}}{\bar{f}}}{f}} \ominus fK_I), \end{aligned}$$

where \(\overline{T_{\frac{If}{{\bar{f}}}}(H^2)}\) denotes the closure of \(T_{\frac{If}{{\bar{f}}}}(H^2)\) in \(H^2\) and \(\overline{I\mathcal {M}(a)}^b\) denotes the closure of \(I\mathcal {M}(a)\) in \(\mathcal {H}(b)\). On the other hand,

$$\begin{aligned} \mathcal {H}(b) = \mathcal {H}(b_0I) =K_I\oplus I\mathcal {H}(b_0)= K_I\oplus I(\mathcal {H}(b_0)\ominus \overline{\mathcal {M}(a)}^{b_0})\oplus I\overline{\mathcal {M}(a)}^{b_0}. \end{aligned}$$

Since \(T_{I}:\, \mathcal {H}(b_0)\rightarrow \mathcal {H}(Ib_0)\) is an isometry ( [20, Proposition 4]), \( I\overline{(\mathcal {M}(a))}^{b_0}= \overline{I \mathcal {M}(a)}^{b}\). It then follows,

$$\begin{aligned} T_{1-b}T_{{\bar{f}}}(\ker T_{\frac{{\bar{I}}{\bar{f}}}{f}}\ominus fK_I)= I(\mathcal {H}(b_0)\ominus \overline{\mathcal {M}(a)}^{b_0}). \end{aligned}$$
(6)

\(\square \)

We remark that the orthogonal complement of \(\mathcal {M}(a)\) in \(\mathcal {H}(b)\) is discussed in Sect. 5.

4 The example

Let, as in the previous sections, f be an outer function in \(H^2\) and let (ba) be the pair associated with f. Let \(b=Ib_0\), where I is an inner function such that \(I(0)=0\) and \(f_0=\frac{a}{1-b_0}\). Then \(fK_I\subset \ker T_{\frac{\bar{I}\bar{f}}{f}}\) and equality holds if and only if the pair \((b_0,a)\) is special and \(f_0^2\) is rigid. Moreover, if the pair \((b_0,a)\) is special and \( f_0^2 \) is rigid, then (ba) is special and \( f^2\) is rigid but the converse implication fails ( [18, p. 158]).

In [4, vol. 2, pp. 541–542] the authors constructed a function h in \(\ker T_{\frac{\bar{I}{\bar{f}}}{f}}\) which is not in \(fK_{I}\) under the assumption that \(f^2\) is not rigid. Here we consider the function f such that \(f^2\) is rigid, the pair \((b_0,a)\) is special but \(f^2_0\) is not rigid, and describe the space \(\ker T_{\frac{{\bar{I}}{\bar{f}}}{f}} \ominus fK_{I} \).

Our example is a slight modification of the one given in see also [4, vol. 2, p. 494]. The corresponding functions f and \(f_0\) are defined by taking \(a(z)=\frac{1}{2}( 1+z)\), \(b_0(z)=\frac{1}{2}z(1-z)\), and \(I(z)=zB(z)\), where B(z) is a Blaschke product with zero sequence \(\{r_n\}_{n=1}^{\infty }\) lying in \((-1,0)\) and converging to \(-1\). It has been proved in [19, pp. 491–492] (see also [4, vol. 2, pp. 494–496] that \(f^2\) is rigid while \(f^2_0\) is not. Notice that the pair \((b_0,a)\) is rational and the point \(-1\) is the only zero of the function a. It then follows from [15, Theorem 4.1] (see also [5]) that \(\mathcal {M}(a)\) is a closed subspace of \(\mathcal {H}(b_0)\) and

$$\begin{aligned} \mathcal {H}(b_0)= \mathcal {M}(a)\oplus {\mathbb {C}}k_{-1}^{b_0}, \end{aligned}$$

where

$$\begin{aligned} k_{-1}^{b_0}(z)=\frac{1-\overline{b_0(-1)}b_0(z)}{1+z}=\frac{2-z}{2}. \end{aligned}$$

Thus we see that

$$\begin{aligned} \mathcal {H}(b_0)\ominus \mathcal {M}(a)= \mathbb {C}k_{-1}^{b_0}. \end{aligned}$$

Moreover (6) implies that

$$\begin{aligned} T_{1-b}T_{{\bar{f}}}(\ker T_{\frac{{\bar{I}}{\bar{f}}}{f}}\ominus fK_I)= \mathbb {C}I k_{-1}^{b_0}. \end{aligned}$$

Our aim is to prove that

$$\begin{aligned} \ker T_{\frac{{\bar{I}}{\bar{f}}}{f}}\ominus fK_I=\mathbb {C}g, \end{aligned}$$
(7)

where the function \(g\in H^2\) is given by \(g=fk_{-1}(I+1),\) with \(k_{-1}(z)=(1+z)^{-1}\), \(z\in \mathbb {D}\).

