1 Introduction

Let \(\mathbb D\) be the open unit disk in the complex plane. Let \(L^2\) denote the Lebesgue space of square integrable functions on the unit circle \(\partial \mathbb D\). The Hardy space \(H^2\) is the subspace of analytic functions on \(\mathbb D\) whose Taylor coefficients are square summable. Then it can also be identified with the subspace of \(L^2\) of functions whose negative Fourier coefficients vanish. Let P and \(P^{\perp }\) be the orthogonal projections from \(L^2\) to \(H^2\) and \([H^2]^\perp \), respectively. Here \([H^2]^{\perp }\) is the orthogonal complement of \(H^2\) in \(L^2\). For \(f\in L^\infty \), the space of essentially bounded Lebesgue measurable functions on \(\partial \mathbb D\), the Toeplitz operator \(T_f\) with symbol \(f\in L^\infty \) is defined by

$$\begin{aligned} T_fh=P(fh), \end{aligned}$$

for \(h\in H^2\).

An analytic function \(\theta \) is called an inner function if \(|\theta |=1\) a.e. on \(\mathbb T\). For each non-constant inner function \(\theta \), the so-called model space is

$$\begin{aligned} K_\theta =H^2\ominus \theta H^2. \end{aligned}$$

It is a reproducing kernel Hilbert space with reproducing kernels

$$\begin{aligned} k_w^{\theta }(z)=\frac{1-\overline{\theta (w)}\theta (z)}{1-\bar{w}z}. \end{aligned}$$

Let \(P_\theta \) denote the orthogonal projection from \(L^2\) onto \(K_\theta \),

$$\begin{aligned} P_{\theta }f=P f-\theta P(\bar{\theta }f). \end{aligned}$$
(1.1)

For \(\varphi \in L^2\), the truncated Toeplitz operator \(A_\phi \) is defined by

$$\begin{aligned} A^{\theta }_\varphi f=P_\theta (\varphi f), \end{aligned}$$

on the dense subset \(K_\theta \cap H^{\infty }\) of \(K_\theta \). In particular, \(K_\theta \cap H^{\infty }\) contains all reproducing kernels \(k_w^{\theta }\). The operator \(A^{\theta }_\varphi \) may be extended to a bounded operator on \(K_\theta \) even for unbounded symbols \(\varphi \). The symbol \(\varphi \) is never unique and it is proved in [2] that

$$\begin{aligned} A^{\theta }_\varphi =0 \end{aligned}$$

if and only if

$$\begin{aligned} \varphi \in \theta H^2+\overline{\theta H^2}. \end{aligned}$$

If \(\theta (0)=0\), then \(A^{\theta }_\varphi \) has a unique symbol

$$\begin{aligned} \varphi \in K_\theta +\overline{K_\theta }. \end{aligned}$$

The set of all bounded truncated Toeplitz operators is denoted by \({\mathcal T}_\theta \).

Recall that a bounded operator T on a Hilbert space \({\mathcal H}\) is normal if \(T^*T=TT^*.\) The characterization of normal truncated Toepltiz operators is first given by Chalendar and Timotin using the algebraic properties of truncated Toeplitz operators obtained by Sarason [2] and Sedlock [3].

Theorem 1.1

[1, Theorem 6.2] Let \(\theta \) be a non-constant inner function vanishing at 0. Then \(A^\theta _\varphi \) is normal if and only if one of the following holds

  1. (1)

    \(A^\theta _\varphi \) belongs to \({\mathscr {B}}_\theta ^\alpha \), for some unimodular constant \(\alpha \).

  2. (2)

    \(A^\theta _\varphi \) is a linear combination of a self-adjoint truncated Toeplitz operator and the identity.

Here \({\mathscr {B}}_\theta ^\alpha \) is a class of truncated Toeplitz operators introduced in [3]. In this note, we give an elementary proof of their result.

2 Proof of the Main Result

In this section we offer a proof of our characterization of normal truncated Toepltiz operators \(A^{\theta }_{\varphi }\). We begin with some reduction. Notice that for any constant C, \(A^{\theta }_{\varphi +C}=A^{\theta }_{\varphi }+CI, \) which implies \(A^{\theta }_{\varphi }\) is normal if and only if \(A^{\theta }_{\varphi +C}\) is normal. Thus we may assume, without losing of generality, that \(\varphi (0)=0\).

