1 Introduction

Let \(\mathbb {T}\) denote the unit circle and \(\mathbb {D}\) denote the unit disc. For a positive integer N, \(L^2(\mathbb {T}^N)\), \(L^{\infty }(\mathbb {T}^N)\) denote the standard \(L^2\) and \(L^{\infty }\) spaces with respect to normalized Lebesque measure \(m_N\) on N-torus \(\mathbb {T}^N\).

Let \(H^2(\mathbb {D}^N)\) denote the Hardy space. Notice that these functions have a dual nature. Once we can see them as a holomorphic function on \(\mathbb {D}^N\). On the other hand, they can be treated as an element of \(L^2(\mathbb {T}^N)\). For \(\varphi \in L^\infty (\mathbb {T}^N)\), a (essentially) bounded element of \(L^2(\mathbb {T}^N)\), \(M_\varphi \) denotes the multiplication operator on \(L^2(\mathbb {T}^N)\), i.e., \(M_\varphi f= \varphi f\) for \(f\in L^2(\mathbb {T}^N)\).

In the paper we study conjugation operators on N–torus. By definition it is a bounded, antilinear operator \(C :L^2(\mathbb {T}^N)\rightarrow L^2(\mathbb {T}^N)\) such that \(C^2 = I\) and \(\langle C f, C g \rangle = \langle g, f \rangle \) for every f, \(g \in L^2(\mathbb {T}^N)\). Recall that such operator is isometric and onto.

In the space \(L^2(\mathbb {T}^N)\) we can consider two natural conjugations: J and \(J^\#\). Operation J conjugates the value of the functions (see (2.1)), while \(J^\#\) conjugates the Fourier coefficients of the functions (see (2.2)). The behaviour of these two conjugations with respect to \(M_{z_i}\), \(i=1,\dots ,N\) is different. The conjugation \(J^\#\) commutes with all \(M_{z_i}\). On the other hand, the conjugation J intertwines \(M_z\) and \(M_{\bar{z}}\) for all \(i=1,\dots ,N\) (see (2.3)).

Theory of conjugations on the unit circle \(\mathbb {T}\) (i.e., \(N=1\)) has been intensively studied in the last few years (see e.g. [4, 5, 8, 9]). In [5, Theorem 2.2 and 2.4] and [4, Theorem 4.1], all conjugations commuting with (or intertwining) multiplication operator by the independent variable were described. One of the consequences of those results are characterizations (see [5, Theorems 5.2, Proposition 5.6]) of all conjugations leaving invariant subspaces being invariant for multiplication by independent variable.

The importance of conjugations is underlined by its connections with physics, where \(\mathcal{PT}\mathcal{}\) symmetries in \(L^2(\mathbb {R})\) are considered, see [1,2,3]. For a connection between conjugations on \(L^2(\mathbb {T})\) and \(L^2(\mathbb {R})\) see [7]. \(\mathcal{PT}\mathcal{}\) symmetries can also be considered in \(L^2(\mathbb {R}^N).\)

The investigation in the following paper goes to multivariable case (\(N>1\)). The latter case becomes more interesting because of two reasons – the holomorphic function theory is much more complicated for \(N>1\) and the Beurling theorem characterizing invariant subspaces is no longer true.

The structure of the paper is as follows. In Sect. 2, we present some natural conjugations in \(L^2(\mathbb {T}^N)\) and give their basic properties. In Sect. 3, the characterization of conjugations intertwining with multiplication operators and its adjoints is settled. Sect. 4 presents a characterization of conjugations commuting with multiplication operators. Sects. 5, 6, 7 are devoted to conjugations and invariant subspaces. In Sect. 5, it is shown that there are no conjugations intertwining multiplication operators and its adjoint and leaving any subspace of \(H^2(\mathbb {D}^N)\) invariant. In Sects. 6 and 7, we characterize all conjugations commuting with multiplication operators and leaving invariant given invariant subspace of \(H^2(\mathbb {D}^N)\). Sect. 6 deals with invariant subspaces of type \(\theta H^2(\mathbb {D}^N)\), where \(\theta \) is an inner function. In Sect. 7, we consider subspaces fulfilling property (*), introduced in [6], which are different from those considered in Sect. 6.

2 Preliminaries

For a vector \(z=(z_1, \ldots , z_N) \in \mathbb {C}^N\), we define its conjugation by \(\bar{z}=(\overline{z_1}, \ldots , \overline{z_N})\).

