1 Introduction and results

We write \({{\mathbb {T}}}\) for the unit circle \(\{\zeta \in {{\mathbb {C}}}:|\zeta |=1\}\) and m for the normalized arc length measure on \({{\mathbb {T}}}\); thus, \(dm(\zeta )=|d\zeta |/(2\pi )\). We then define the spaces \(L^p:=L^p({{\mathbb {T}}},m)\) in the usual way and let \(\Vert \cdot \Vert _p\) denote the standard norm on \(L^p\). Also, for \(1\le p\le \infty \), we introduce the Hardy space \(H^p\) by putting

$$\begin{aligned} H^p:=\{f\in L^p:\,{\widehat{f}}(n)=0\,\,\,{\mathrm{for}}\,n=-1,-2,\dots \}, \end{aligned}$$

where \({\widehat{f}}(n)\) is the nth Fourier coefficient of f given by

$$\begin{aligned} {\widehat{f}}(n):=\int _{{\mathbb {T}}}\overline{\zeta }^nf(\zeta )\,dm(\zeta ),\quad n\in {{\mathbb {Z}}}. \end{aligned}$$

The Poisson integral (i.e., harmonic extension) of an \(H^p\) function being holomorphic on the disk

$$\begin{aligned} {{\mathbb {D}}}:=\{z\in {{\mathbb {C}}}:\,|z|<1\} \end{aligned}$$

(see [15, Chapter II]), we may use this extension to view elements of \(H^p\) as holomorphic functions on \({{\mathbb {D}}}\) when convenient.

Furthermore, we write \(P_+\) (resp., \(P_-\)) for the orthogonal projection from \(L^2\) onto \(H^2\) (resp., onto \(\overline{z}\overline{H^2}=L^2\ominus H^2\)). By a classical theorem of M. Riesz (see [15, Chapter III]), each of these projections admits a bounded extension—or restriction—to \(L^p\), with \(1<p<\infty \), and maps \(L^p\) onto \(H^p\) (resp., onto \(\overline{z}\overline{H^p}\)).

Now suppose \(\theta \) is an inner function, meaning that \(\theta \in H^\infty \) and \(|\theta |=1\) a.e. on \({{\mathbb {T}}}\). The corresponding star-invariant (or model) subspace \(K^p_\theta \) is then defined by

$$\begin{aligned} K^p_\theta :=\{f\in H^p:\,\overline{z}\overline{f}\theta \in H^p\},\quad 1\le p\le \infty , \end{aligned}$$
(1.1)

so that \(K^p_\theta =H^p\cap \theta \overline{z}\overline{H^p}\). (When \(p=2\), yet another equivalent definition is \(K^2_\theta =H^2\ominus \theta H^2\).) It is clear from (1.1) that the antilinear isometry

$$\begin{aligned} f\mapsto \overline{z}\overline{f}\theta =:{\widetilde{f}} \end{aligned}$$
(1.2)

leaves \(K^p_\theta \) invariant. Also, it is well known (see [6, 20]) that each \(K^p_\theta \) is invariant under the backward shift operator

$$\begin{aligned} {\mathfrak {B}}:f\mapsto \frac{f-f(0)}{z},\quad f\in H^p, \end{aligned}$$

and conversely, that every closed and nontrivial \({\mathfrak {B}}\)-invariant subspace of \(H^p\), with \(1\le p<\infty \), arises in this way.

The functions belonging to some \(K^p_\theta \) space (i.e., noncyclic vectors of \({\mathfrak {B}}\)) are known as pseudocontinuable functions. In fact, they are characterized by the property of having a meromorphic pseudocontinuation to \({{\mathbb {D}}}_-:={{\mathbb {C}}}{\setminus }({{\mathbb {D}}}\cup {{\mathbb {T}}})\); that is, the function in question should agree a.e. on \({{\mathbb {T}}}\) with the boundary values of some meromorphic function of bounded characteristic in \({{\mathbb {D}}}_-\) (see [6] for details).

The orthogonal projection from \(H^2\) onto \(K^2_\theta \) is given by \(f\mapsto \theta P_-(\overline{\theta }f)\), and the M. Riesz theorem shows that the same formula provides, for \(1<p<\infty \), a bounded projection from \(H^p\) onto \(K^p_\theta \) parallel to \(\theta H^p\). This yields the direct sum decomposition

$$\begin{aligned} H^p=K^p_\theta \oplus \theta H^p,\quad 1<p<\infty , \end{aligned}$$
(1.3)

with orthogonality for \(p=2\).