For \(a\in \mathbb D\), let \(u_a\) be the Möbius transform

$$\begin{aligned} u_a(z)=\frac{z-a}{1-\bar{a}z}. \end{aligned}$$

The Crofoot transform is the unitary operator \(J: K_\theta \rightarrow K_{ u_a\circ \theta }\) defined by

$$\begin{aligned} J(f)=\frac{\sqrt{1-|a|^2}}{1-\bar{a}\theta }f. \end{aligned}$$

It is proved in [2] that

$$\begin{aligned} J{\mathcal T}_\theta J^*={\mathcal T}_{u_a\circ \theta }. \end{aligned}$$

Taking \(a=\theta (0)\), we see that it is sufficient to consider the normal truncated Toeplitz operators for \(\theta (0)=0\). In this case, constant functions are in \(K_{\theta }\). Write \(\varphi =\varphi _1+\overline{\varphi _2}\), where \(\varphi _1, \varphi _2\) are in \(K_{\theta }\). We may also assume \(\varphi _1(0)=\varphi _2(0)=0\).

It is easy to see that

$$\begin{aligned} (A^{\theta }_{\varphi })^*=A^{\theta }_{\bar{\varphi }}. \end{aligned}$$

Our approach to characterizing normal truncated Toeplitz operators starts with a computation of

$$\begin{aligned} ||A^{\theta }_{\varphi }u||^2-||(A^{\theta }_{\varphi })^*u||^2. \end{aligned}$$

Lemma 2.1

Let \(\theta \) be a non-constant inner function. Suppose

$$\begin{aligned} \varphi =\varphi _1+\overline{\varphi _2}, \end{aligned}$$

where \(\varphi _1, \varphi _2\) are in \(K_{\theta }\). Then for every \(u\in K_\theta \cap H^{\infty }\),

$$\begin{aligned}&||A^{\theta }_{\varphi }u||^2-||(A^{\theta }_{\varphi })^*u||^2\\&\quad =||P^{\perp }(\bar{\theta }\varphi _1 u)||^2-||P(\bar{\varphi _1}u) ||^2-( ||P^{\perp }(\bar{\theta }\varphi _2 u)||^2-||P(\bar{\varphi _2}u) ||^2). \end{aligned}$$

Proof

By (1.1), we have for every \(u\in K_\theta \cap H^\infty \)

$$\begin{aligned} A^{\theta }_{\varphi }u&=P_{\theta }(\varphi u)\\&=P(\varphi u)-\theta P(\bar{\theta }\varphi u)\\&=\varphi _1 u+P(\bar{\varphi _2}u)-\theta P(\bar{\theta }\varphi _1 u+\bar{\theta }\bar{\varphi _2} u )\\&=\varphi _1 u-\theta P(\bar{\theta }\varphi _1 u) +P(\bar{\varphi _2}u). \end{aligned}$$

Then

$$\begin{aligned} ||A^{\theta }_{\varphi }u||^2&=||(\varphi _1 u-\theta P(\bar{\theta }\varphi _1 u))+P(\bar{\varphi _2}u) ||^2\\&=||\varphi _1 u-\theta P(\bar{\theta }\varphi _1 u)||^2+||P(\bar{\varphi _2}u) ||^2+2\, \text{ Re } \langle \varphi _1 u-\theta P(\bar{\theta }\varphi _1 u), P(\bar{\varphi _2}u) \rangle \\&=||\bar{\theta }\varphi _1 u||^2-||P(\bar{\theta }\varphi _1 u)||^2+||P(\bar{\varphi _2}u) ||^2\\&\quad +\,2\, \text{ Re } \langle \varphi _1 u-\theta P(\bar{\theta }\varphi _1 u), P(\bar{\varphi _2}u) \rangle \\&=||P^{\perp }(\bar{\theta }\varphi _1 u)||^2+||P(\bar{\varphi _2}u) ||^2+2\, \text{ Re } \langle \varphi _1 u-\theta P(\bar{\theta }\varphi _1 u), P(\bar{\varphi _2}u)\rangle . \end{aligned}$$