In the space \(L^2(\mathbb {T}^N)\) there are two conjugations, which seems very natural. First of them is J, which is simply a conjugation of the function, i.e.,

$$\begin{aligned} J f = \bar{f}, \quad f \in L^2(\mathbb {T}^N).\end{aligned}$$
(2.1)

The second one is \(J^\#\), which derives from conjugating Fourier coefficients of functions in \(L^2(\mathbb {T}^N)\). Precisely,

$$\begin{aligned} J^\# f = f^\#, \quad f \in L^2(\mathbb {T}^N), \end{aligned}$$
(2.2)

where

$$\begin{aligned} f^\# (z) = \overline{f(\bar{z})}, \quad z \in \mathbb {T}^N. \end{aligned}$$

The conjugation \(J^\#\) keeps the Hardy space invariant (\(J^\#H^2(\mathbb {D}^N)=H^2(\mathbb {D}^N)\)). On the other hand, \(JH^2(\mathbb {D}^N)=\overline{H^2(\mathbb {D}^N)}\).

Let \(\mathcal {M}\) be a subspace of \(H^2(\mathbb {D}^N)\). We denote the set \(J^\# \mathcal {M}\) by \(\mathcal {M}^\#\). Moreover, if \(\mathcal {M}\) is closed, then \(\mathcal {M}\) is called invariant if \(z_i\mathcal {M}\subset \mathcal {M}\) for \(i=1,\dots ,N\).

Let us summarize some basic properties below.

Proposition 2.1

Let \(\mathcal {M}\) be a subspace of \(H^2(\mathbb {D}^N)\). Then:

  1. (a)

    \(\mathcal {M}^\# \subset H^2(\mathbb {D}^N)\),

  2. (b)

    If \(\mathcal {M}\) is invariant, then \(\mathcal {M}^\#\) is also invariant,

  3. (c)

    \(J\mathcal {M}\subset \overline{H^2(\mathbb {D}^N)}\).

The behaviour of J and \(J^\#\) with respect to the multiplication operators \(M_{z_i}\), \(i=1,\dots , N\), is also completely different. Precisely,

$$\begin{aligned} JM_{z_i}=M_{\bar{z}_i}J\qquad \text {and}\qquad J^\#M_{z_i}=M_{ z_i}J^\#\quad \text {for}\ i=1,\dots ,N. \end{aligned}$$
(2.3)

Let \(\psi \in L^\infty (\mathbb {T}^N)\). Then

$$\begin{aligned} (M_\psi J)M_{z_i}=M_{\bar{z}_i}(M_\psi J)\ \text {and}\ (M_\psi J^\#)M_{z_i}=M_{ z_i}(M_\psi J^\#)\ \text {for}\ i=1,\dots ,N. \end{aligned}$$
(2.4)

A unimodular function \(\psi \in L^\infty (\mathbb {T}^N)\) is called symmetric if \(\psi \psi ^\#=1 \) or equivalently \(\psi (z)=\psi (\bar{z})\) for \(z\in \mathbb {T}^N\).

Now, we set apart functions \(\psi \in L^\infty (\mathbb {T}^N)\) such that operators \(M_\psi J\) and \(M_\psi J^\#\) are conjugations.

Proposition 2.2

Let \(\psi \in L^\infty (\mathbb {T}^N)\). Then:

  1. (a)

    \(C=M_\psi J\) is a conjugation if and only if \(\psi \) is a unimodular function,

  2. (b)

    \(C=M_\psi J^\#\) is a conjugation if and only if \(\psi \) is a unimodular symmetric function.

Proof

We give a proof only for (b). Assume that \(C= M_\psi J^\#\) is a conjugation. Then, for any \(f,g\in L^2(\mathbb {T}^N)\), we have

$$\begin{aligned} \int g\bar{f}\,dm&=\langle g, f\rangle = \langle Cf,Cg\rangle =\langle \psi f^{\#},\psi g^{\#}\rangle \\ {}&=\int |\psi (z)|^2\,\overline{f(\bar{z})}g(\bar{z})\,dm(z)\\&=\int |\psi (\bar{z}) |^2\,\overline{f(z)}g(z)\,dm(z), \quad z \in \mathbb {T}^N. \end{aligned}$$

Hence \(|\psi |=1\) a.e. on \(\mathbb {T}^N\). For all \(f\in L^2(\mathbb {T}^N)\) we have

$$\begin{aligned} f=C^2 f= M_\psi J^{\#}M_\psi J^{\#}f= M_\psi J^{\#}({\psi f^{\#}})=\psi \psi ^\# f, \end{aligned}$$

which implies that \(\psi \psi ^\#=1\) a.e. on \(\mathbb {T}^N\). Therefore, \(\psi \) is symmetric, i.e., \(\psi (z)=\psi (\bar{z})\) a.e. on \(\mathbb {T}^N\). The reverse implication is clear. \(\square \)

Let \(\alpha \in L^\infty (\mathbb {T}^N)\) be a unimodular function. If \(\alpha \in H^2(\mathbb {D}^N)\), then we say that \(\alpha \) is inner. We say that an inner function \(\alpha \) is irreducible if it is not a product of two non-constant inner functions. For any inner functions \(\alpha ,\beta \), we write \(\alpha \leqslant \beta \) if \(\frac{\beta }{\alpha }\) is inner. In such case \(\beta H^2(\mathbb {D}^N)\subset \alpha H^2(\mathbb {D}^N)\). Finally, we say that \(\alpha \) is self–reflected if \(\alpha \leqslant \alpha ^\#\) and \(\alpha ^\# \leqslant \alpha \).