Among the inner functions \(\theta \), of special relevance to us are Blaschke products. Recall that, for a sequence \({\mathcal {Z}}=\{z_j\}\subset {{\mathbb {D}}}\) with

$$\begin{aligned} \sum _j(1-|z_j|)<\infty , \end{aligned}$$
(1.4)

the associated Blaschke product is given by

$$\begin{aligned} B(z)=B_{{\mathcal {Z}}}(z):=\prod _j\frac{|z_j|}{z_j}\frac{z_j-z}{1-\overline{z}_jz} \end{aligned}$$

(if \(z_j=0\), then we set \(|z_j|/z_j=-1\)). The product converges uniformly on compact subsets of \({{\mathbb {D}}}\) and defines an inner function that vanishes precisely at the \(z_j\)’s; see [15, Chapter II]. If, in addition,

$$\begin{aligned} \inf _j|B'(z_j)|\,(1-|z_j|)>0, \end{aligned}$$
(1.5)

then we say that B is an interpolating Blaschke product. Accordingly, the sequences \({\mathcal {Z}}=\{z_j\}\) in \({{\mathbb {D}}}\) that satisfy (1.4) and (1.5), with \(B=B_{{\mathcal {Z}}}\), are called interpolating (or \(H^\infty \)-interpolating) sequences. By a celebrated theorem of Carleson (see [3] or [15, Chapter VII]), these are precisely the sequences \({\mathcal {Z}}\) with the property that

$$\begin{aligned} H^\infty \big |_{{\mathcal {Z}}}=\ell ^\infty . \end{aligned}$$

Here and below, the following standard notation (and terminology) is used. Given a sequence \({\mathcal {Z}}=\{z_j\}\) of pairwise distinct points in \({{\mathbb {D}}}\), the trace \(f\big |_{{\mathcal {Z}}}\) of a function \(f:{{\mathbb {D}}}\rightarrow {{\mathbb {C}}}\) is defined to be the sequence \(\{f(z_j)\}\); and if \({\mathcal {X}}\) is a certain function space on \({{\mathbb {D}}}\), then the corresponding trace space is

$$\begin{aligned} {\mathcal {X}}\big |_{{\mathcal {Z}}}:=\left\{ f\big |_{{\mathcal {Z}}}:\,f\in {\mathcal {X}}\right\} . \end{aligned}$$

We shall be concerned with interpolation problems for functions in star-invariant subspaces—specifically, for those in \(K^p_B\), where B is an interpolating Blaschke product. Some of the earlier results in this area can be found in [1, 7, 16, 18], while others, more relevant to our current topic, will be recalled presently.

First, we need yet another piece of notation. Given numbers \(p>0\), \(\gamma \in {{\mathbb {R}}}\) and a sequence \({\mathcal {Z}}=\{z_j\}\subset {{\mathbb {D}}}\), we write \(\ell ^p_\gamma ({\mathcal {Z}})\) for the set of all sequences \(\{w_j\}\subset {{\mathbb {C}}}\) satisfying

$$\begin{aligned} \sum _j|w_j|^p(1-|z_j|)^{\gamma }<\infty . \end{aligned}$$

Now, if \(1<p<\infty \) and if \(B=B_{{\mathcal {Z}}}\) is an interpolating Blaschke product with zero sequence \({\mathcal {Z}}\), then we have

$$\begin{aligned} K^p_B\big |_{{\mathcal {Z}}}=H^p\big |_{{\mathcal {Z}}}=\ell ^p_1({\mathcal {Z}}). \end{aligned}$$

Indeed, the left-hand equality follows from (1.3) with \(\theta =B\), while the other holds by a well-known theorem of Shapiro and Shields [21]. In addition, for each sequence \({\mathcal {W}}=\{w_j\}\) in \(\ell ^p_1({\mathcal {Z}})\), there is a unique function \(f\in K^p_B\) with \(f\big |_{{\mathcal {Z}}}={\mathcal {W}}\); the uniqueness is due to the fact that \(K^p_B\cap BH^p=\{0\}\).

The case of \(K^\infty _B\) is subtler, as the next result shows.

Theorem A

Suppose that \({\mathcal {Z}}=\{z_j\}\) is an interpolating sequence in \({{\mathbb {D}}}\) and \(B=B_{{\mathcal {Z}}}\) is the associated Blaschke product. Then we have

$$\begin{aligned} K^\infty _B|_{{\mathcal {Z}}}=\ell ^\infty \end{aligned}$$
(1.6)

if and only if

$$\begin{aligned} \sup \left\{ \sum _j\frac{1-|z_j|}{|\zeta -z_j|}:\,\zeta \in {{\mathbb {T}}}\right\} <\infty . \end{aligned}$$
(1.7)

This theorem is essentially a consequence of Hruščev and Vinogradov’s work in [19]; see also [5, Section 3] for details.

Condition (1.7) above is known as the uniform Frostman condition, and the sequences \({\mathcal {Z}}=\{z_j\}\) in \({{\mathbb {D}}}\) that obey it are called Frostman sequences. While a Frostman sequence need not be interpolating (in fact, its points are not even supposed to be pairwise distinct), it does necessarily split into finitely many interpolating sequences; see [19] for a proof. Finally, a Blaschke product whose zeros form a Frostman sequence will be referred to as a Frostman Blaschke product.

We mention in passing that, by a theorem of Vinogradov [22], the identity

$$\begin{aligned} K^\infty _{B^2}|_{{\mathcal {Z}}}=\ell ^\infty \end{aligned}$$

is valid whenever \({\mathcal {Z}}\) is an interpolating sequence and \(B=B_{{\mathcal {Z}}}\). It should be noted, however, that \(K^\infty _{B^2}\) is strictly larger than \(K^\infty _B\).

