1 Introduction

In the proof of [2, Theorem 4.3 (iii)], the inclusion relationship “\(\{\gamma +b_\gamma :\ \gamma \in \Gamma \}\subset \sum \nolimits _{j=1}^{\ell _n}(\{0\}\cup U_j)\)" maybe wrong in some cases. Actually, this inclusion relationship need a precondition “\(\ell _j\le \ell _n\) for all \(j<\ell _n\)". The following example shows that, there are examples that \(\ell _j>\ell _n\) holds for at least one integer \(j<\ell _n\) for all \(n>0\). Hence, the sufficiency of [2, Theorem 1.1] needs to be reproved.

Example

Let \(p_{2n-1}=4,\ p_{2n}=9\) and \(d_{2n-1}=1,\ d_{2n}=8\) for all \(n\ge 1\). Then, the definition of \(k_n\) and \(\ell _n\) shows

$$\begin{aligned} k_{2n}=v_2(\frac{p_1p_2\cdots p_{2n}}{2d_{2n}})=v_2(\frac{36^n}{16})=2n-4, \quad \forall \ n\ge 1, \end{aligned}$$
$$\begin{aligned} k_{2n-1}=v_2(\frac{p_1p_2\cdots p_{2n-1}}{2d_{2n-1}})=v_2(\frac{36^{n-1}\times 4}{2})=2n-1, \quad \forall \ n\ge 1. \end{aligned}$$

Also, \(\ell _{2n-1}=2n+2\) and \(\ell _{2n}=2n\) for all \(n\ge 1\). This means \(\ell _{\ell _{n}-1}>\ell _{n}\) for all \(n\ge 1\).

We recall the definition of Moran-type Bernoulli convolution. Let \(\{p_n\}_{n\ge 1}\) and \(\{d_n\}_{n\ge 1}\) be two sequences of integers satisfying \(|p_n|\ge 2,\ |d_n|\ge 1\) and

$$\begin{aligned} \sum _{n=1}^{+\infty }|p_1^{-1}p_2^{-1}\cdots p_n^{-1}d_n|<+\infty . \end{aligned}$$

The weak limit of the following convolutions is called a Moran-type Bernoulli convolution

$$\begin{aligned} \mu _n=\delta _{p_1^{-1}D_1}*\delta _{p_1^{-1}p_2^{-1}D_2}*\cdots *\delta _{p_1^{-1}p_2^{-1}\cdots p_n^{-1}D_n}.\end{aligned}$$

And we denote it by

$$\begin{aligned} \mu =\delta _{p_1^{-1}D_1}*\delta _{p_1^{-1}p_2^{-1}D_2}*\cdots *\delta _{p_1^{-1}p_2^{-1}\cdots p_n^{-1}D_n}\cdots . \end{aligned}$$
(1.1)

We shall reprove the sufficiency of the following result (i.e. [2, Theorem 1.1]).

Theorem 1.1

For the measure \(\mu \) defined by (1.1) with \(|p_n|>|d_n|\) for all \(n\ge 2\), assume that the sequence \(\{|d_n|\}_{n=1}^{+\infty }\) is bounded. Then, \(\mu \) is a spectral measure if and only if \(k_j\ne k_i\) for all \(j>i\ge 1\), where

$$\begin{aligned} k_n=v_2\Big (\frac{p_1p_2\dots p_n}{2d_n}\Big )=v_2(p_1p_2\dots p_n)-v_2(2d_n),\quad n=1,2,3,\dots . \end{aligned}$$
(1.2)

2 Proof of the Sufficiency of Theorem 1.1

In order to make the proof more readable, we first simplify our model.

Proposition 2.1

For the measure \(\mu \) defined by (1.1), there exist two sequences of integers \(\{c_n\}_{n=1}^\infty \) and \(\{q_n\}_{n=1}^\infty \) such that for \(n\ge 1\), we have \( \gcd (q_n,c_n)=1 \) and

$$\begin{aligned} q_1^{-1}q_2^{-1}\cdots q_n^{-1}c_n=p_1^{-1}p_2^{-1}\cdots p_n^{-1}d_n. \end{aligned}$$
(2.1)

Furthermore, we have \(|q_n|>|c_n|\) when \(|p_n|>|d_n|\) (\(n=1,2,\dots \)). Hence, we can rewrite \(\mu \) as

$$\begin{aligned} \mu =\delta _{q_1^{-1}C_1}*\delta _{q_1^{-1}q_2^{-1}C_2}*\delta _{q_1^{-1}q_2^{-1}q_3^{-1}C_3}\cdots *\delta _{q_1^{-1}q_2^{-1}\cdots q_n^{-1}C_n}*\cdots , \end{aligned}$$
(2.2)

where \(C_n=\{0,c_n\}\).

Proof

Write \(g_0=1\), and define inductively

$$\begin{aligned} g_{n}=\gcd (|g_{n-1}p_{n}|,\ |d_{n}|),\ q_{n}=\frac{g_{n-1}p_{n}}{g_{n}}\ \text { and} \ c_{n}=\frac{d_{n}}{g_{n}},\quad \forall \ n\ge 1. \end{aligned}$$

It is clear that for any \(n\ge 1\), we have \(\gcd (q_n,\ c_n)=1\) and (2.1). By writing \(C_n=\{0,c_n\}\), we have

$$\begin{aligned} \delta _{q_1^{-1}q_2^{-1}\cdots q_n^{-1}C_n}=\delta _{p_1^{-1}p_2^{-1}\cdots p_n^{-1}D_n}, \end{aligned}$$

which implies that (2.2) holds.