3 Characterization of \(M_z\)–Conjugations

Consider a conjugation C on \(L^2(\mathbb {T}^N)\). It is called \(M_z\)conjugationFootnote 1 if \(M_{z_i}C=CM_{\bar{z}_i}\), for \(i=1,\dots ,N\). Observe that the natural conjugation J on \(L^2(\mathbb {T}^N)\) defined by (2.1) is an \(M_z\)-conjugation.

Let us define a space \(\mathcal {L}^{\infty }(\mathbb {T}^N)=\{M_\varphi : \varphi \in L^{\infty }(\mathbb {T}^N)\}\). It will be used in the proof of the next theorem, which completely characterizes all \(M_z\)–conjugations on \(L^2(\mathbb {T}^N)\).

Theorem 3.1

Let C be a conjugation on \(L^2(\mathbb {T}^N)\). The following conditions are equivalent:

  1. (a)

    C is \(M_z\)–conjugation,

  2. (b)

    \(M_\varphi C=CM_{\bar{\varphi }}\) for all \(\varphi \in L^{\infty }(\mathbb {T}^N)\),

  3. (c)

    there is a unimodular function \(\psi \in L^{\infty }(\mathbb {T}^N)\) such that \(C=JM_{\psi }\),

  4. (d)

    there is a unimodular function \(\psi ^\prime \in L^{\infty }(\mathbb {T}^N)\) such that \(C=M_{\psi ^\prime } J\).

Note that function \(\psi ^\prime \) from condition (d) corresponds to function \(\bar{\psi }\) from condition (c).

Proof

Equivalence of (a) and (b) is a consequence of standard approximation procedure. Condition (c) is equivalent to (d). In fact, it is enough to show that (b) implies (c).

If \(\varphi \in L^{\infty }(\mathbb {T}^N)\), then

$$\begin{aligned} JCM_\varphi = JM_{\bar{\varphi }}C=M_\varphi JC. \end{aligned}$$

Thus JC is a bounded linear operator and JC commutes with all \(M_\varphi \in \mathcal {L}^{\infty }(\mathbb {T}^N)\). By [10, Theorem 1.20], \(\mathcal {L}^{\infty }(\mathbb {T}^N)\) is a m.a.s.a. (maximal abelian selfadjoint algebra) and as a consequence \(JC \in \mathcal {L}^{\infty }(\mathbb {T}^N)\). Therefore, there is a function \(\psi \in L^{\infty }(\mathbb {T}^N)\) such that \(JC=M_\psi \). Finally, \(C=JM_\psi \). \(\square \)

4 Characterization of \(M_z\)–Commuting conjugations

Let C be a conjugation on \(L^2(\mathbb {T}^N)\). C is called \(M_z\)commuting conjugation if \( M_{z_i} C=C M_{z_i}\), \(i=1,\dots ,N\). Notice that the natural conjugation \(J^\#\) on \(L^2(\mathbb {T}^N)\) defined by (2.2) is \(M_z\)–commuting conjugation.

The theorem below completely characterizes all \(M_z\)–commuting conjugations on \(L^2(\mathbb {T}^N)\).

Theorem 4.1

Let C be a conjugation on \(L^2(\mathbb {T}^N)\). The following conditions are equivalent:

  1. (a)

    C is an \(M_z\)–commuting conjugation,

  2. (b)

    \(M_\varphi C=C M_{\varphi ^{\#}}\) for all \(\varphi \in L^\infty (\mathbb {T}^N)\),

  3. (c)

    there is a symmetric unimodular function \(\psi \in L^\infty (\mathbb {T}^N)\) such that \(C=J^{\#}\,M_{\psi }\),

  4. (d)

    there is a symmetric unimodular function \(\psi ^\prime \in L^\infty (\mathbb {T}^N)\) such that \(C=M_{\psi ^\prime } J^{\#}\).

Note that function \(\psi ^\prime \) from condition (d) corresponds to function \(\psi ^\#\) from condition (c).

Proof

Assume (a) and note that \(M_\varphi C=C M_{\varphi ^{\#}}\) if \(\varphi \) is a polynomial. Now, equivalence of (a) and (b) is a consequence of standard approximation procedure. Clearly, condition (c) is equivalent to (d). It remains to show the implication from (b) to (c). If \(\varphi \in L^{\infty }(\mathbb {T}^N)\), then

$$\begin{aligned} J^\#CM_\varphi = J^\#M_{\varphi ^\#}C=M_\varphi J^\#C. \end{aligned}$$

Using again the fact that \(\mathcal {L}^{\infty }(\mathbb {T}^N)\) is a m.a.s.a., we get \(J^\#C \in \mathcal {L}^{\infty }(\mathbb {T}^N)\). Therefore, there is a function \(\psi \in L^\infty (\mathbb {T}^N)\) such that \(J^{\#}C=M_{\psi }\). Hence \(C=J^\#M_{\psi }=M_{\psi ^{\#}}J^\#\). By Proposition 2.2, the function \(\psi \) is unimodular and symmetric. \(\square \)

5 Invariant Subspaces of \(H^2(\mathbb {D}^N)\) and \(M_z\)–Conjugations

In this section, we study the image of a given invariant subspace of the Hardy space under \(M_z\)–conjugations.