For \(\lambda \) in \(\mathbb {D}\) let \(k_\lambda \) denote the kernel function in \(H^2\) for the functional of evaluation at \(\lambda \), \(k_\lambda (z)=(1-{\bar{\lambda }}z)^{-1}\). In the proof of (7) we will apply the following

Lemma

[9, Lemma 2]

  1. (i)

    \(P_{+}\left( |f|^2Ik_\lambda \right) =\dfrac{Ik_\lambda }{1-b}+\dfrac{\overline{b_0(\lambda )}k_\lambda }{1-\overline{b(\lambda )}}\).

  2. (ii)

    \(P_{+}\left( |f|^2k_\lambda \right) =\dfrac{k_\lambda }{1-b}+\dfrac{\overline{b(\lambda )}k_\lambda }{1-\overline{b(\lambda )}}\).

Since \(I(r_n)=0\), (i) and (ii) in the Lemma yield

$$\begin{aligned} T_{1-b}T_{{\bar{f}}}(f I k_{r_n})&= Ik_{r_n}(1-\overline{b_0(r_n)}b_0)+ \overline{b_0(r_n)}k_{r_n}, \\ T_{1-b}T_{{\bar{f}}}(f k_{r_n})&=k_{r_n}. \end{aligned}$$

Hence

$$\begin{aligned} T_{1-b}T_{{\bar{f}}}(f k_{r_n}(I-\overline{b_0(r_n)}))=Ik_{r_n}(1-\overline{b_0(r_n)}b_0)=Ik_{r_n}^{b_0}. \end{aligned}$$
(8)

It follows from [4, Theorem 21.1] that

$$\begin{aligned} \Vert k_{r_n}^{b_0}- k_{-1}^{b_0}\Vert _{b_0}\underset{n\rightarrow \infty }{\longrightarrow }0. \end{aligned}$$

Next, since \(T_{I}:\, \mathcal {H}(b_0)\rightarrow \mathcal {H}(Ib_0)=\mathcal {H}(b)\) is an isometry and \(T_{1-b}T_{{\bar{f}}}\) is an isometry of \(H^2\) onto \(\mathcal {H}(b)\), we see that \( \{fk_{r_n}(I-\overline{b_0(r_n)}) \}_{n\in \mathbb {N}}\) is a bounded sequence in \(H^2\). So it contains a subsequence that converges weakly, say, to a function \(g\in H^2\). Without loss of generality, we may assume that the sequence \(\{fk_{r_n}(I-\overline{b_0(r_n)})\}\) itself converges weakly to g. Then for any point \(z\in \mathbb {D}\),

$$\begin{aligned} g(z)&= \langle g, k_z\rangle = \lim \limits _{n\rightarrow \infty }\langle fk_{r_n}(I-\overline{b_0(r_n)}), k_z\rangle \\&=\lim _{n\rightarrow \infty }\frac{f(z)(I(z)-\overline{b_0(r_n)})}{1-r_nz}= \frac{f(z)(I(z)+1)}{1+z}. \end{aligned}$$

Now observe that since

$$\begin{aligned} \Vert fk_{r_n}(I-\overline{b_0(r_n)})\Vert _{2}=\Vert k_{r_n}^{b_0}\Vert _{b_0}\quad \text {and}\quad \Vert fk_{-1}(I+1)\Vert _{2}=\Vert k_{-1}^{b_0}\Vert _{b_0}, \end{aligned}$$

\(fk_{r_n}(I-\overline{b_0(r_n)})\rightarrow fk_{-1}(I+1)\) in \(H^2\) strongly. Finally, passing to the limit in (8) gives

$$\begin{aligned} T_{1-b}T_{{\bar{f}}}(fk_{-1}(I+1))=Ik_{-1}(1+b_0)=Ik_{-1}^{b_0}, \end{aligned}$$

which proves (7).