If \(|p_n|>|d_n|\), noting \(|g_{n-1}p_{n}|\ge |p_n|\), it is obvious that \(|q_n|>|c_n|\). \(\square \)

The above Proposition 2.1 shows that, in order to prove the sufficiency of Theorem 1.1, without loss of generality, we can assume that \(\gcd (d_n,\ p_n)=1\). By the argument in [2], we will always assume that [2, (2.9)] holds without loss of generality. Therefore, we shall assume that the following conditions hold in the sequel:

$$\begin{aligned} p_n\ge 2,\ \ d_n\ge 1,\ \ \gcd (d_n,\ p_n)=1,\ \frac{p_1p_2\cdots p_n}{2d_n}\in {\mathbb {N}},\ \ \ \ \ \forall \ n\ge 1. \end{aligned}$$
(2.3)

The following Proposition 2.2 is obviously true.

Proposition 2.2

Let \(\nu \) be a probability measure and its support has finite cardinality N. If \(L^2(\nu )\) has an orthogonal set \(\{e^{2\pi i\lambda x}:\ \lambda \in \Lambda \}\) and \(\#\Lambda \) is at least N, then \(\Lambda \) is a spectrum of \(\nu \) and \(\#\Lambda =N\).

We will continue to use notations \(\ell _n,\ k_n,\ U_n,\ t_n,\ r_n\) defined in [2] and the constant c is defined in [2, Lemma 4.1 (i)]. Given a nonzero integer n, we denote by \(\theta (n)\) the odd part of n, i.e. \(\theta (n)=\frac{n}{2^{\nu _2(n)}}.\) Then, we rewrite

$$\begin{aligned} U_n=2^{k_n}\theta (p_1p_2\cdots p_{\ell _n})(2\mathbb {Z}+1), \quad n\ge 1. \end{aligned}$$
(2.4)

The following set will play an important role in the sequel

$$\begin{aligned} \mathcal {I}=\{n:\ell _{n}=n\}. \end{aligned}$$
(2.5)

Lemma 2.3

Assume that (2.3) holds and \(k_i\ne k_j\) for all \(i>j\ge 1\). Then, we have the following statements.

  1. (i).

    \(\#{\mathcal {I}}=+\infty \).

  2. (ii).

    Let B be a finite nonempty subset of positive integers. Then, \(\Delta :=\sum \limits _{j\in B}\{0,\ a_j\}\) is a spectrum (with cardinality \(2^{\#B}\)) of \(*_{j\in B} \delta _{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}\) for any \(a_j\in U_j\).

  3. (iii).

    Let B be a finite nonempty subset of positive integers. Assume that \(\Lambda \subset \sum \limits _{j\in B}(\{0\}\cup U_j)\) is a spectrum of the probability measure \(*_{j\in B} \delta _{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}\). For any \(m\ge \max \{\ell _j:\ j\in B\}\) and \(\lambda \in \Lambda \), we take an integer \(b_\lambda \in p_1p_2\cdots p_m{\mathbb {Z}}\) with \(b_0=0\). Then, the set \(\{\lambda +b_\lambda :\ \gamma \in \Lambda \}\) is also a spectrum of the probability measure \(*_{j\in B} \delta _{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}\). Furthermore, we have \(\{\gamma +b_\lambda :\ \gamma \in \Lambda \}\subset \sum \limits _{j\in B}(\{0\}\cup U_j)\).

Proof

(i) It is sufficient to prove that for any integer \(N>0\), there exists a positive integer \(n>N\) such that \(\ell _n=n\).

Indeed, since \(\{d_n\}_{n\ge 1}\) is bounded, there is an integer \(z_0\) such that \(k_n\ge z_0\) for all \(n>0\). Hence, there is an integer \(n> N\) such that \(k_n=\min \{k_j:\ j>N\}\). Since \(k_i\ne k_j\) for all \(i>j\ge 1\), we see that \(k_n<k_j\) for all \(j>n\). Hence, the definition of \(\ell _n\) shows \(\ell _n=n\). The conclusion is proven.

(ii) Suppose \(B=\{j_1,\ j_2,\ \cdots ,\ j_s\}\) with \(k_{j_1}<k_{j_2}<\cdots < k_{j_s}\). From [2, Lemma 4.2], it follows that \(k_{i}<k_{j}\) implies \(\ell _{i}\le \ell _{j}\). Then, the definition of \(U_n\) shows

$$\begin{aligned} U_{j_t}+\sum \limits _{t<i\le s}(\{0\}\cup U_{j_i})=U_{j_t},\quad t=1,2,\dots , s-1. \end{aligned}$$
(2.6)

For any \(\xi =\sum \limits _{j\in B}\xi _j\) and \( \eta =\sum \limits _{j\in B}\eta _j\in \Delta \) with \(\xi _j,\ \eta _j\in \{0,\ a_j\}\) and \(\xi _j\ne \eta _j\) for at least one \(j\in B\), it is easy to see \(\xi -\eta \in \sum \limits _{j\in B}\{0,\ \pm a_j\}\). Write \(t=\min \{i:\ \xi _{j_i}\ne \eta _{j_i},\ 1\le i\le s\}\). Then, we have \(\xi -\eta \in U_{j_t}+\sum \limits _{t<i\le s}(\{0\}\cup U_{j_i})\). From (2.6) it follows \(\xi -\eta \in U_{j_t}\). This implies \(\xi -\eta \in U_{j_t}\) and \(\#\Delta =2^s\). Furthermore, \(\xi -\eta \in U_{j_t}\) shows that \(\xi -\eta \) is a zero point of the Fourier transformation of the probability measure \(*_{j\in B} \delta _{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}\), i.e. \( \prod \limits _{j\in B}\widehat{\delta }_{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}(\xi -\eta )=0\).