To describe the trace class \(K^\infty _B|_{{\mathcal {Z}}}\) in the general case (i.e., when (1.7) no longer holds), we first introduce a bit of notation. Once the interpolating sequence \({\mathcal {Z}}=\{z_j\}\) is fixed, we associate with each sequence \({\mathcal {W}}=\{w_j\}\) from \(\ell ^1_1({\mathcal {Z}})\) the conjugate sequence \(\widetilde{{\mathcal {W}}}=\{\widetilde{w}_k\}\) whose elements are

$$\begin{aligned} {\widetilde{w}}_k:=\sum _j\frac{w_j}{B'(z_j)\cdot (1-z_j\overline{z}_k)}\quad (k=1,2,\dots ). \end{aligned}$$
(1.8)

The absolute convergence of the series in (1.8) is ensured, for any \(k\in {{\mathbb {N}}}\), by the fact that \({\mathcal {W}}\in \ell ^1_1({\mathcal {Z}})\) in conjunction with (1.5). Because \(\ell ^1_1({\mathcal {Z}})\) contains \(\ell ^\infty \), as well as every \(\ell ^p_1({\mathcal {Z}})\) with \(1<p<\infty \), the sequence \(\widetilde{{\mathcal {W}}}\) is well defined whenever \({\mathcal {W}}\) belongs to one of these spaces.

The following result was established in [13].

Theorem B

Suppose that \({\mathcal {Z}}=\{z_j\}\) is an interpolating sequence in \({{\mathbb {D}}}\) and \(B=B_{{\mathcal {Z}}}\) is the associated Blaschke product. Given a sequence \({\mathcal {W}}\in \ell ^\infty \), one has \({\mathcal {W}}\in K^\infty _B\big |_{{\mathcal {Z}}}\) if and only if \(\widetilde{{\mathcal {W}}}\in \ell ^\infty \).

It was further conjectured in [13, 14] that the trace space \(K^1_B\big |_{{\mathcal {Z}}}\) is describable in similar terms, i.e., that the necessary conditions \({\mathcal {W}}\in \ell ^1_1({\mathcal {Z}})\) and \(\widetilde{{\mathcal {W}}}\in \ell ^1_1({\mathcal {Z}})\) are also sufficient for \({\mathcal {W}}\) to be in \(K^1_B\big |_{{\mathcal {Z}}}\). To the best of our knowledge, the conjecture is still open.

Here, our purpose is to supplement Theorem B by characterizing the values of smooth, not just bounded, functions in \(K^2_B\) on the (interpolating) sequence \({\mathcal {Z}}=B^{-1}(0)\). To be more precise, of concern are trace spaces of the form \(\left( K^2_B\cap X\right) \big |_{{\mathcal {Z}}}\), where X is a certain smoothness class on \({{\mathbb {T}}}\). Specifically, X will be one of the following spaces.

\(\bullet \) The Lipschitz–Zygmund space \(\Lambda ^\alpha =\Lambda ^\alpha ({{\mathbb {T}}})\) with \(\alpha >0\). This is the set of functions \(f\in C({{\mathbb {T}}})\) satisfying

$$\begin{aligned} \Vert \Delta _h^nf\Vert _\infty =O(|h|^\alpha ),\quad h\in {{\mathbb {R}}}, \end{aligned}$$

where n is some (any) integer with \(n>\alpha \), and \(\Delta _h^n\) denotes the nth order difference operator with step h. (As usual, the difference operators \(\Delta _h^k\) are defined inductively: we put

$$\begin{aligned} (\Delta _h^1f)(\zeta ):=f(e^{ih}\zeta )-f(\zeta ),\quad \zeta \in {{\mathbb {T}}}, \end{aligned}$$

and \(\Delta _h^kf:=\Delta _h^1\Delta _h^{k-1}f\) for \(k\ge 2\).)

\(\bullet \) \(\mathrm{BMO}=\mathrm{BMO}({{\mathbb {T}}})\), the space of functions of bounded mean oscillation on \({{\mathbb {T}}}\). Recall that an integrable function f on \({{\mathbb {T}}}\) belongs to \(\mathrm{BMO}\) if and only if

$$\begin{aligned} \Vert f\Vert _*:=\left| \int _{{\mathbb {T}}}f\,dm\right| +\sup _I\frac{1}{m(I)}\int _I|f-f_I|\,dm<\infty , \end{aligned}$$

where \(f_I:=m(I)^{-1}\int _I f\,dm\); the supremum is taken over the open arcs \(I\subset {{\mathbb {T}}}\). Even though \(\mathrm{BMO}\) contains discontinuous and unbounded functions, there are reasons for viewing it as a smoothness class. In a sense, it corresponds to the endpoint as \(\alpha \rightarrow 0\) of the \(\Lambda ^\alpha \) scale. We also need the analytic subspace \(\mathrm{BMOA}:=\mathrm{BMO}\cap H^2\).

\(\bullet \) The Gevrey class \(G_\alpha =G_\alpha ({{\mathbb {T}}})\) with \(\alpha >0\). This is the set of functions \(f\in C^\infty ({{\mathbb {T}}})\) satisfying

$$\begin{aligned} \Vert f^{(n)}\Vert _\infty \le Q_f^{n+1}(n!)^{1+1/\alpha },\quad n=0,1,2,\dots , \end{aligned}$$

with some constant \(Q_f>0\). Here, we write \(f^{(n)}(e^{it})\) for the nth order derivative of the function \(t\mapsto f(e^{it})\), which is assumed to be \(C^\infty \)-smooth on \({{\mathbb {R}}}\).