Theorem 5.1

Let \(\mathcal {M}\subset H^2(\mathbb {D}^N)\) be a non–zero invariant subspace. Then there does not exist \(M_z\)–conjugation C such that \(C\mathcal {M}\subset H^2(\mathbb {D}^N)\).

Proof

Let C be an \(M_z\)–conjugation such that \(C\mathcal {M}\subset H^2(\mathbb {D}^N)\). Let \(n_1,\dots ,n_N\) be non–negative integers. Since \(z_i\mathcal {M}\subset \mathcal {M}\), we get

$$\begin{aligned} z_1^{n_1}\cdots z_N^{n_N} \mathcal {M}\subset \mathcal {M}. \end{aligned}$$

From the fact that C is an \(M_z\)–conjugation, we deduce

$$\begin{aligned} \bar{z}_1^{n_1}\cdots \bar{z}_N^{n_N}\,C\mathcal {M}=C\, \big (z_1^{n_1}\cdots z_N^{n_N}\mathcal {M}\big ) \subset C\mathcal {M}\subset H^2(\mathbb {D}^N). \end{aligned}$$

Finally,

$$\begin{aligned} C\mathcal {M}\subset z_1^{n_1}\cdots z_N^{n_N} H^2(\mathbb {D}^N). \end{aligned}$$

Therefore,

$$\begin{aligned} C\mathcal {M}\subset \bigcap _{(n_1,\dots , n_N)\in \mathbb {N}^N} z_1^{n_1}\cdots z_N^{n_N} H^2(\mathbb {D}^N)=\{0\} \end{aligned}$$

and as a consequence \(\mathcal {M}=\{0\}\). \(\square \)

Corollary 5.2

Let \(\mathcal {M}\subset H^2(\mathbb {D}^N)\) be a non–zero invariant subspace. Then there does not exist \(M_z\)–conjugation C such that \(C\mathcal {M}\subset \mathcal {M}\).

Corollary 5.3

There are no \(M_z\)–conjugations on \(L^2(\mathbb {T}^N)\) which preserve \(H^2(\mathbb {D}^N)\).

6 Invariant Subspaces of Type \(\alpha H^2(\mathbb {D}^N)\) and \(M_z\)–Commuting Conjugations

Let \(\alpha \in H^2(\mathbb {D}^N)\) be an inner function. Then \(\alpha H^2(\mathbb {D}^N)\) is invariant. We will now investigate conjugations in \(L^2(\mathbb {T}^N)\) which preserve subspaces of this form.

Let \(\alpha \), \(\beta \) be two inner functions. Since \(H^2(\mathbb {D}^N)\) is invariant for \(J^\#\), the operator

$$M_{\beta }J^{\#}M_{\bar{\alpha }}=M_{\beta \overline{\alpha ^\#}}J^\#\ :\ L^2(\mathbb {T}^N)\rightarrow L^2(\mathbb {T}^N)$$

is an antilinear isometry which maps \(\alpha H^2(\mathbb {D}^N)\) onto \(\beta H^2(\mathbb {D}^N)\) and commutes with \(M_{z_i},\ 1,\dots ,N\). However, this operator does not have to be an involution. By Proposition 2.2, the necessary and sufficient condition is that \( \beta \overline{\alpha ^\#}\) is symmetric (or equivalently \(\alpha \alpha ^{\#}=\beta \beta ^{\#}\)).

The theorem below characterizes all \(M_z\)–commuting conjugations mapping one invariant subspace of the considered type into another.

Theorem 6.1

Let \(\theta \), \(\alpha \in H^2(\mathbb {D}^N)\) be two inner functions. The following conditions are equivalent:

  1. (a)

    there exists an \(M_z-\)commuting conjugation C such that

    $$\begin{aligned} C(\alpha H^2(\mathbb {D}^N))\subset \theta H^2(\mathbb {D}^N),\end{aligned}$$
    (6.1)
  2. (b)

    there exists an inner function \(\beta \in H^2(\mathbb {D}^N)\) such that \(\theta \leqslant \beta \), \(\alpha \alpha ^\# = \beta \beta ^\#\).