And

$$\begin{aligned} \langle \varphi _1 u-\theta P(\bar{\theta }\varphi _1 u), P(\bar{\varphi _2}u) \rangle&=\langle \varphi _1 u, P(\bar{\varphi _2}u) \rangle -\langle \theta P(\bar{\theta }\varphi _1 u), P(\bar{\varphi _2}u) \rangle \\&=\langle \varphi _1 u, \bar{\varphi _2}u \rangle -\langle P(\bar{\theta }\varphi _1 u), \bar{\theta }P(\bar{\varphi _2}u) \rangle \\&=\langle \varphi _1 u, \bar{\varphi _2}u \rangle -\langle P(\bar{\theta }\varphi _1 u), \bar{\theta }u\bar{\varphi _2}-\bar{\theta }P^{\perp }(\bar{\varphi _2}u)\rangle \\&=\langle \varphi _1 u, \bar{\varphi _2} u \rangle . \end{aligned}$$

Thus

$$\begin{aligned} ||A^{\theta }_{\varphi }u||^2=||P^{\perp }(\bar{\theta }\varphi _1 u)||^2+||P(\bar{\varphi _2}u) ||^2+2\, \text{ Re } \langle \varphi _1 u, \bar{\varphi _2} u \rangle . \end{aligned}$$
(2.1)

Similarly

$$\begin{aligned} ||(A^{\theta }_{\varphi })^*u||^2=||A^{\theta }_{\varphi _2+\bar{\varphi _1}} u||^2=||P^{\perp }(\bar{\theta }\varphi _2 u)||^2+||P(\bar{\varphi _1}u) ||^2+2\, \text{ Re } \langle \varphi _2 u, \bar{\varphi _1}u\rangle . \end{aligned}$$
(2.2)

Subtracting (2.2) from (2.1), we get the desired identity. \(\square \)

For \(w\in \mathbb D\), let

$$\begin{aligned} k_w(z)=\frac{1}{1-\bar{w}z} \end{aligned}$$

be the reproducing kernel of \(H^2\).

First we show that if \(A^{\theta }_{\varphi }\) is normal then \(\varphi _1/\varphi _2\) is a unimodular function.

Lemma 2.2

Let \(\theta \) be a non-constant inner function vanishing at 0. Suppose \(\varphi =\varphi _1+\overline{\varphi _2}\), where \(\varphi _1, \varphi _2\) are in \(K_{\theta }\), and \(\varphi _1(0)=\varphi _2 (0)=0\). If \(A^{\theta }_{\varphi }\) is normal then

$$\begin{aligned} |\varphi _1|=|\varphi _2|, \end{aligned}$$

a.e. on \(\mathbb T\).

Proof

By Lemma 2.1, \(A^{\theta }_{\varphi }\) is normal implies

$$\begin{aligned} ||P^{\perp }(\bar{\theta }\varphi _1 u)||^2-||P(\bar{\varphi _1}u) ||^2=||P^{\perp }(\bar{\theta }\varphi _2 u)||^2-||P(\bar{\varphi _2}u) ||^2, \end{aligned}$$
(2.3)

for every \(u\in K_\theta \cap H^{\infty }\). Take \(u=1\), we get

$$\begin{aligned} ||P^{\perp }(\bar{\theta }\varphi _1)||^2-||P(\bar{\varphi _1}) ||^2= ||P^{\perp }(\bar{\theta }\varphi _1)||^2-||P(\bar{\varphi _2}) ||^2. \end{aligned}$$

Since

$$\begin{aligned} P^{\perp }(\bar{\theta }\varphi _j)=\bar{\theta }\varphi _j, \end{aligned}$$
(2.4)

and

$$\begin{aligned} P(\bar{\varphi _j})=0, \end{aligned}$$
(2.5)

we have

$$\begin{aligned} ||\varphi _1||=||\varphi _2||. \end{aligned}$$
(2.6)

Next we consider the reproducing kernels of \(K_\theta \):

$$\begin{aligned} k_w^{\theta }(z)=\frac{1-\overline{\theta (w)}\theta (z)}{1-\bar{w}z}, \end{aligned}$$