\(\bullet \) The Sobolev space \({\mathcal {L}}^p_s={\mathcal {L}}^p_s({{\mathbb {T}}})\) with \(1<p<\infty \) and \(s>0\), defined by

$$\begin{aligned} {\mathcal {L}}^p_s:=\{f\in L^p:\,f^{(s)}\in L^p\}, \end{aligned}$$

with the appropriate interpretation of the (possibly fractional) derivative \(f^{(s)}\). Precisely speaking, we write \(f^{(s)}\in L^p\) to mean that there is a function \(g\in L^p\) satisfying \({\widehat{g}}(n)=(in)^s{\widehat{f}}(n)\) for all \(n\in {{\mathbb {Z}}}\).

For each of these choices of X, we now characterize the sequences \({\mathcal {W}}\) from the trace space \(\left( K^2_B\cap X\right) \big |_{{\mathcal {Z}}}\) in terms of the conjugate sequence \(\widetilde{{\mathcal {W}}}\), as defined by (1.8) above. The description always involves a certain decay condition (or growth restriction) on \(\widetilde{{\mathcal {W}}}\), as we shall presently see.

Theorem 1.1

Let \(\alpha >0\), \(1<p<\infty \) and \(s>0\). Also, let X be one of the following spaces: \(\Lambda ^\alpha \), \(\mathrm{BMO}\), \(G_\alpha \) or \({\mathcal {L}}^p_s\). Given an interpolating Blaschke product \(B=B_{{\mathcal {Z}}}\) with zeros \({\mathcal {Z}}=\{z_k\}\) and a sequence \({\mathcal {W}}=\{w_k\}\in \ell ^2_1({\mathcal {Z}})\), we have

$$\begin{aligned} {\mathcal {W}}\in \left( K^2_B\cap X\right) \big |_{{\mathcal {Z}}} \end{aligned}$$

if and only if

(a) \(|\widetilde{w}_k|=O\left( (1-|z_k|)^\alpha \right) \) when \(X=\Lambda ^\alpha \),

(b) \(\widetilde{{\mathcal {W}}}\in \ell ^\infty \) when \(X=\mathrm{BMO}\),

(c) there is a constant \(c>0\) such that

$$\begin{aligned} |{\widetilde{w}}_k|=O\left( \exp \left( -\frac{c}{(1-|z_k|)^\alpha }\right) \right) \end{aligned}$$

when \(X=G_\alpha \),

(d) \(\widetilde{{\mathcal {W}}}\in \ell ^p_{1-sp}({\mathcal {Z}})\) when \(X={\mathcal {L}}^p_s\).

The intersection \(K^2_B\cap \mathrm{BMO}\), which corresponds to case (b) above, will be henceforth denoted by \(K_{*B}\). Similarly, for a general inner function \(\theta \), we define

$$\begin{aligned} K_{*\theta }:=K^2_\theta \cap \mathrm{BMO}. \end{aligned}$$

Comparing Theorem B with the \(\mathrm{BMO}\) part of Theorem 1.1, we see that the structure of the trace space \(K^\infty _B\big |_{{\mathcal {Z}}}\) is remarkably similar to that of \(K_{*B}\big |_{{\mathcal {Z}}}\). In light of this observation, we may wonder what the \(\mathrm{BMO}\) counterpart of Theorem A could look like. Specifically, we may ask if there exist infinite Blaschke products \(B=B_{{\mathcal {Z}}}\) for which the trace space \(K_{*B}\big |_{{\mathcal {Z}}}\) is completely determined by the natural (and necessary) logarithmic growth condition on the values.

To be more precise, suppose that \({\mathcal {Z}}=\{z_k\}\) is a sequence of pairwise distinct points in \({{\mathbb {D}}}\), and write \(\ell ^\infty _{\log }({\mathcal {Z}})\) for the space of sequences \({\mathcal {W}}=\{w_k\}\subset {{\mathbb {C}}}\) with

$$\begin{aligned} |w_k|=O\left( \log \frac{2}{1-|z_k|}\right) . \end{aligned}$$

It is well known (and easily shown) that every \(f\in \mathrm{BMOA}\) satisfies

$$\begin{aligned} |f(z)|=O\left( \log \frac{2}{1-|z|}\right) ,\quad z\in {{\mathbb {D}}}, \end{aligned}$$

so \(\mathrm{BMOA}\big |_{{\mathcal {Z}}}\) is always contained in \(\ell ^\infty _{\log }({\mathcal {Z}})\). The equality

$$\begin{aligned} \mathrm{BMOA}\big |_{{\mathcal {Z}}}=\ell ^\infty _{\log }({\mathcal {Z}}) \end{aligned}$$
(1.9)

obviously need not hold in general, but it does actually occur for some infinite sequences \({\mathcal {Z}}=\{z_k\}\) (which form a tiny subfamily among the \(H^\infty \)-interpolating sequences). For instance, (1.9) will be valid provided that

$$\begin{aligned} |z_j-z_k|\ge c(1-|z_j|)^s,\quad j\ne k, \end{aligned}$$

for some constants \(c>0\) and \(s\in (0,\frac{1}{2})\); see [10, Theorem 11].