Moreover, every conjugation C fullfiling condition (a) has a form \(C = M_{\overline{\alpha ^\#} \beta } J^\#\) for some \(\beta \in H^2(\mathbb {D}^N)\) such that \(\theta \leqslant \beta \), \(\alpha \alpha ^\# = \beta \beta ^\#\), and \(C(\alpha H^2(\mathbb {D}^N))= \beta H^2(\mathbb {D}^N)\).

Proof

Assume firstly that C is an \(M_z\)–commuting conjugation satisfying (6.1). By Theorem 4.1, \(C=M_{\psi }J^{\#}\) for some unimodular symmetric function \(\psi \in L^{\infty }(\mathbb {T}^N)\). In particular,

$$\begin{aligned} \theta H^2(\mathbb {D}^N)\ni C(\alpha ) =M_{\psi }J^{\#}\,\alpha =\psi \alpha ^{\#}, \end{aligned}$$

and there exists \(u\in H^2(\mathbb {D}^N)\) such that \(\psi \alpha ^{\#}=\theta u\). Note that u must be inner and so \(\psi =\beta \overline{\alpha ^{\#}}\) with \(\beta =\theta u\), \(\theta \leqslant \beta \). Hence \(C = M_{\overline{\alpha ^\#} \beta } J^\#\). Moreover, \(\beta \overline{\alpha ^{\#}}\) is symmetric, i.e., \(\beta \beta ^{\#}=\alpha \alpha ^{\#}\).

Now, let us assume that (b) holds. We define

$$\begin{aligned} C=M_{\overline{\alpha ^\#} \beta } J^\#. \end{aligned}$$

The condition \(\beta \beta ^{\#}=\alpha \alpha ^{\#}\) guarantees that C is a conjugation. Then, for any \(f\in H^2(\mathbb {D}^N)\), since \(\theta \leqslant \beta \), we have

$$\begin{aligned} C(\alpha f)=M_{\beta \overline{\alpha ^{\#}}}J^{\#}(\alpha f)=M_{\beta \overline{\alpha ^{\#}}}(\alpha ^{\#}f^\#)=\beta f^\#\in \beta H^2(\mathbb {D}^N)\subset \theta H^2(\mathbb {D}^N). \end{aligned}$$

\(\square \)

The corollary below characterizes all possible subspaces of \(H^2(\mathbb {D}^N)\) which are images of \(\alpha H^2(\mathbb {D}^N)\) under some \(M_z\)–commuting conjugation. We can take \(\theta =1\) in Theorem 6.1 (b) to get the following.

Corollary 6.2

Let \(\alpha \in H^2(\mathbb {D}^N)\) be an inner function and let \(\mathcal {M}\subset H^2(\mathbb {D}^N)\) be a closed subspace. Then the following conditions are equivalent:

  1. (a)

    there exists an \(M_z\)–commuting conjugation C such that \(C(\alpha H^2(\mathbb {D}^N))=\mathcal {M}\),

  2. (b)

    there exists an inner function \(\beta \in H^2(\mathbb {D}^N)\) such that \(\beta \beta ^{\#}=\alpha \alpha ^{\#}\) and \(\mathcal {M}= \beta H^2(\mathbb {D}^N)\)

The another consequence of Theorem 6.1 is the following corollary.

Corollary 6.3

Let \(\alpha \) be an inner function and let C be an \(M_z\)–commuting conjugation in \(L^2(\mathbb {T}^N)\). Then:

  1. (a)

    \(C(\alpha H^2(\mathbb {D}^N))\subset \alpha H^2(\mathbb {D}^N)\) if and only if \(C= M_{\alpha \overline{\alpha ^\#}}J^{\#}\),

  2. (b)

    \(C(\alpha H^2(\mathbb {D}^N))\subset \alpha ^{\#} H^2(\mathbb {D}^N)\) if and only if \(C=J^{\#}\).

Moreover, if \(\alpha \) is self-reflected, then \(C(\alpha H^2(\mathbb {D}^N))\subset \alpha H^2(\mathbb {D}^N)\) if and only if \(C= J^{\#}\).

Proof

By Theorem 6.1, \(C(\alpha H^2(\mathbb {D}^N))\subset \alpha H^2(\mathbb {D}^N)\) if and only if there exists an inner function \(\beta \) such that \(\alpha \leqslant \beta \) and \(\beta \beta ^{\#}=\alpha {\alpha ^{\#}}\). Thus \(\beta =u\alpha \) for some inner function u and \(u\alpha u^{\#}\alpha ^{\#}=\alpha {\alpha ^{\#}}\). Hence \(uu^{\#}=1\). This is only possible if \(u=\lambda \in \mathbb {T}\) since \(u\in H^2(\mathbb {D}^N)\). This gives (a). The proof of (b) is similar. \(\square \)

In particular, we have the following fact.