and take \(u=u_w =k_w^{\theta }+1\) in (2.3). Using (2.4) and (2.5), we have

$$\begin{aligned} ||P^{\perp }(\bar{\theta }\varphi _j u_w)||^2= & {} ||P^{\perp }(\bar{\theta }\varphi _j k_w^{\theta })||^2+||P^{\perp }(\bar{\theta }\varphi _j)||^2+2\text{ Re }\,\langle P^{\perp }(\bar{\theta }\varphi _j k_w^{\theta }), P^{\perp }(\bar{\theta }\varphi _j )\rangle \\= & {} ||P^{\perp }(\bar{\theta }\varphi _j k_w^{\theta })||^2+||\bar{\theta }\varphi _j||^2+2\text{ Re }\,\langle P^{\perp }(\bar{\theta }\varphi _j k_w^{\theta }), \bar{\theta }\varphi _j \rangle , \end{aligned}$$

and

$$\begin{aligned} ||P(\bar{\varphi _j}u_w)||^2= & {} ||P(\bar{\varphi _j}k_w^{\theta })||^2+||P(\bar{\varphi _j})||^2+2\text{ Re }\,\langle P(\bar{\varphi _j}k_w^{\theta }), P(\bar{\varphi _j})\rangle \\= & {} ||P(\bar{\varphi _j}k_w^{\theta })||^2. \end{aligned}$$

This together with Lemma 2.1 and (2.6) implies

$$\begin{aligned} \text{ Re }\,\langle P^{\perp }(\bar{\theta }\varphi _1 k_w^{\theta }), \bar{\theta }\varphi _1\rangle =\text{ Re }\,\langle P^{\perp }(\bar{\theta }\varphi _2 k_w^{\theta }), \bar{\theta }\varphi _2 \rangle . \end{aligned}$$
(2.7)

Since

$$\begin{aligned} k_w^{\theta }=(1-\overline{\theta (w)}\theta )k_w, \end{aligned}$$

we get

$$\begin{aligned} P^{\perp }(\bar{\theta }\varphi _j k_w^{\theta })= & {} P^{\perp }(\bar{\theta }\varphi _j(1-\overline{\theta (w)}\theta ) k_w)=P^{\perp }(\bar{\theta }\varphi _jk_w)-\overline{\theta (w)}P^{\perp }(\varphi _j k_w)\\= & {} P^{\perp }(\bar{\theta }\varphi _jk_w). \end{aligned}$$

Hence

$$\begin{aligned} \quad&\text{ Re }\,\langle P^{\perp }(\bar{\theta }\varphi _j k_w^{\theta }), \bar{\theta }\varphi _j\rangle \\&=\text{ Re }\,\langle P^{\perp }(\bar{\theta }\varphi _jk_w), \bar{\theta }\varphi _j\rangle \\&=\text{ Re }\,\langle \bar{\theta }\varphi _jk_w, \bar{\theta }\varphi _j\rangle \\&=\text{ Re }\,\int _{0}^{2\pi }\frac{ |\varphi _j(e^{it})|^2}{1-\bar{w}e^{it}}\frac{dt}{2\pi }\\&=\int _{0}^{2\pi } |\varphi _j(e^{it})|^2 \Big (\text{ Re }\,\frac{1}{1-\bar{w}e^{it}}\Big )\frac{dt}{2\pi }\\&=\frac{1}{2}\int _{0}^{2\pi } |\varphi _j(e^{it})|^2 \Big (1+\text{ Re }\,\frac{1+\bar{w}e^{it}}{1-\bar{w}e^{it}}\Big )\frac{dt}{2\pi }\\&=\frac{1}{2}||\varphi _j||^2+\frac{1}{2}\int _{0}^{2\pi } |\varphi _j(e^{it})|^2 \Big (\text{ Re }\,\frac{1+\bar{w}e^{it}}{1-\bar{w}e^{it}}\Big )\frac{dt}{2\pi }\\&=\frac{1}{2}(||\varphi _j||^2+\widehat{|\varphi _j|^2}(w)). \end{aligned}$$

The last equality holds because

$$\begin{aligned} \text{ Re }\,\frac{1+\bar{w}e^{it}}{1-\bar{w}e^{it}} \end{aligned}$$

is the Poisson kernel at w. Here \(\widehat{|\varphi _j|^2}\) is the harmonic extension of the function \(|\varphi _j|^2\). It follows from (2.7) and (2.6) that

$$\begin{aligned} \widehat{|\varphi _1|^2}(w)=\widehat{|\varphi _2|^2}(w). \end{aligned}$$

Let \(w\rightarrow \zeta \in \mathbb T\) nontangentially, we see that

$$\begin{aligned} |\varphi _1|=|\varphi _2|, \end{aligned}$$

a.e. on \(\mathbb T\). \(\square \)