The question is what happens to (1.9) when \(\mathrm{BMOA}\) gets replaced by its subspace \(K_{*B}\), with \(B=B_{{\mathcal {Z}}}\). The property that arises is thus

$$\begin{aligned} K_{*B}\big |_{{\mathcal {Z}}}=\ell ^\infty _{\log }({\mathcal {Z}}), \end{aligned}$$
(1.10)

and we regard it as an analogue of (1.6) in the \(\mathrm{BMO}\) setting. In contrast to (1.6), however, (1.10) does not lead to any nontrivial class of sequences. Indeed, our next result shows that (1.10) is only possible when \({\mathcal {Z}}=B^{-1}(0)\) is a finite set.

Proposition 1.2

Whenever \(B=B_{{\mathcal {Z}}}\) is an infinite Blaschke product with simple zeros, the trace space \(K_{*B}\big |_{{\mathcal {Z}}}\) is properly contained in \(\ell ^\infty _{\log }({\mathcal {Z}})\).

This will be deduced from another result, which deals with the case of a general inner function \(\theta \) and asserts an amusing lack of duality between the star-invariant subspaces \(K^1_\theta \) and \(K_{*\theta }\).

It is well known that, for \(1<p<\infty \), the dual of the Hardy space \(H^p\) (under the pairing \(\langle f,g\rangle =\int _{{\mathbb {T}}}f\overline{g}\,dm\)) is \(H^q\) with \(q=p/(p-1)\), while the dual of \(H^1\) is \(\mathrm{BMOA}\); see, e.g., [15, Chapter VI]. The former duality relation has a natural counterpart in the \(K^p_\theta \) setting, namely \((K^p_\theta )^*=K^q_\theta \) for p and q as above (see [4, Lemma 4.2]), and one may wonder if the identity \((K^1_\theta )^*=K_{*\theta }\) has any chance of being true, at least for some inner functions \(\theta \). Our last theorem says that this is never the case, except when \(\theta \) is a finite Blaschke product.

Theorem 1.3

Given an inner function \(\theta \), other than a finite Blaschke product, there exists a non-\(\mathrm{BMO}\) function \(g\in K^2_\theta \) such that the functional

$$\begin{aligned} f\mapsto \int _{{\mathbb {T}}}f\overline{g}\,dm, \end{aligned}$$

defined initially for \(f\in K^2_\theta \), extends continuously to \(K^1_\theta \).

In other words, whenever \(\theta \) is an “interesting” (i.e., nonrational) inner function, \(K_{*\theta }\) is properly contained in \((K^1_\theta )^*\). One might compare this non-duality result with Bessonov’s duality theorem for \(K^1_\theta \) that appears in [2]. There, \(\theta \) was assumed to be a one-component inner function, meaning that the set \(\{z\in {{\mathbb {D}}}:|\theta (z)|<\varepsilon \}\) is connected for some \(\varepsilon \in (0,1)\), and the dual of \(K^1_\theta \cap zH^1\) was identified with a certain discrete \(\mathrm{BMO}\) space on \({{\mathbb {T}}}\).

In the remaining part of the paper, we first list a number of auxiliary facts (these are collected in Sect. 2) and then use them to prove our current results. The proofs are in Sects. 3 and 4.

2 Preliminaries

Several background results will be needed. When stating the first of these, we shall assume that X is one of our smoothness spaces (namely, \(\Lambda ^\alpha \), \(\mathrm{BMO}\), \(G_\alpha \) or \({\mathcal {L}}^p_s\)), the admissible values of the parameters \(\alpha \), p and s being as above.

Lemma 2.1

Let \(f\in H^2\) and let \(B=B_{{\mathcal {Z}}}\) be an interpolating Blaschke product with zeros \({\mathcal {Z}}=\{z_k\}\). In order that \(P_-(\overline{B}f)\in X\), it is necessary and sufficient that

(a) \(|f(z_k)|=O\left( (1-|z_k|)^\alpha \right) \) when \(X=\Lambda ^\alpha \),

(b) \(\{f(z_k)\}\in \ell ^\infty \) when \(X=\mathrm{BMO}\),

(c) for some \(c>0\),

$$\begin{aligned} |f(z_k)|=O\left( \exp \left( -\frac{c}{(1-|z_k|)^\alpha }\right) \right) \end{aligned}$$

when \(X=G_\alpha \),

(d) \(\{f(z_k)\}\in \ell ^p_{1-sp}({\mathcal {Z}})\) when \(X={\mathcal {L}}^p_s\).

The statements corresponding to parts (a) and (b) were proved in [8] as Theorems 4.1 and 5.2. For parts (c) and (d), we refer to [11]; specifically, see Theorems 1 and 7 therein.

Another (well-known) fact to be used below is that the space \(\mathrm{BMOA}\) enjoys the so-called K-property of Havin, as defined in [17]. The precise meaning of this assertion is as follows.

Lemma 2.2

For every \(\psi \in H^\infty \), the Toeplitz operator \(T_{\overline{\psi }}\) given by

$$\begin{aligned} T_{\overline{\psi }}f:=P_+(\overline{\psi }f),\quad f\in \mathrm{BMOA}, \end{aligned}$$

maps \(\mathrm{BMOA}\) boundedly into itself.

To prove this, it suffices to observe (in the spirit of [17]) that \(T_{\overline{\psi }}\) is the adjoint of the multiplication operator \(g\mapsto \psi g\), which is obviously bounded on \(H^1\).