Corollary 6.4

Let C be an \(M_z\)–commuting conjugation on \(L^2(\mathbb {T}^N)\). Then

$$\begin{aligned} C(H^2(\mathbb {D}^N))\subset H^2(\mathbb {D}^N)\text { if and only if } C=J^{\#}. \end{aligned}$$

Remark 6.5

Let \(\alpha \) and \(\theta \) be two inner functions satisfying condition \(\theta H^2(\mathbb {D}^N)\varsubsetneq \alpha H^2(\mathbb {D}^N).\) Then there does not exist \(M_z\)–commuting conjugation C such that

$$\begin{aligned} C(\alpha H^2(\mathbb {D}^N))\subset \theta H^2(\mathbb {D}^N). \end{aligned}$$

Indeed, if so,

$$\begin{aligned} C(\alpha H^2(\mathbb {D}^N))\subset \theta H^2(\mathbb {D}^N)\subset \alpha H^2(\mathbb {D}). \end{aligned}$$

Thus, by Corollary 6.3(a), \(C= M_{\alpha \overline{\alpha ^\#}}J^{\#}\) and \(\theta H^2(\mathbb {D}^N)= \alpha H^2(\mathbb {D})\). This is a contradiction. Compare this remark with Example 7.5.

Notice that, for given inner functions \(\alpha \), \(\theta \), the number of existing \(M_z\)–commuting conjugations fulfilling (6.1) can be different. We provide three examples. First of them is based on [11, Theorem 5.5.2.].

Example 6.6

Let \(c\in (0,\frac{1}{8})\) and let \(\alpha \in H^2(\mathbb {T}^2)\) be an inner function defined by

$$\begin{aligned} \alpha (z_1,z_2)=\frac{2z_1^2z_2-z_1+c}{2-z_1z_2+cz_1^2z_2}. \end{aligned}$$

Then \(\alpha ^{\#}=\alpha \). Suppose that C is an \(M_z\)–commuting conjugation such that

$$\begin{aligned} C(\alpha H^2(\mathbb {T}^2))\subset \theta H^2(\mathbb {T}^2)\end{aligned}$$
(6.2)

for some inner function \(\theta \in H^2(\mathbb {T}^2)\). By Theorem 6.1, there is an inner function \(\beta \) such that \(\theta \leqslant \beta \) and \(\alpha ^2=\beta \beta ^{\#}\). Note that \(\alpha \) is continuous on \(\overline{\mathbb {D}^2}\). Then, by [12, Theorem 2.3.], \(\alpha ^2\) has the unique Rudin-Ahern factorization. Since \(\alpha \) is irreducible and \(\alpha ^2=\beta \beta ^{\#}\), \(\beta =\beta ^\#=\alpha \). Hence there is only one possible conjugation C fulfilling (6.2), namely \(C=M_{\alpha \overline{\alpha ^\#}}J^\#=J^\#\). On the other hand, there are only two functions \(\theta \) (\(\theta =1\) or \(\theta =\alpha \)) which can satisfy (6.2), since \(\theta \leqslant \beta =\alpha \).

Example 6.7

Let \(\alpha _1\) and \(\alpha _2\) be two different inner functions such that

  • \(\alpha _1\) and \(\alpha _2\) are irreducible,

  • \(\alpha _j^\#\ne \alpha _k\) for \(j,k\in \{1,2\}, \)

  • both \(\alpha _1(z_1,\cdot )\) and \(\alpha _2(z_1, \cdot )\) are rational for \(z_1\) in some subset of positive measure in \(\mathbb {T}\).Footnote 2

We can consider \(\alpha _j(z_1,z_2)=\frac{z_1z_2-\lambda _j}{1-\overline{\lambda _j}z_1z_2}\) for \(j=1,2\), where \(\lambda _1,\lambda _2\in \mathbb {D}\,\setminus \,\mathbb {R}\) and \(\lambda _1\ne \lambda _2\), \(\lambda _1\ne \bar{\lambda }_2\). These functions satisfy the above three condition. The conditions will allow us to apply [12, Theorem 2.3].

Now, let us take \(\alpha =\alpha _1\alpha _2\) and \(\theta =\alpha ^\#_2\). Note that \(\alpha H^2(\mathbb {D})\not \subset \theta H^2(\mathbb {D})\). We would like to find all \(M_z\)–commuting conjugations such that

$$\begin{aligned} C(\alpha H^2(\mathbb {D}))\subset \theta H^2(\mathbb {D}).\end{aligned}$$
(6.3)

Realize that, by the uniqueness of the Rudin-Ahern factorization, there are exactly two inner function satisfying condition (b) in Theorem 6.1: \(\beta _1=\alpha _1\alpha ^\#_2\) and \(\beta _2=\alpha ^\#_1\alpha ^\#_2\). Applying this theorem we obtain only two conjugations \(C_1=M_{\alpha _1\overline{\alpha ^\#_1}} J^\#\) and \(C_2=J^\#\).