Remark

One can check directly that the function \(g=fk_{-1}(I+1)\) is in \(\ker T_{\frac{{\bar{I}}{\bar{f}}}{f}}\ominus fK_I\).

Indeed, we have

$$\begin{aligned}&T_{\frac{{\bar{I}}\bar{f}}{f}}\left( fk_{-1}(I+1)\right) =P_{+}\left( \bar{f}\bar{I}\frac{I+1}{1+z}\right) \\&\quad =P_{+}\left( \bar{f}\frac{\bar{z}(\bar{I}+1)}{\bar{z}+1}\right) =P_{+}\left( \bar{z}\bar{f}\overline{k_{-1}}(\bar{I}+1)\right) =0. \end{aligned}$$

To see that the functions \((fk_{r_n}I-b_0(r_n)fk_{r_n})\) are orthogonal to \(fK_{I}\) note that a function \(h\in H^2\) is in \(K_{I}\) if and only if \(h= h - I P_{+}({\bar{I}}h)\). So, we have to check that for any \(h\in H^2\),

$$\begin{aligned} \langle fk_{r_n}I-b_0(r_n)fk_{r_n}, f(h-I P_{+}({\bar{I}}h))\rangle =0. \end{aligned}$$

Since the functions \(\{k_{\lambda }, \lambda \in {\mathbb {D}}\}\) are dense in \(H^2\), it is enough to show that for any \(\lambda \in {\mathbb {D}}\),

$$\begin{aligned}&\langle fk_{r_n}I-b_0(r_n)fk_{r_n}, f(k_{\lambda }-I P_{+}(\bar{I}k_{\lambda }))\rangle \\&\quad =\langle fk_{r_n}I-b_0(r_n)fk_{r_n}, f(k_{\lambda }-\overline{I(\lambda )}Ik_{\lambda })\rangle = 0. \end{aligned}$$

Finally, the last equality follows from

$$\begin{aligned} \langle fk_{r_n}I,fk_{\lambda }\rangle&= \frac{I(\lambda )k_{r_n}(\lambda )}{1- b(\lambda )}+ b_0(r_n)k_{r_n}(\lambda ),\\ \langle fk_{r_n}I, -\overline{I(\lambda )}fIk_{\lambda }\rangle&= -\frac{I(\lambda )k_{r_n}(\lambda )}{1-b(\lambda )},\\ \langle -b_0(r_n)fk_{r_n}, fk_{\lambda }\rangle&= -\frac{b_0(r_n)k_{r_n}(\lambda )}{1-b(\lambda )},\\ \end{aligned}$$

and

$$\begin{aligned} \langle -b_0(r_n)fk_{r_n},-\overline{I(\lambda )}fIk_{\lambda }\rangle&= b_0(r_n)I(\lambda )\frac{b_0(\lambda )k_{r_n}(\lambda )}{1-b(\lambda )}= \frac{b_0(r_n)b(\lambda )k_{r_n}(\lambda )}{1-b(\lambda )}. \end{aligned}$$