It is easy to see that the support of the measure \(*_{j\in B} \delta _{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}\) has cardinality at most \(2^s\). Proposition 2.2 shows that \(\Delta \) is a spectrum of the measure \(*_{j\in B} \delta _{p_1^{-1}p_2^{-1}\cdots p_{j}^{-1}D_{j}}\) and \(\#\Delta =2^s=2^{\# B}\).

(iii) Note a fact that for all \(j\in B\), the integer \(p_1p_2\cdots p_{j}\) is a period of \(\widehat{\delta }_{p_1^{-1}\cdots p_j^{-1}D_j}\). For \(m\ge \max \{\ell _j:\ j\in B\}\ge j\), we have \(\widehat{\delta }_{p_1^{-1}\cdots p_j^{-1}D_j}(x+\lambda +b_\lambda )=\widehat{\delta }_{p_1^{-1}\cdots p_j^{-1}D_j}(x+\lambda )\) for all \(x\in {\mathbb {R}}\), \(j\in B\) and \(\lambda \in \Lambda \). Hence

$$\begin{aligned} \sum _{\gamma \in \{\lambda +b_\lambda :\lambda \in \Lambda \}} \left| \prod _{j\in B}\widehat{\delta }_{p_1^{-1}\cdots p_j^{-1}D_j}(x+\gamma )\right| ^2=\sum _{\lambda \in \Lambda } \left| \prod _{j\in B}\widehat{\delta }_{p_1^{-1}\cdots p_j^{-1}D_j}(x+\lambda )\right| ^2,\quad \forall \ x\in {\mathbb {R}}. \end{aligned}$$

That means that the first conclusion in (iii) is proven by using [2, Proposition 2.3].

For any \(\lambda \in \Lambda \), it can be written as \(\lambda =\sum \limits _{j\in B}b_j\) with \(b_j\in (\{0\}\cup U_j)\). Let \(B=\{j_1,\ j_2,\ \cdots ,\ j_s\}\) with \(k_{j_1}<k_{j_2}<\cdots < k_{j_s}\) as in (ii) and define \(t=\min \{i:\ b_{j_i}\ne 0\}\). Then, (2.6) shows \(\lambda \in U_{j_t}\). Since \(m\ge \max \{\ell _j:\ j\in B\}\), the definition of \(U_j\) implies that \(U_j+b_\lambda =U_j\) for all \(j\in B\), which implies \(\lambda +b_\lambda \in U_{j_t}\). Hence, we have \(\{\lambda +b_\lambda :\ \lambda \in \Lambda \}\subset \sum \limits _{j\in B}(\{0\}\cup U_j)\) for any \(b_\lambda \in p_1p_2\cdots p_m{\mathbb {Z}}\) with \(b_0=0\). The second conclusion in (iii) is proven. \(\square \)

The following two lemmas deal with the possible case that \(\ell _j>\ell _n\) for some \(j<\ell _n\).

Lemma 2.4

Assume that \(k_n\ne k_m\) for all \(n\ne m\) and (2.3) holds. Furthermore, assume that there exists a positive integer \(n_0\) such that for any \(n\ge n_0\) there exists an integer \(j_n<\ell _n\) satisfying \(\ell _{j_n}>\ell _n\).

  1. (i)

    . For any \(i\ge n_0\), there is at least one member of the group \(p_i,\ p_{i+1},\ \cdots ,\ p_{i+c}\) which is an odd integer larger than or equal to 3.

  2. (ii)

    . There exists a positive integer \(N_1\ge 0\) such that

    $$\begin{aligned} \frac{d_{n}}{\theta (p_{\ell _i+1}\cdots p_{n})}\le 1 ,\ n_0\le i\le n-N_1. \end{aligned}$$
    (2.7)
  3. (iii)

    . For any \(n\ge n_0+c\), we have \(v_2(p_n)<\max \{v_2(d_j):\ j>0\}\).

Proof

(i) Given \(i\ge n_0\), suppose \(p_i,\ p_{i+1},\ \cdots ,\ p_{i+c}\) are all even. From the assumption \(\gcd (p_n,d_n)=1\), it is clear that \(d_i,\ d_{i+1},\ \cdots ,\ d_{i+c}\) are all odd. Hence, \(k_i<k_{i+1}< \cdots <k_{i+c}\), which implies \(\ell _i=i\) since [2, Lemma 4.1 (i)] shows \(i\le \ell _i\le i+c\). On the other hand, however, our assumption shows for the integer i, there is a positive integer \(j<\ell _i\) such that \(\ell _j>\ell _i\). Then, we have \(k_{j}>k_{\ell _j}>k_i\). In virtue of \(\ell _i=i\), we have \(j<i\). Since \(p_i\) is even and \(d_i\) is odd, we have \(k_i=v_2(p_1p_2\cdots p_{i})-1>v_2(p_1p_2\cdots p_{j})-1\ge k_j\), which leads to a contradiction. Therefore, at least one member of \(p_i,\ p_{i+1},\ \cdots ,\ p_{i+c}\) is odd which is larger than or equal to 3. The conclusion (i) is proven.

(ii) Choose a positive integer \(s>0\) such that \(3^s\ge \max \{d_n:\ n>0\}\). Thus for \(i\ge n_0\), we have

$$\begin{aligned} \frac{\max \{d_n:\ n>0\}}{p_{\ell _i+1}p_{\ell _i+2}\cdots p_{\ell _i+sc}}\le 1. \end{aligned}$$

It is clear that we finish the proof by taking \(N_1=sc\).