Example 6.8

Consider \(\alpha \) as above and \(\theta =1\), i.e., condition \(C(\alpha H^2(\mathbb {D}))\subset H^2(\mathbb {D})\). Then there are exactly four inner functions satisfying condition (b) in Theorem 6.1: \(\beta _1=\alpha _1\alpha _2\), \(\beta _2=\alpha _1^\#\alpha _2\), \(\beta _3=\alpha _1\alpha _2^\#\), and \(\beta _4=\alpha _1^\#\alpha _2^\#\). Hence, we obtain four possible conjugations: \(C_1=M_{\alpha _1\alpha _2\overline{\alpha _1^\#}\overline{\alpha _2^\#}}J^\#\), \(C_2=M_{\alpha _2\overline{\alpha _2^\#}}J^\#\), \(C_3=M_{\alpha _1\overline{\alpha _1^\#}}J^\#\), and \(C_4=J^\#\).

7 Invariant Subspaces With Condition (*) and \(M_z\)–Commuting Conjugations

The aim of this section is to study different invariant subspaces fulfilling condition (*) introduced in [6]. This class of subspaces is completely different from those one considered in Sect. 6 if \(N>1\).

Following [11, Chapter 2], we introduce some function theory notation. We define multivariable Poisson kernel by the formula

$$\begin{aligned} P(z,w) = \prod _{i=1}^N P_{r_i}(\theta _i-\varphi _i), \quad z_j=r_je^{i\theta _j}, \quad w_j=e^{i\varphi _j}, \end{aligned}$$

where \(P_r\) denotes one dimensional Poisson kernel

$$\begin{aligned} P_r(\theta )= \frac{1-r^2}{1-2r \cos \theta +r^2}. \end{aligned}$$

For a complex Borel measure \(\mu \) on \(\mathbb {T}^N\) we define its Poisson integral by the formula

$$\begin{aligned} P[\mu ](z) = \int _{\mathbb {T}^N} P(z,w) d \mu (w), \quad z \in \mathbb {D}^N. \end{aligned}$$

Let \(f\in H^2(\mathbb {D}^N)\). Then there is a real singular measure \(\sigma _f\) on \(\mathbb {T}^N\) such that the least harmonic majorant of function \(\log |f|\), denoted by \(u(\log |f|)\), is given by

$$\begin{aligned} u(\log |f|)(z)= P[\log |f|dm_N+d\sigma _f](z), \quad z \in \mathbb {D}^N. \end{aligned}$$
(7.1)

It is known that \(d\sigma _f\leqslant 0\) for every \(f\in H^2(\mathbb {D}^N)\).

For the invariant subspace \(\mathcal {M}\), following [6], we define the zero set

$$\begin{aligned} Z(\mathcal {M})=\{z\in \mathbb {D}^N:f(z)=0 \text { for all } f\in \mathcal {M}\} \end{aligned}$$

and the measure

$$\begin{aligned} Z_{\partial }(\mathcal {M})=\inf \{-d\sigma _f:f\in \mathcal {M}, f\not =0\}. \end{aligned}$$

The invariant subspace \(\mathcal {M}\) is said to satisfy condition (*) if \(Z_\partial (\mathcal {M})=0\) and the real \((2N-2)\)–dimensional Hausdorff measure of \(Z(\mathcal {M})\) is 0.

Proposition 7.1

Let \(\mathcal {M}\subset H^2(\mathbb {D}^N)\) be an invariant subspace. If \(\mathcal {M}\) satisfies condition (*), then \(\mathcal {M}^\#\) satisfies condition (*).

Proof

Let \(\mu \) be a measure on \(\mathbb {T}^N\) and let \(\mu ^c\) be defined as follows

$$\begin{aligned} \mu ^c(\omega )=\mu (\{\bar{w}:w\in \omega \}). \end{aligned}$$

Notice that \(m_N^c=m_N\). Moreover,

$$\begin{aligned} h_\#^{(r)}(z)&= \int _{\mathbb {T}^N} P\Big (\frac{z}{r}, w\Big ) \log |f^\#(rw)| d m_N(w) \\&=\int _{\mathbb {T}^N} P\Big (\frac{z}{r}, w\Big ) \log |f(r\bar{w})| d m_N(w) \\&=\int _{\mathbb {T}^N} P\Big (\frac{z}{r}, \bar{w}\Big ) \log |f(rw)| d m_N(w) \\&=\int _{\mathbb {T}^N} P\Big (\frac{\bar{z}}{ r}, w \Big ) \log |f(rw)| d m_N(w)=h^{(r)}(\bar{z}). \end{aligned}$$

From [11, Theorem 3.2.4] we get the connection between least N–harmonic majorants

$$u(\log |f|)(z) = \lim \limits _{r\rightarrow 1^-} h^{(r)}(z)= \lim \limits _{r\rightarrow 1^-} h_\#^{(r)}(\bar{z}) = u(\log |f^\#|)(\bar{z}).$$