5 A remark on orthogonal complement of \(\mathcal {M}(a)\) in \(\mathcal {H}(b)\)

In this section we continue to assume that b is nonextreme. Recall that if the pair (ba) is special and \(f^2=\left( \frac{a}{1-b}\right) ^2 \)is not rigid or the pair (ba) is not special, then the space \(\mathcal {M}(a)\) is not dense in \(\mathcal {H}(b)\) ( [20]). In such a case let \(\mathcal {H}_0(b)\) denote the orthogonal complement of \(\mathcal {M}(a)\) in \(\mathcal {H}(b)\). Let Y be the restriction of the shift operator S to \(\mathcal {H}(b)\) and let \(Y_0\) be the compression of Y to the subspace \(\mathcal {H}_0(b)\). Then the spectrum of \(Y_0\) is contained in the unit circle. Moreover, if \(z_0\in \partial {\mathbb {D}}\) and k is a positive integer, then \(\ker (Y^*-\bar{z_0})^k\), which actually equals \(\ker (Y^*_0-\bar{z_0})^k\), lies in \({\mathcal {H}}_0(b)\). The necessary and sufficient conditions for \(\mathcal {H}_0(b)\) to have finite dimension are given in Chapter X of [21] (see also [4, Theorem 29.11]). In particular, the dimension of \( \mathcal {H}_0(b)\) is N if and only if the operator \(Y_0\) has distinct eigenvalues \(z_1\), \(z_2\), ..., \(z_s\) with their algebraic multiplicities \(n_1\), ..., \(n_s\), \(N=n_1+n_2+\dots +n_s\). Then \(\bar{z_1}\), \(\bar{z_2}\), ..., \(\bar{z_s}\) are the eigenvalues of \(Y^*_0\) with the same multiplicities, i.e., \(\dim \ker (Y_0-z_j)^{n_j}=\dim \ker (Y^*_0-\bar{z_j})^{n_j}\) and \(\mathcal {H}_0(b)\) is the direct sum of the subspaces \(\ker (Y^*_0-\bar{z_j})^{n_j}\), \(j=1,2,\ldots , s\).

On the other hand, if \(z_0\) is a point of \(\partial {\mathbb {D}}\) and b has an angular derivative in the sense of Carathéodory at \(z_0\), then the function given by

$$\begin{aligned} k_{z_0}^b(z)= \dfrac{1-\overline{b(z_0)}b(z)}{1-\bar{z_0}z}, \end{aligned}$$
(9)

where \(b(z_0) \) is the nontangential limit of b at \(z_0\), is in \(\mathcal {H}(b)\) (see [21, VI-4,5], [4, Theorem 21.1]). In this section we actually show that \(k_{z_0}^b\) is in \(\mathcal {H}_0(b)\).

Here we consider the case when the eigenspaces corresponding to eigenvalues \(z_1\), \(z_2\), ..., \(z_s\) are one dimensional and show that then the space \({\mathcal {H}}_0(b)\) is spanned by the functions \(k_{z_1}^b\), \(k_{z_2}^b\), ..., \(k_{z_s}^b\).

For \(|\lambda |=1\) let \(\mu _{\lambda }\) denote the measure for which equality in (1) holds when b is replaced by \({\bar{\lambda }}b\). If we put \(F_{\lambda }(z)= \frac{a}{1-{\overline{\lambda }}b}\), then the Radon-Nikodym derivative of the absolutely continuous component of \(\mu _{\lambda }\) is \(|F_{\lambda }|^2\). Note also that \(\mathcal {H}(b)=\mathcal {H}({{\bar{\lambda }}} b)\).

In the proof of our main result in this section we use the following theorem proved in [21, X-13].

Sarason’s Theorem

Let \(z_0\) be a point of \(\partial \mathbb {D}\) and \(\lambda \) a point of \(\partial \mathbb {D}\) such that the measure \(\mu _\lambda \) is absolutely continuous. The following conditions are equivalent.

  1. (i)

    \(\bar{z_0}\) is an eigenvalue of \(Y^*\).

  2. (ii)

    The function \(\dfrac{F_\lambda (z)}{1-\bar{z_0}z}\) is in \(H^2\).

  3. (iii)

    The function b has an angular derivative in the sense of Carathéodory at \(z_0\).

In view of remark in Sect. 3 under the assumption of Sarason’s Theorem the operator \(T_{1-{\bar{\lambda }}b}T_{{\overline{F}}_\lambda }\) is an isometry of \(H^2\) onto \(\mathcal {H}(b)\). Let \(A_{\lambda }\) be an operator on \(H^2\) that intertwines \(T_{1-{\bar{\lambda }}b}T_{{\overline{F}}_\lambda }\) with the operator \(Y^*\), i.e.,

$$\begin{aligned} T_{1-{\bar{\lambda }}b}T_{{\overline{F}}_\lambda }A_\lambda =Y^*T_{1-{\bar{\lambda }}b}T_{{\overline{F}}_\lambda }. \end{aligned}$$
(10)

The operator \(A_\lambda \) is given by

$$\begin{aligned} A_\lambda =S^*-F_\lambda (0)^{-1}(S^*F_\lambda \otimes 1). \end{aligned}$$

It follows from the proof of Sarason’s theorem that if one of conditions (i)–(iii) holds true, then the space \(\ker (A_\lambda -\bar{z_0})\) is one dimensional and is spanned by the function

$$\begin{aligned} g(z)=\frac{F_{\lambda }(z)}{1-\bar{z_0}z}= F_{\lambda }(z)k_{z_0}(z). \end{aligned}$$
(11)

We also note that condition (iii) in Sarason’s Theorem is equivalent to the fact that the function \(k_{z_0}^b\) given by (9) is in \(\mathcal {H}(b)\).