(iii) Suppose \(v_2(p_n)\ge \max \{v_2(d_j):\ j>0\}\) for some \(n\ge n_0+c\). For any j and m with \(j<n\le m\), we have \(k_m= v_2(p_1p_2\cdots p_{m})-1-v_2(d_m)\ge v_2(p_1p_2\cdots p_{j})+v_2(p_n)-1-v_2(d_m)\ge v_2(p_1p_2\cdots p_{j})-1\ge k_j\). In fact, by the assumption that \(k_m\ne k_j\), we have \(k_m>k_j\). Let \(k_s=\max \{k_j:\ n_0\le j\le n-1\}\). According to the definition of \(\ell _s\), we have \(\ell _s=n-1\).

On the other hand, however, for the integer s, there exits a positive integer \(j_0\) with \(j_0<\ell _s\) such that \(\ell _{j_0}>\ell _s\), which implies \(\ell _{j_0}\ge n\). Noting that \(j_0<n\), according to the above argument, we get \(k_{\ell _{j_0}}>k_{j_0}\), which is a contradiction to the definition of \(\ell _{j_0}\). The statement (iii) is proven. \(\square \)

Lemma 2.5

Assume that \(k_n\ne k_m\) for all \(n\ne m\) and (2.3) hold. Furthermore, assume that there exists a positive integer \(n_0\) such that for any \(n\ge n_0\) there exists an integer \(j_n<\ell _n\) satisfying \(\ell _{j_n}>\ell _n\). Then, there are small constants \(\varepsilon >0\) and \(\theta _0>0\) such that for any \(n_1\) and \(n_2\in \mathcal {I}\) with \(n_2>n_1+N_1\), there exists a spectrum \(\Lambda =\sum _{i=n_1+1}^{n_2}\{0,\ a_i\}\) of \(*_{i=n_1+1}^{n_2} \delta _{p_1^{-1}p_2^{-1}\cdots p_{i}^{-1}D_{i}}\) such that

$$\begin{aligned} \inf _{\lambda \in \Lambda , \ |y|\le \theta _0}\left\{ \prod _{j=1}^c \left| m_{p_{n_2+1}^{-1}\cdots p_{n_2+j}^{-1} D_{n_2+j}}\big (\frac{\lambda }{p_1p_2\cdots p_{n_2}}+y\big )\right| \right\} >\varepsilon . \end{aligned}$$
(2.8)

Proof

We first construct the spectrum \(\Lambda \). Write \(D=\max \{d_n:n\ge 1\}\) and \(\mathcal {S}=\{i:n_1+1\le i\le n_2\}\). We divide the set \(\mathcal {S}\) into two parts \(\mathcal {S}_1\) and \(\mathcal {S}_2\), where

$$\begin{aligned} \mathcal {S}_1=\left\{ i\in \mathcal {S}:\frac{D\theta (p_1\cdots p_{\ell _i})}{\theta (p_1\cdots p_{n_2+1})}\ge 1\right\} \ \text {and}\ \mathcal {S}_2=\left\{ i\in \mathcal {S}:\frac{D\theta (p_1\cdots p_{\ell _i})}{\theta (p_1\cdots p_{n_2+1})}<1\right\} . \end{aligned}$$

Take

$$\begin{aligned} a_i=\left\{ \begin{array}{ll} 2^{k_i}\theta (p_1\cdots p_{n_2+c}),&{}i\in \mathcal {S}_1,\\ 2^{k_i}\theta (p_1\cdots p_{\ell _i}),&{}i\in \mathcal {S}_2. \end{array} \right. \end{aligned}$$
(2.9)

[2, Lemma 4.1 (i)] shows that \(n_2+c\ge \ell _i\) for all \(i\in \mathcal {S}_1\). By the definition of \(U_n\), it is clear that \(a_i\in U_i\) for all \(i\in \mathcal {S}\). Then, Lemma 2.3 shows that \(\Lambda =\sum _{i=n_1+1}^{n_2}\{0,\ a_i\}\) is a spectrum of \(*_{i=n_1+1}^{n_2} \delta _{p_1^{-1}p_2^{-1}\cdots p_{i}^{-1}D_{i}}\). By the definition of the function \(\theta \), for \(i\in \mathcal {S}_1\) we have

$$\begin{aligned} \frac{d_{n_2+j}a_i}{p_1\cdots p_{n_2+j}}= \frac{\theta (d_{n_2+j})\theta (p_{n_2+j+1}\cdots p_{n_2+c})}{2^{k_{n_2+j}+1-k_i}} \in 2^{k_i-k_{n_2+j}-1}(2{\mathbb {Z}}+1). \end{aligned}$$
(2.10)

According to the definition of \(\mathcal {S}_2\), we have \(\ell _i\le n_2\) for any \(i\in \mathcal {S}_2\). Thus we have

$$\begin{aligned} \frac{d_{n_2+j}a_i}{p_1\cdots p_{n_2+j}}=\frac{2^{k_i-k_{n_2+j}-1}\theta (d_{n_2+j})}{\theta (p_{\ell _i+1}\cdots p_{n_2+j})},\ i\in \mathcal {S}_2. \end{aligned}$$
(2.11)

Given \(\lambda =\sum \limits _{i=n_1+1}^{n_2}b_i\in \Lambda \) with \(b_i\in \{0,\ a_i\}(n_1+1\le i\le n_2)\), write \(k_{i_1}=\min \{k_i:\ i\in \mathcal {S}_1,\ b_i\ne 0\}\) when the set \(\{i:\ i\in \mathcal {S}_1,\ b_i\ne 0\}\) is not empty. According to (2.10) and the assumption that \(k_i\ne k_j\) for any \(i\ne j\), we have

$$\begin{aligned} \left\{ \begin{array}{ll} \sum _{i\in \mathcal {S}_1}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}\in 2^{k_{i_1}-1-k_{n_2+j}}(2{\mathbb {Z}}+1),&{}\{i:\ i\in \mathcal {S}_1,\ b_i\ne 0\}\ne \emptyset ,\\ \sum _{i\in \mathcal {S}_1}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}=0,&{}\{i:\ i\in \mathcal {S}_1,\ b_i\ne 0\}=\emptyset . \end{array}\right. \end{aligned}$$
(2.12)