Hence by (7.1)

$$\begin{aligned} \int _{\mathbb {T}^N} P(z,w)d\sigma _{f^\#}(w)= \int _{\mathbb {T}^N} P(z,w)d\sigma ^c_{f}(w), \quad z \in \mathbb {D}^N. \end{aligned}$$

Consequently, by uniqueness of the measure, \(\sigma _{f^\#}=\sigma _f^c\). Now, one can easily see that \(Z_{\partial }(\mathcal {M})=Z_{\partial }(\mathcal {M}^\#)\). The easy observation \(Z(\mathcal {M}^\#)=\overline{Z(\mathcal {M})}\) finishes the proof. \(\square \)

Theorem 7.2

Let C be an \(M_z\)–commuting conjugation on \(L^2(\mathbb {T}^N)\). Let \(\mathcal {M}\subset H^2(\mathbb {D}^N)\) be an invariant subspace fulfilling condition (*). If \(C\mathcal {M}\subset H^2(\mathbb {D}^N)\) then there is a symmetric inner function \(\psi \) such that \(C=M_\psi J^\#\) and \(C\mathcal {M}=\psi \mathcal {M}^\#\).

Proof

Since C is \(M_z\)–commuting conjugation thus by Theorem 4.1 there is a unimodular symmetric function \(\psi \in L^\infty (\mathbb {T}^N)\) such that \(C=M_\psi J^\#=J^\# \psi ^\#\). Thus

$$\begin{aligned} \psi ^\#\mathcal {M}=J^\# J^\# {\psi }^\#\mathcal {M}=J^\#C\mathcal {M}\subset J^\#H^2(\mathbb {D}^N)=H^2(\mathbb {D}^N). \end{aligned}$$

The subspace \(\mathcal {M}\) is invariant subspace fulfilling condition (*). Moreover \(\psi ^\#\mathcal {M}\subset H^2(\mathbb {D}^N)\). Hence by [6, Theorem 1] function \(\psi ^\# \) is holomorphic and in particular inner. Therefore \(\psi \) is a symmetric inner function. \(\square \)

Notice that the above theorem can also be proved using Proposition 7.1.

Now, for any \(\alpha \in \mathbb {D}^N\), let \(H^2_\alpha \) be defined as

$$\begin{aligned} H^2_\alpha =\{f\in H^2(\mathbb {D}^N):f(\alpha )=0\}. \end{aligned}$$

In [13, Example 2.9] it was shown that \(H^2_\alpha \) is invariant and fulfills property (*).

As a consequence of Theorem 7.2 and the fact that \(J^\# H^2_\alpha =H^2_{\bar{\alpha }}\) for any \(\alpha \in \mathbb {D}^N\), we get the following.

Theorem 7.3

Let C be an \(M_z\)–commuting conjugation in \(L^2(\mathbb {T}^N)\). Let \(\alpha \in \mathbb {D}^N\). If \(C H^2_\alpha \subset H^2(\mathbb {D}^N)\) then there is an inner symmetric function \(\psi \) such that \(C=M_\psi J^\#\) and \(C H^2_\alpha =\psi H^2_{\bar{\alpha }}\).

To illustrate the above theorem we give two examples, showing that there exists a lot of conjugations leaving invariant space \(H^2_\alpha \).

Example 7.4

Let \(\alpha \in \mathbb {D}^N\cap \mathbb {R}^N\). Then there are infinitely many conjugations C on \(L^2(\mathbb {T}^N)\) such that \(C H^2_\alpha \subset H^2_\alpha \). In fact, there is a one-to-one correspondence between such conjugations and the set of all inner symmetric functions.

Example 7.5

Let \(\alpha =(\alpha _1,\ldots ,\alpha _N)\in \mathbb {D}^N\, \setminus \,\mathbb {R}^N\). Then there are infinitely many conjugations C on \(L^2(\mathbb {T}^N)\) such that \(C H^2_\alpha \subset H^2_\alpha \). If \(\psi \) is any symmetric inner function such that \(\psi (\alpha )=0\), then \(C=M_\psi J^\#\) is a conjugation such that \(C H^2_\alpha \subset H^2_\alpha \). It is clear that, in this case, the inclusion is strict. Compare this example with Remark 6.5.

To give a concrete example of \(\psi \), denote by \(B_\alpha \) the Blaschke product, i.e.,

$$\begin{aligned} B_\alpha (z)=\prod _{j=1}^N\frac{z_j-\alpha _j}{1-\bar{\alpha _j}z_j} \end{aligned}$$

for \(z=(z_1,\ldots ,z_N)\in \mathbb {D}^N\). Now, take a function \(\psi \) defined by \(\psi (z)=B_\alpha (z)B_{\overline{\alpha }}(z)\varphi \), \(z\in \mathbb {D}^N\), where \(\varphi \) is a symmetric inner function (e.g. \(\varphi \equiv 1\)). It is easily seen that \(\psi \) is symmetric and \(C=M_\psi J^\#\) satisfies \(CH^2_\alpha \subset H^2_\alpha \).