Before proceeding, we need to introduce a bit of notation. Namely, with an inner function \(\theta \) and a number \(\varepsilon \in (0,1)\) we associate the sublevel set

$$\begin{aligned} \Omega (\theta ,\varepsilon ):=\{z\in {{\mathbb {D}}}:\,|\theta (z)|<\varepsilon \}. \end{aligned}$$

The following result is a restricted version of [9, Theorem 1].

Lemma 2.3

Suppose that \(f\in \mathrm{BMOA}\) and \(\theta \) is an inner function. Then \(f\overline{\theta }\in \mathrm{BMO}\) if and only if

$$\begin{aligned} \sup \{|f(z)|:\,z\in \Omega (\theta ,\varepsilon )\}<\infty \end{aligned}$$
(2.1)

for some (or every) \(\varepsilon \) with \(0<\varepsilon <1\).

Next, we recall a remarkable maximum principle for \(K^2_\theta \) functions that was established by Cohn in [5].

Lemma 2.4

Let \(\theta \) be inner, and suppose \(f\in K^2_\theta \) is a function that satisfies (2.1) for some \(\varepsilon \in (0,1)\). Then \(f\in H^\infty \).

Our last lemma reproduces yet another result of Cohn (see [4, p. 737]), which characterizes the inner functions \(\theta \) with the property that \(K_{*\theta }\) contains only bounded functions. This characterization is, in turn, a consequence of Hruščev and Vinogradov’s earlier work from [19] on the multipliers of Cauchy type integrals.

Lemma 2.5

Let \(\theta \) be an inner function. Then \(K_{*\theta }=K^\infty _\theta \) if and only if \(\theta \) is a Frostman Blaschke product.

We also refer to [12, Theorem 1.7] for a refinement of this result in terms of \({\mathrm{inn}}(K_{*\theta })\), the set of inner factors for functions from \(K_{*\theta }\).

3 Proof of Theorem 1.1

We shall only give a detailed proof of part (a), the other cases being similar. Since \({\mathcal {W}}=\{w_k\}\in \ell ^2_1({\mathcal {Z}})\), we know that there exists a unique \(f\in K^2_B\) such that \(f\big |_{{\mathcal {Z}}}={\mathcal {W}}\). Therefore, in order that

$$\begin{aligned} {\mathcal {W}}\in \left( K^2_B\cap \Lambda ^\alpha \right) \big |_{{\mathcal {Z}}} \end{aligned}$$
(3.1)

it is necessary and sufficient that

$$\begin{aligned} f\in \Lambda ^\alpha . \end{aligned}$$
(3.2)

To find out when the latter condition holds, we apply Lemma 2.1, part (a), to the function \(g:=\overline{z}\overline{f}B\) in place of f. (Note that \(g\in H^2\) because \(f\in K^2_B\).) This tells us that \(P_-(\overline{B}g)\in \Lambda ^\alpha \) if and only if

$$\begin{aligned} |g(z_k)|=O\left( (1-|z_k|)^\alpha \right) ,\quad k\in {{\mathbb {N}}}. \end{aligned}$$
(3.3)

On the other hand,

$$\begin{aligned} P_-(\overline{B}g)=P_-(\overline{z}\overline{f})=\overline{z}\overline{f}, \end{aligned}$$

and it is clear that the function \(\overline{z}\overline{f}\) belongs to \(\Lambda ^\alpha \) if and only if f does. Thus, we may rephrase (3.2) as (3.3). To arrive at a further—and definitive—restatement of (3.3), we need to express the numbers \(g(z_k)\) in terms of \({\mathcal {W}}\). For \(z\in {{\mathbb {D}}}\), Cauchy’s formula yields

$$\begin{aligned} g(z)=\int _{{{\mathbb {T}}}}\frac{g(\zeta )}{1-\overline{\zeta }z}\,dm(\zeta ). \end{aligned}$$

Consequently,

$$\begin{aligned} \begin{aligned} \overline{g(z_k)}&=\int _{{{\mathbb {T}}}}\frac{\overline{g(\zeta )}}{1-\zeta \overline{z}_k}\,dm(\zeta ) =\int _{{{\mathbb {T}}}}\frac{\zeta f(\zeta )\overline{B(\zeta )}}{1-\zeta \overline{z}_k}\,dm(\zeta )\\&=\frac{1}{2\pi i}\int _{{{\mathbb {T}}}}\frac{f(\zeta )}{B(\zeta )\cdot (1-\zeta \overline{z}_k)}\,d\zeta . \end{aligned} \end{aligned}$$

Computing the last integral by residues, while recalling that \(f(z_j)=w_j\), we find that

$$\begin{aligned} \overline{g(z_k)}=\sum _j\frac{w_j}{B'(z_j)\cdot (1-z_j\overline{z}_k)}={\widetilde{w}}_k \end{aligned}$$
(3.4)

for each \(k\in {{\mathbb {N}}}\). (To justify the application of the residue theorem, one may begin by evaluating the integral over the circle \(r_n{{\mathbb {T}}}\), where \(\{r_n\}\subset (0,1)\) is a suitable sequence tending to 1, and then pass to the limit as \(n\rightarrow \infty \).)