Let \(k_{i_2}=\max \{k_i:\ i\in \mathcal {S}_2,b_i\ne 0\}\) when the set \(\{i:i\in \mathcal {S}_2, b_i\ne 0\}\) is not empty. According to the definition of \(\mathcal {S}_2\), we have \(\theta (d_{n_2+j})\le D<\theta (p_{\ell _i+1}\cdots p_{n_2+1})\) for \(i\in \mathcal {S}_2\). Thus we have \(\theta (d_{n_2+j})+2\le \theta (p_{\ell _i+1}\cdots p_{n_2+1})\). Hence, \(\frac{\theta (d_{n_2+j})}{\theta (p_{\ell _i+1}\cdots p_{n_2+1})}\le \frac{\theta (d_{n_2+j})}{\theta (d_{n_2+j})+2}\le \frac{D}{D+2}\) for any \(i\in \mathcal {S}_2\). Also by (2.11) and the assumption that \(k_n\ne k_m\) for any \(n\ne m\), we get

$$\begin{aligned} 0\le \sum _{i\in \mathcal {S}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}<\frac{D}{D+2}\sum _{s\ge 0}2^{k_{i_2}-1-k_{n_2+j }-s}=\frac{D}{D+2}2^{k_{i_2}-k_{n_2+j}}. \end{aligned}$$
(2.13)

Given \(1\le j\le c\), we consider

$$\begin{aligned} m_{p^{-1}_{n_2+1}\cdots p^{-1}_{n_2+j}D_{n_2+j}}\left( \frac{\lambda }{p_1p_2\cdots p_{n_2}}\right) =m_{\{0,1\}}\left( \frac{\lambda d_{n_2+j}}{p_1p_2\cdots p_{n_2+j}}\right) ,\ \lambda \in \Lambda . \end{aligned}$$
(2.14)

And then we will deal with two cases.

Case A. \(\{i:\ i\in \mathcal {S}_1,\ b_i\ne 0\}=\emptyset \) or \(k_{i_1}>k_{n_2+j}\).

From (2.12) it is clear that

$$\begin{aligned} \sum _{i\in \mathcal {S}_1}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}\in {\mathbb {Z}}. \end{aligned}$$
(2.15)

Noting that \(i_2\in {\mathcal {S}}_2\), we see \(\ell _{i_2}\le n_2\), which implies \(k_{n_2+j}>k_{i_2}\). In virtue of (2.13), we get

$$\begin{aligned} \sum \limits _{i\in {\mathcal {S}}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}\in \left( 0,\ \frac{1}{2}\frac{D}{D+2}\right) . \end{aligned}$$

Since \(m_{\{0,1\}}\) has period 1, we have

$$\begin{aligned} \begin{array}{rl} m_{\{0,1\}}\left( \frac{\lambda d_{n_2+j}}{p_1p_2\cdots p_{n_2+j}}\right) &{}=m_{\{0,1\}}\left( \sum \limits _{i=n_1+1}^{n_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}\right) \\ &{}=m_{\{0,1\}}\left( \sum \limits _{i\in {\mathcal {S}}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}\right) \\ &{}=m_{\{0,1\}}(\alpha ) \end{array} \end{aligned}$$
(2.16)

for some \(\alpha \in \left( 0,\ \frac{1}{2}\frac{D}{D+2}\right) \).

Case B. \(k_{i_1}<k_{n_2+j}\).

Without loss of generality, we assume that the set \(\{i:i\in \mathcal {S}_2,b_i\ne 0\}\) is not empty. Otherwise, we have \(\sum \limits _{i\in {\mathcal {S}}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}=0\). From the definitions of \(i_1\) and \(i_2\), it follows that \(k_{i_1}>k_{i_2}\). By (2.13), we get

$$\begin{aligned} 0\le \sum _{i\in \mathcal {S}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}<\frac{D}{D+2}2^{k_{i_2}-k_{n_2+j}}\le \frac{D}{D+2}2^{k_{i_1}-k_{n_2+j}-1}. \end{aligned}$$

Combining (2.12), this shows there exists an integer z such that

$$\begin{aligned} \sum _{i\in \mathcal {S}_1\cup \mathcal {S}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}}=2^{k_{i_1}-k_{n_2+j}-1}(2z+1+\eta ) \end{aligned}$$

for some \(\eta \in \left( 0,\frac{D}{D+2}\right) \).