Theorem 3

If the assumptions of Sarason’s theorem are satisfied and \(\bar{z_0}\) is an eigenvalue of \(Y_0^*\), then \(\ker (Y_0^*-\bar{z_0})\) is spanned by \(k_{z_0}^b\).

Proof

According to the remark at the beginning of this section \(\ker (Y_0^*-\bar{z_0})\) is equal to \(\ker (Y^*-\bar{z_0})\). Since the operator \(T_{1-{\bar{\lambda }}b}T_{{\overline{F}}_\lambda }\) is an isometry of \(H^2\) onto \(\mathcal {H}(b)\), (11) and (10) imply that the space \(\ker (Y^*-\bar{z_0})\) is spanned by

$$\begin{aligned} h= T_{1-{\overline{\lambda }}b}T_{{\overline{F}}_{\lambda }}g= T_{1-{\overline{\lambda }}b}T_{\overline{F}_{\lambda }}(F_{\lambda }k_{z_0}). \end{aligned}$$

We will show that \(h= Ck_{z_0}^b\). To this end we use the operator \(V_b\) given by (5). We know that for \(w\in \mathbb {D}\),

$$\begin{aligned} V_b((1-\overline{b(w)})k_w)=k^b_w \end{aligned}$$

(see [21, III-7], [4, Theorem 20.5]). Since \(\mathcal {H}(b)=\mathcal {H}({\bar{\lambda }}b)\), we have

$$\begin{aligned} V_{{\bar{\lambda }}b}((1-\lambda \overline{b(w)})k_{w})(z)&= (1-{\bar{\lambda }}b(z))( 1-\lambda \overline{b(w)}) \int _{\partial \mathbb {D}}\frac{|F_{\lambda }(e^{i\theta })|^2d\theta }{(1-\bar{w}e^{i\theta })(1-ze^{-i\theta })}\\&= (1-{\bar{\lambda }}b(z))T_{{\overline{F}}_{\lambda }}((1-\lambda \overline{b(w)})F_{\lambda }k_w)(z)=k_{w}^b(z). \end{aligned}$$

Let \(\{z_n\}\) be a sequence in \({\mathbb {D}}\) converging nontangentially to \(z_0\). Then

$$\begin{aligned} T_{1-{\overline{\lambda }}b}T_{\overline{F}_{\lambda }}((1-\lambda \overline{b(z_n)})F_{\lambda }k_{z_n})=k_{z_n}^b. \end{aligned}$$

Observe also that since \(\mu _{\lambda }\) is absolutely continuous, \(b(z_0)\ne \lambda \) ( [21, VI-7, VI-9]). Moreover, \(k_z^b\) tends to \(k_{z_0}^b\) in norm as z tends to \(z_0\) nontagentially (see [21, VI-4,5], [4, Theorem 21.1]). It then follows that the sequence \(\{(1-\lambda \overline{b(z_n)})F_{\lambda }k_{z_n}\}\) converges in \(H^2\), which in turn implies compact and pointwise convergence. Hence passing to the limit in the last equality yields

$$\begin{aligned} T_{1-{\overline{\lambda }}b}T_{\overline{F}_{\lambda }}(F_{\lambda }k_{z_0})= Ck_{z_0}^b, \end{aligned}$$

where \(C=(1-\lambda \overline{b(z_0)})^{-1}\). \(\square \)

Our last theorem is an immediate consequence of Theorem 3.

Theorem 4

If \(z_1\), \(z_2\), ..., \(z_s\) are the only eigenvalues of \(Y_0\) and each of them is of multiplicity one, then \(\mathcal {H}_0(b)\) is spanned by the functions \(k_{z_1}^b\), \(k_{z_2}^b\), ..., \(k_{z_s}^b\).

Finally, we remark that this theorem generalizes results obtained in [5] and in [15] for the case when pairs (ba) are rational.