Finally, we use (3.4) to rewrite (3.3) in the form

$$\begin{aligned} |{\widetilde{w}}_k|=O\left( (1-|z_k|)^\alpha \right) ,\quad k\in {{\mathbb {N}}}. \end{aligned}$$
(3.5)

The equivalence of (3.1) and (3.5) is thereby established, proving the \(\Lambda ^\alpha \) part of the theorem.

The remaining statements (i.e., those involving \(\mathrm{BMO}\), \(G_\alpha \) and \({\mathcal {L}}^p_s\)) are proved similarly, by combining the appropriate parts of Lemma 2.1 with identity (3.4).

4 Proofs of Proposition 1.2 and Theorem 1.3

We begin by proving Theorem 1.3. Once this is done, Proposition 1.2 will be derived as a corollary.

Proof of Theorem 1.3

Given an inner function \(\theta \) distinct from a finite Blaschke product, we want to find a function \(g\in K^2_\theta {\setminus }\mathrm{BMO}\) that induces a bounded linear functional on \(K^1_\theta \). We shall distinguish two cases.

Case 1

Assume that \(\theta \) is an infinite Frostman Blaschke product. Its zero sequence, say \({\mathcal {Z}}=\{z_j\}\), must then have a limit point on \({{\mathbb {T}}}\). Of course, nothing is lost by assuming that \({\mathcal {Z}}\) clusters at 1. Now let

$$\begin{aligned} \varphi (z):=\log (1-z), \end{aligned}$$

where “log” stands for the holomorphic branch of the logarithm that lives on the right half-plane and satisfies \(\log 1=0\). We have \(\varphi \in \mathrm{BMOA}\) (because \({\mathrm{Im}}\,\varphi \in L^\infty \)), so the corresponding linear functional acts boundedly on \(H^1\) and hence on \(K^1_\theta \). Clearly, the same functional on \(K^1_\theta \) is also induced, in a similar manner, by the function

$$\begin{aligned} g:=\theta P_-(\overline{\theta }\varphi ), \end{aligned}$$

which is the orthogonal projection (in \(H^2\)) of \(\varphi \) onto \(K^2_\theta \). Precisely speaking, the functional

$$\begin{aligned} f\mapsto \int _{{\mathbb {T}}}f\overline{g}\,dm\left( =\int _{{\mathbb {T}}}f\overline{\varphi }\,dm\right) ,\quad f\in K^2_\theta , \end{aligned}$$

extends continuously to \(K^1_\theta \).

We know that \(g\in K^2_\theta \), and to conclude that g does the job, we only need to check that

$$\begin{aligned} g\notin \mathrm{BMO}. \end{aligned}$$
(4.1)

To this end, observe first that \(\sup _j|\varphi (z_j)|=\infty \) and hence, a fortiori,

$$\begin{aligned} \sup \{|\varphi (z)|:\,z\in \Omega (\theta ,\varepsilon )\}=\infty \end{aligned}$$

for every \(\varepsilon \in (0,1)\). By Lemma 2.3, this implies that \(\overline{\theta }\varphi \notin \mathrm{BMO}\). On the other hand,

$$\begin{aligned} \overline{\theta }\varphi =P_-(\overline{\theta }\varphi )+P_+(\overline{\theta }\varphi ), \end{aligned}$$

where the last term, \(P_+(\overline{\theta }\varphi )\), is in \(\mathrm{BMOA}(\subset \mathrm{BMO})\) thanks to Lemma 2.2. It follows readily that \(P_-(\overline{\theta }\varphi )\notin \mathrm{BMO}\). In particular, \(P_-(\overline{\theta }\varphi )\notin L^\infty \) (just note that \(L^\infty \subset \mathrm{BMO}\)). Equivalently, the function \(\theta P_-(\overline{\theta }\varphi )=g\) is not in \(L^\infty \).

Now, if g were in \(\mathrm{BMO}\), then we would have \(g\in K_{*\theta }\); and since our current assumption on \(\theta \) yields \(K_{*\theta }=K^\infty _\theta \) (in accordance with Lemma 2.5), g would have to be bounded, which it is not. This proves (4.1).

Case 2

Assume that \(\theta \) is not a Frostman Blaschke product. This time, using Lemma 2.5 again, we can find an unbounded function \(h\in K_{*\theta }\). We have then \({\widetilde{h}}:=\overline{z}\overline{h}\theta \in K^2_\theta \), and we go on to claim that \({\widetilde{h}}\notin \mathrm{BMO}\). (Here and below, the “tilde operation” (1.2) is being used repeatedly.) Indeed, if \({\widetilde{h}}\) were in \(\mathrm{BMO}\), then so would be \(h\overline{\theta }\), and Lemma 2.3 would tell us that

$$\begin{aligned} \sup \{|h(z)|:\,z\in \Omega (\theta ,\varepsilon )\}<\infty \end{aligned}$$

for some (any) \(\varepsilon \in (0,1)\). This, however, would imply that \(h\in H^\infty \) by virtue of Lemma 2.4, whereas h is actually unbounded by assumption.

Now we know that \({\widetilde{h}}\in K^2_\theta {\setminus }\mathrm{BMO}\), and we proceed by showing that \({\widetilde{h}}\) generates a continuous linear functional on \(K^1_\theta \). This will allow us to conclude that \({\widetilde{h}}\) is eligible as g (the function we are looking for), and the proof will be complete.