Given a real number \(r\in \mathbb {R}\), we denote by \(||r||_\frac{1}{2}\) the distance between r and it’s nearest middle point of two neighboring integer points, i.e.

$$\begin{aligned} ||r||_\frac{1}{2}=\min _{z\in \mathbb {Z}}\left\{ \left| r-z-\frac{1}{2}\right| , \left| r-z+\frac{1}{2}\right| \right\} . \end{aligned}$$

Thus the assumption \(k_{i_1}<k_{n_2+j}\) implies

$$\begin{aligned} \left| \left| 2^{k_{i_1}-k_{n_2+j}-1}(2z+1)\right| \right| _{\frac{1}{2}}\ge 2^{k_{i_1}-k_{n_2+j}-1}. \end{aligned}$$

By noting that \(\left| 2^{k_{i_1}-k_{n_2+j}-1}\eta \right| <2^{k_{i_1}-k_{n_2+j}-1}\frac{D}{D+2}\), we get

$$\begin{aligned} \begin{array}{rl} \left| \left| \frac{\lambda d_{n_2+j}}{p_1p_2\cdots p_{n_2+j}}\right| \right| _{\frac{1}{2}}&{}= \left| \left| \sum _{i\in \mathcal {S}_1\cup \mathcal {S}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}} \right| \right| _{\frac{1}{2}}\\ &{}=\left| \left| 2^{k_{i_1}-k_{n_2+j}-1}(2z+1+\eta ) \right| \right| _{\frac{1}{2}} \\ &{} \ge 2^{k_{i_1}-k_{n_2+j}-1}\frac{2}{D+2}. \end{array} \end{aligned}$$
(2.17)

Since \(i_1\in \mathcal {S}_1\), from Lemma 2.4 (i) it follows that \(i_1>n_2-N_1\). By Lemma 2.4 (iii), we have

$$\begin{aligned} \begin{array}{rl} k_{n_2+j}-k_{i_1}&{}=\nu _2(p_1\cdots p_{n_2+j})-\nu _2(2d_{n_2+j})-\nu _2(p_1\cdots p_{i_1})+\nu _2(2d_{i_1})\\ &{}\le \nu _2(p_{i_1+1}\cdots p_{n_2+j})+\nu _2(d_{i_1})\\ &{}\le (n_2+j-i_1+1)\max \{\nu _2(d_n):n\ge 1\}\\ &{}\le (N_1+c+2)\max \{\nu _2(d_n):n\ge 1\}. \end{array} \end{aligned}$$
(2.18)

Together with (2.17) and the boundedness of \(\{d_n\}_{n=1}^\infty \), we conclude that there is a positive constant \(0<\theta <\frac{1}{2}\) such that

$$\begin{aligned}\left| \left| \frac{\lambda d_{n_2+j}}{p_1p_2\cdots p_{n_2+j}}\right| \right| _{\frac{1}{2}}= \left| \left| \sum _{i\in \mathcal {S}_1\cup \mathcal {S}_2}\frac{d_{n_2+j}b_i}{p_1\cdots p_{n_2+j}} \right| \right| _{\frac{1}{2}}>\theta . \end{aligned}$$

Therefore, combining the conclusions of Case A and Case B we see the modulus of (2.14) has a positive lower bound. Furthermore, there is a constant \(\varepsilon >0\) such that

$$\begin{aligned} \prod _{j=1}^c \left| m_{p_{\ell _n+1}^{-1}\cdots p_{\ell _n+j}^{-1} D_{\ell _n+j}}\left( \frac{\lambda }{p_1p_2\cdots p_{n_2}}\right) \right| >2\varepsilon ,\quad \forall \ \lambda \in \Lambda . \end{aligned}$$

Finally, note that

$$\begin{aligned} \{\delta _{p_{n+1}^{-1}D_{n+1}}*\delta _{p_{n+1}^{-1} p_{n+2}^{-1} D_{n+2}}*\cdots *\delta _{p_{n+1}^{-1}\cdots p_{n+c}^{-1} D_{n+c}}:\ n>0\} \end{aligned}$$

is a family of probability measures supported on subsets of \([0,\ 1]\). Hence, their Fourier transformations are equi-continuous (cf [2, Definition 4.4 (iii)]). Thus we see that there is a small positive number \(\theta _0>0\) such that (2.8) holds for some constant \(\varepsilon >0\). The proof is completed. \(\square \)

Furthermore, we have the following Lemma 2.6. For \(k\ge 1\), we write

$$\begin{aligned} \mu _{>k}:=\delta _{p_{k+1}^{-1}D_{k+1}}*\delta _{p_{k+1}^{-1}p_{k+2}^{-1}D_{k+2}}*\cdots . \end{aligned}$$

Lemma 2.6

Assume that \(k_n\ne k_m\) for all \(n\ne m\) and (2.3) holds. Furthermore, assume that there exists a positive integer \(n_0\) such that for any \(n\ge n_0\), there exists an integer \(j_n<\ell _n\) satisfying \(\ell _{j_n}>\ell _n\). Consider the set \(\Lambda \) defined in Lemma 2.5 for \(n_1,\ n_2\in \mathcal {I}\) satisfying \(n_2>n_1+N_1\). There are small positive constants \(\varepsilon _1>0\) and \(\theta _1>0\) such that for any \(\lambda \in \Lambda \), there exists an integer \(b_\lambda \in {\mathbb {Z}}\) with \(b_0=0\) such that

$$\begin{aligned} \left| \widehat{\mu _{>n_2}}(y+\frac{\lambda +p_1p_2\cdots p_{n_2+c}b_\lambda }{p_1p_2\cdots p_{n_2}})\right| >\varepsilon _1,\ \ \forall \ y\in [-\theta _1,\ \theta _1],\ \lambda \in \Lambda . \end{aligned}$$
(2.19)

Proof

By [2, Lemma 4.5], there are small positive constants \(\varepsilon '>0\) and \(\theta _0>\theta _1>0\) such that for any \(\lambda \in \Lambda \), there exists an integer \(b_\lambda \) with \(b_0=0\) such that

$$\begin{aligned} \left| \widehat{\mu _{>n_2+c}}(y+b_\lambda +\frac{\lambda }{p_1p_2\cdots p_{n_2+c}})\right| >\varepsilon ',\ \ \forall \ y\in [-\theta _1,\ \theta _1]. \end{aligned}$$
(2.20)