Let U is the unitary operator on \(L^2\) defined by

$$\begin{aligned} Uh(z)=\bar{z}\tilde{h}(z), \end{aligned}$$

where \(\tilde{h}(z)=h(\bar{z})\). Let \(V_\theta \) be the operator

$$\begin{aligned} V_\theta h=P(\theta h), \end{aligned}$$

for \(h\in L^2\). Consider the decomposition

$$\begin{aligned}{}[H^2]^{\perp }=\bar{\theta }K_\theta \oplus \bar{\theta }[H^2]^{\perp }. \end{aligned}$$

It is easy to check that \(V_\theta \) maps \(\bar{\theta }K_\theta \) onto \(K_\theta \), and maps \(\bar{\theta }[H^2]^{\perp }\) to 0. Thus \(V_\theta \) maps \([H^2]^\perp \) onto \(K_\theta \). Since U maps \(H^2\) onto \([H^2]^\perp \), we see that

$$\begin{aligned} V_\theta U: H^2\rightarrow K_\theta \end{aligned}$$

is also onto.

We shall use the following identity.

Lemma 2.3

Let \(\theta \) be an inner function and let g be in \(H^2\). Then for every function \(f\in H^\infty \)

$$\begin{aligned} ||P(\bar{g}V_\theta U f)||=||P^\perp (\bar{\theta } g f^*)||, \end{aligned}$$

where \(f^*(z)=\overline{f(\bar{z})}\).

Proof

Notice that for all \(h\in L^2\), we have

$$\begin{aligned} (Uh)^*=U(h^*) \end{aligned}$$

and

$$\begin{aligned} (Ph)^*=P(h^*). \end{aligned}$$

Thus

$$\begin{aligned} P(\bar{g}V_\theta U f)&=P(\bar{g}P(\theta U f))=P(\bar{g}\theta U f)=P(\bar{z}\theta \bar{g}\tilde{f})\\&=PU((\bar{\theta }g)^*f)=P(U(\bar{\theta }g f^*))^*\\&=(PU(\bar{\theta }g f^*))^*=(UP^\perp (\bar{\theta }g f^*))^*. \end{aligned}$$

Here we used \(PU=UP^\perp \) in the last equality. Since \(||h||=||h^*||,\) for all \(h\in L^2\) and U is an isometry, we get the desired identity. \(\square \)

The following result is well-known (see e.g. [4, Lemma 8]).

Theorem 2.1

If \(f\in H^2\), then for every \(w\in \mathbb D\),

$$\begin{aligned} P(\bar{f}k_w) = \overline{f(w)} k_w. \end{aligned}$$

Now we can prove the main result.

Theorem 2.2

Let \(\theta \) be a non-constant inner function vanishing at 0. Suppose \(\varphi =\varphi _1+\overline{\varphi _2}\), where \(\varphi _1, \varphi _2\) are in \(K_{\theta }\). Then \(A^{\theta }_{\varphi }\) is normal if and only if either

$$\begin{aligned} \varphi _2-\varphi _2(0)=\alpha (\varphi _1-\varphi _1(0)) \end{aligned}$$

or

$$\begin{aligned} \varphi _2-\varphi _2(0)=\alpha \theta (\overline{\varphi _1}-\overline{\varphi _1(0)}), \end{aligned}$$

for some unimodular constant \(\alpha \).

Proof

We may assume \(\varphi _1(0)=\varphi _2(0)=0\). Sufficiency follows easily from Lemma 2.1.

Suppose \(A^{\theta }_{\varphi }\) is normal. By (2.3) and Lemma 2.2, we have

$$\begin{aligned} ||P(\bar{\varphi _1}u)||^2+||P(\bar{\theta }\varphi _1 u)||^2=||P(\bar{\varphi _2}u)||^2+||P(\bar{\theta }\varphi _2 u)||^2, \end{aligned}$$
(2.8)

for every \(u\in K_\theta \cap H^\infty \). According to the discussion before Lemma 2.3, if we write \(u=V_\theta Uf\), where \(f\in H^\infty \), (2.8) is equivalent to

$$\begin{aligned} ||P(\bar{\varphi _1}V_\theta Uf)||^2+||P(\bar{\theta }\varphi _1 V_\theta Uf)||^2=||P(\bar{\varphi _2}V_\theta Uf)||^2+||P(\bar{\theta }\varphi _2 V_\theta Uf)||^2, \end{aligned}$$