Given \(f\in K^2_\theta \), we have the elementary identity \(\overline{f}{\widetilde{h}}={\widetilde{f}}\overline{h}\). Recalling also the facts that \({\widetilde{f}}\in H^2(\subset H^1)\) and \(h\in \mathrm{BMOA}\), we use the duality relation \((H^1)^*=\mathrm{BMOA}\) to infer that

$$\begin{aligned} \begin{aligned} \left| \int _{{{\mathbb {T}}}}f\,\overline{{\widetilde{h}}}\,dm\right|&= \left| \int _{{{\mathbb {T}}}}\overline{f}\,{\widetilde{h}}\,dm\right| =\left| \int _{{{\mathbb {T}}}}{\widetilde{f}}\,\overline{h}\,dm\right| \\&\le C\Vert {\widetilde{f}}\Vert _1\Vert h\Vert _*=C\Vert f\Vert _1\Vert h\Vert _* \end{aligned} \end{aligned}$$

with some absolute constant \(C>0\). Consequently, for \(g={\widetilde{h}}\), the (densely defined) functional

$$\begin{aligned} f\mapsto \int _{{{\mathbb {T}}}}f\overline{g}\,dm \end{aligned}$$
(4.2)

is indeed continuous on \(K^1_\theta \), and we are done. \(\square \)

Proof of Proposition 1.2

Assuming that \(B=B_{\mathcal {Z}}\) is an infinite Blaschke product with zeros \({\mathcal {Z}}=\{z_j\}\), where the \(z_j\)’s are pairwise distinct, we want to find a sequence of values \({\mathcal {W}}=\{w_j\}\) in \(\ell ^\infty _{\log }({\mathcal {Z}})\) that is not the trace of any \(K_{*B}\) function on \({\mathcal {Z}}\). Consider, for each \(j\in {{\mathbb {N}}}\), the function

$$\begin{aligned} f_j(z):=\frac{1}{1-\overline{z}_jz} \end{aligned}$$

and note that \(f_j\in K^2_B\). Observe also that

$$\begin{aligned} \Vert f_j\Vert _1\le M\log \frac{2}{1-|z_j|},\quad j\in {{\mathbb {N}}}, \end{aligned}$$
(4.3)

for some fixed constant \(M>0\).

Now, Theorem 1.3 provides us with a function \(g\in K^2_B{\setminus }\mathrm{BMO}\) such that the associated functional (4.2), defined initially for \(f\in K^2_B\), acts boundedly on \(K^1_B\), say with norm \(N_g\). When applied to \(f=f_j\), this functional takes the value \(\overline{g(z_j)}\); indeed, Cauchy’s formula gives

$$\begin{aligned} \int _{{\mathbb {T}}}\overline{f}_jg\,dm=g(z_j) \end{aligned}$$

for each j. In conjunction with (4.3), this yields

$$\begin{aligned} |g(z_j)|\le N_g\Vert f_j\Vert _1\le MN_g\log \frac{2}{1-|z_j|},\quad j\in {{\mathbb {N}}}. \end{aligned}$$
(4.4)

Finally, we put

$$\begin{aligned} w_j:=g(z_j),\quad j\in {{\mathbb {N}}}. \end{aligned}$$

The sequence \({\mathcal {W}}=\{w_j\}\) is then in \(\ell ^\infty _{\log }({\mathcal {Z}})\), as (4.4) shows, while

$$\begin{aligned} {\mathcal {W}}\notin K_{*B}\big |_{\mathcal {Z}} \end{aligned}$$
(4.5)

as required. To verify (4.5), it suffices to note that g is the only function in \(K^2_B\) that interpolates \({\mathcal {W}}\) on \({\mathcal {Z}}\) (indeed, a \(K^2_B\) function is uniquely determined by its trace on \({\mathcal {Z}}=B^{-1}(0)\)), whereas \(g\notin \mathrm{BMO}\). The proof is complete. \(\square \)

Remark

We have seen above that if \(g\in K^2_B\), with \(B=B_{{\mathcal {Z}}}\), and if the functional (4.2) is continuous on \(K^1_B\), then \(g\big |_{{\mathcal {Z}}}\in \ell ^\infty _{\log }({\mathcal {Z}})\). Now, if \({\mathcal {Z}}\) has the \(\mathrm{BMOA}\)-interpolating property (1.9), then every sequence \({\mathcal {W}}\) in \(\ell ^\infty _{\log }({\mathcal {Z}})\) is actually writable as \(g\big |_{{\mathcal {Z}}}\) for some \(g\in K^2_B\) that induces a continuous linear functional on \(K^1_B\). (To see why, take \(G\in \mathrm{BMOA}\) with \(G\big |_{{\mathcal {Z}}}={\mathcal {W}}\) and then put \(g=BP_-(\overline{B}G)\), so that g is the orthogonal projection of G onto \(K^2_B\).) Of course, things become different if condition (1.9) is dropped. For instance, there are interpolating sequences \({\mathcal {Z}}\) for which \(\ell ^\infty _{\log }({\mathcal {Z}})\not \subset \ell ^2_1({\mathcal {Z}})\); and if this is the case, then no sequence in \(\ell ^\infty _{\log }({\mathcal {Z}}){\setminus }\ell ^2_1({\mathcal {Z}})\) is the trace of any \(H^2\) function on \({\mathcal {Z}}\).