Recall a fact that the mask function \(m_{\{0,1\}}(x)\) has period 1. For any \(\lambda \in \Lambda \), we have

$$\begin{aligned} \begin{array}{rl} &{}\left| \widehat{\mu _{>n_2}}(y+\frac{\lambda +p_1p_2\cdots p_{n_2+c}b_\lambda }{p_1p_2\cdots p_{n_2}})\right| \\ &{}\quad = \left| \widehat{\mu _{>n_2+c}}(y+\frac{\lambda +p_1p_2\cdots p_{n_2+c}b_\lambda }{p_1p_2\cdots p_{n_2+c}})\right| \cdot \prod _{j=1}^c \left| m_{p_{n_2+1}^{-1}\cdots p_{n_2+j}^{-1} D_{n_2+j}}\big (y+\frac{\lambda +p_1p_2\cdots p_{n_2+c}b_\lambda }{p_1p_2\cdots p_{n_2}}\big )\right| \\ &{}\quad = \left| \widehat{\mu _{>n_2+c}}(y+b_\lambda +\frac{\lambda }{p_1p_2\cdots p_{n_2+c}})\right| \cdot \prod _{j=1}^c \left| m_{p_{n_2+1}^{-1}\cdots p_{n_2+j}^{-1} D_{n_2+j}}\big (y+\frac{\lambda }{p_1p_2\cdots p_{n_2}}\big )\right| .\\ \end{array} \end{aligned}$$
(2.21)

Lemma 2.5 shows there are small constants \(\varepsilon >0\) and \(\theta _0\) such that

$$\begin{aligned} \prod _{j=1}^c \left| m_{p_{n_2+1}^{-1}\cdots p_{n_2+j}^{-1} D_{n_2+j}}\big (y+\frac{\lambda }{p_1p_2\cdots p_{n_2}}\big )\right| >\varepsilon ,\ \ \forall \ y\in [-\theta _0,\ \theta _0]. \end{aligned}$$
(2.22)

Letting \(\varepsilon _1=\varepsilon \varepsilon '\), the inequality (2.19) follows from (2.20), (2.21) and (2.22). The proof is completed. \(\square \)

Now we are in the place to reprove the sufficiency of [2, Theorem 1.1].

Proof of the sufficiency of [2, Theorem 1.1].

We shall deal with two cases.

(A) If there is an infinite subset \({\mathcal {I}}_0\subset {\mathcal {I}}\) (\({\mathcal {I}}\) is defined in (2.5)) such that \(\ell _i\le n\) for any \(i\le n\) and \(n\in {\mathcal {I}}_0\). Then, the proof in [2] works by replacing \({\mathcal {B}}\) by \({\mathcal {I}}_0\).

(B) If there are only finitely many \(n\in {\mathcal {I}}\) such that \(\ell _i\le n\) for any \(i\le n\). Then, there is an integer \(n_0>0\) such that for any \(n\in {\mathcal {I}}\) with \(n\ge n_0\), there exists at least one integer \(j_n<\ell _n\) satisfying \(\ell _{j_n}>\ell _n\). Also, as stated in the beginning of this section, all conditions in (2.3) can be assumed without loss of generality.

Then, we extend the idea of [1, Lemma 2.6] and [1, Theorem 2.7] to construct a spectrum of \(\mu \). This spectrum is different from the one in [2]. Let \({\mathcal {I}}_1=\{n\in \mathcal {I}:\ n>n_0\}\)

We first choose \(n_1\in {\mathcal {I}}_1\) and define

$$\begin{aligned} \Lambda _1=\{0,\ a_1\}+\{0,\ a_2\}+\cdots +\{0,\ a_{n_1}\}, \end{aligned}$$

where \(a_i=2^{k_i}\theta (p_1\cdots p_{\ell _i})\in U_i\) for \(1\le i\le n_1\). Since \({\mathcal {I}}_1\) is infinite and \(p_n\ge 2\), we can find a sufficiently large integer \(n_2\in {\mathcal {I}}_1\) such that \(n_2>n_1+N_1\) and

$$\begin{aligned} (p_1p_2\cdots p_{n_2})^{-1} \Lambda _1\subset \Big [-\frac{\theta _1}{2^2},\ \frac{\theta _1}{2^2}\Big ], \end{aligned}$$

where \(N_1\) and \(\theta _1\) are defined in Lemma 2.4 and 2.6, respectively. Let \(\epsilon _1\) be the constant in Lemma 2.6 and \(\Lambda _{1,2}\) be a spectrum of \(*_{i=n_1+1}^{n_2} \delta _{p_1^{-1}p_2^{-1}\cdots p_{i}^{-1}D_{i}}\) as stated in Lemma 2.5, i.e.

$$\begin{aligned} \Lambda _{1,2}=\{0,\ a_{n_1+1}\}+\{0,\ a_{n_1+2}\}+\cdots +\{0,\ a_{n_2}\}, \end{aligned}$$

where

$$\begin{aligned} a_i=\left\{ \begin{array}{ll} 2^{k_i}\theta (p_1\cdots p_{n_2+c}),&{} \text {if}\ \frac{D\theta (p_1\cdots p_{\ell _i})}{\theta (p_1\cdots p_{n_2+1})}\ge 1,\\ 2^{k_i}\theta (p_1\cdots p_{\ell _i}),&{}\text {if}\ \frac{D\theta (p_1\cdots p_{\ell _i})}{\theta (p_1\cdots p_{n_2+1})}<1. \end{array} \right. \end{aligned}$$
(2.23)