for every \(f\in H^\infty \). Using Lemma 2.3 and that \(f\mapsto f^*\) is a bijection on \(H^\infty \), we have

$$\begin{aligned} ||P^\perp (\bar{\theta }\varphi _1 f)||^2+||P^\perp (\bar{\varphi _1}f)||^2=||P^\perp (\bar{\theta }\varphi _2 f)||^2+||P^\perp (\bar{\varphi _2}f)||^2, \end{aligned}$$
(2.9)

for every \(f\in H^\infty \). By Lemma 2.2,

$$\begin{aligned} ||\bar{\theta }\varphi _1 f||=||\bar{\theta }\varphi _2 f||, \end{aligned}$$

and

$$\begin{aligned} ||\bar{\varphi _1}f||=||\bar{\varphi _2}f||. \end{aligned}$$

We see that (2.9) implies

$$\begin{aligned} ||P(\bar{\theta }\varphi _1 f)||^2+||P(\bar{\varphi _1}f)||^2=||P(\bar{\theta }\varphi _2 f)||^2+||P(\bar{\varphi _2}f)||^2, \end{aligned}$$
(2.10)

for every \(f\in H^\infty \).

Take \(f=k_w\) in (2.10). By Theorem 2.1, we get

$$\begin{aligned} |\varphi _1(w)|^2+|(\theta \bar{\varphi _1})(w)|^2=|\varphi _2(w)|^2+|(\theta \bar{\varphi _2})(w)|^2, \end{aligned}$$
(2.11)

for every \(w\in \mathbb D\). Here \((\theta \bar{\varphi _1})(w)\) means \(\langle \theta \bar{\varphi _1}, k_w\rangle \).

On the other hand, using Lemma 2.2, we have

$$\begin{aligned} {\varphi _1(w)}(\theta \bar{\varphi _1})(w)= & {} \langle \varphi _1(\theta \bar{\varphi _1}), k_w\rangle =\langle \theta |\varphi _1|^2, k_w\rangle =\langle \theta |\varphi _2|^2, k_w\rangle \nonumber \\= & {} \langle \varphi _2(\theta \bar{\varphi _2}), k_w\rangle ={\varphi _2(w)}(\theta \bar{\varphi _2})(w). \end{aligned}$$
(2.12)

for every \(w\in \mathbb D\).

Multiplying both sides of (2.11) by \(|\varphi _2(w)|^2\) and using (2.12), we have

$$\begin{aligned} |\varphi _1(w)\varphi _2(w)|^2+|\varphi _2(w)(\theta \bar{\varphi _1})(w)|^2&=|\varphi _2(w)|^4+|\varphi _2(w)(\theta \bar{\varphi _2})(w)|^2\\&=|\varphi _2(w)|^4+|\varphi _1(w)(\theta \bar{\varphi _1})(w)|^2, \end{aligned}$$

which is equivalent to

$$\begin{aligned} (|\varphi _1(w)|^2-|\varphi _2(w)|^2)(|(\theta \bar{\varphi _1})(w)|^2-|\varphi _2(w)|^2)=0. \end{aligned}$$

Thus for every \(w\in \mathbb D\), either

$$\begin{aligned} |\varphi _1(w)|=|\varphi _2(w)|, \end{aligned}$$

or

$$\begin{aligned} |\varphi _2(w)|=|(\theta \bar{\varphi _1})(w)|. \end{aligned}$$

Then it follows from the properties of analytic functions that either

$$\begin{aligned} \varphi _1=\alpha \varphi _2, \end{aligned}$$

or

$$\begin{aligned} \varphi _2=\alpha \theta \bar{\varphi _1}, \end{aligned}$$

for some unimodular constant \(\alpha \). \(\square \)

Remark 2.1

The characterization given in Theorem 2.2 is equivalent to that in Theorem 1.1. In fact, if we write \(\varphi =\varphi _1+\bar{\varphi _2}+\varphi (0)\), where \(\varphi _1, \varphi _2\) are in \(K_\theta \cap zH^2\), it is shown in [1, Section 5] that \(A^\theta _\varphi \in {\mathscr {B}}_\theta ^\alpha \) if and only if \(\theta \bar{\varphi _2}=\alpha \varphi _1\).