According to Lemma 2.6, for any \(\lambda \in \Lambda _{1,2}\), there exits an integer \(k_{1,\lambda }\in {\mathbb {Z}}\) with \(k_{1,\ 0}=0\) such that

$$\begin{aligned} \Big |\widehat{\mu _{>n_2}}\Big (\frac{ \gamma }{p_1p_2\cdots p_{n_2}}+ \frac{ \lambda +p_1\cdots p_{n_2+c}k_{1,\ \lambda }}{p_1p_2\cdots p_{n_2}} \Big )\Big |>\varepsilon _1,\quad \forall \ \gamma \in \Lambda _1,\ \lambda \in \Lambda _{1,2}. \end{aligned}$$

Lemma 2.3 (ii) and (iii) show that \(\Lambda _2:=\{\gamma +\lambda +p_1p_2\cdots p_{n_2+c}k_{1,\ \lambda }:\ \gamma \in \Lambda _1,\ \lambda \in \Lambda _{1,2}\}\) is a spectrum of the probability measure \(\delta _{p_1^{-1}D_1}*\delta _{p_1^{-1}p_2^{-1}D_2}*\cdots *\delta _{p_1^{-1}p_2^{-1}\cdots p_{n_2}^{-1}D_{n_2}}\). Furthermore, [2, Lemma 4.1] and the definitions of \(U_i\) show that \(U_i+p_1p_2\cdots p_{n_2+c}k_{1,\ \lambda }=U_i\) for all \(i\le n_2\). Hence, by \(k_{1,\ 0}=0\) and the definitions of \(\Lambda _1\) and \(\Lambda _2\), we see \(\Lambda _1\subset \Lambda _2\subset \sum \limits _{j=1}^{n_2}(\{0\}\cup U_j)\). In a word, we have

$$\begin{aligned} \Big |\widehat{\mu _{>n_2}}\Big (\frac{ \lambda }{p_1p_2\cdots p_{n_2}}\Big )\Big |>\varepsilon _1,\quad \forall \ \lambda \in \Lambda _2. \end{aligned}$$
(2.24)

Continuing in this way, we can find a strictly increasing sequence \(\{n_k\}_{k=1}^\infty \subset {\mathcal {I}}_1\) and \(\Lambda _k\) such that the following properties (2.25), (2.26), (2.27) and claim hold.

$$\begin{aligned} 0\in \Lambda _{k}\subset \Lambda _{k+1}\subset \sum \limits _{j=1}^{n_{k+1}}(\{0\}\cup U_j),\quad k=1,\ 2,\cdots , \end{aligned}$$
(2.25)
$$\begin{aligned} (p_1p_2\cdots p_{n_{k+1}})^{-1} \Lambda _k \subset \Big [-\frac{\theta _1}{2^{k+1}},\ \frac{\theta _1}{2^{k+1}}\Big ],\quad k=1,\ 2,\cdots ,\end{aligned}$$
(2.26)
$$\begin{aligned} \Big |\widehat{\mu _{>n_k}}\Big (\frac{\lambda }{p_1p_2\cdots p_{n_k}}\Big )\Big |>\varepsilon _1,\quad \forall \ \lambda \in \Lambda _k,\ k=2,\ 3,\cdots . \end{aligned}$$
(2.27)

Claim. The set \(\Lambda _k\) is a spectrum of the probability measure \(\delta _{p_1^{-1}D_1}*\delta _{p_1^{-1}p_2^{-1}D_2}*\cdots *\delta _{p_1^{-1}p_2^{-1}\cdots p_{n_k}^{-1}D_{n_k}}\) for all \(k=1,\ 2,\cdots \).

Let \(\Gamma =\bigcup \limits _{k=1}^\infty \Lambda _k\). We shall prove \(\Gamma \) is a spectrum of \(\mu \).

For any \(a\ne b\in \Gamma \), from (2.25) it follows that \(a\ne b\in \Lambda _k\) for some \(k>0\). Hence, \(a-b\) is a zero point of the Fourier transform of \(\delta _{p_1^{-1}D_1}*\delta _{p_1^{-1}p_2^{-1}D_2}*\cdots *\delta _{p_1^{-1}p_2^{-1}\cdots p_{n_k}^{-1}D_{n_k}}\). Hence, \(\widehat{\mu }(a-b)=0\), which implies the exponential function set \(E_\Gamma =\{e^{2\pi i\gamma x}:\gamma \in \Gamma \}\) is an orthogonal family of \(L^2(\mu )\).

Assume, on the contrary, that \(\Gamma \) is not a spectrum of \(\mu \). Then, [2, Proposition 2.3] shows that \(Q_{\mu ,\Gamma }(x_0)<1\) for some \( x_0\in {\mathbb {R}}\).

Recall that \(\lim \limits _{k\rightarrow \infty }(p_1p_2\cdots p_{n_{k}})^{-1}x_0=0\) and \(\widehat{\Phi }:=\{\widehat{\nu }:\nu \in \Phi \}\) (here \(\Phi =\{\mu _{>n}:\ n\ge 1\}\)) is equi-continuous. From (2.26) it follows that

$$\begin{aligned} \beta _{k}:=\inf _{\lambda \in \Lambda _k}|\widehat{\mu _{>n_k}}((p_1p_2\cdots p_{n_{k+1}})^{-1}(\lambda +x_0))|\rightarrow 1 \text{ as } k\rightarrow \infty . \end{aligned}$$
(2.28)

Furthermore, from (2.27) it follows that there exists a positive integer \(k_0>0\) such that for any \(k\ge k_0\) and \(\lambda \in \Lambda _k\), we have

$$\begin{aligned} |\widehat{\mu _{>n_k}}((p_1p_2\cdots p_{n_{k}})^{-1}(\lambda +x_0))|\ge \frac{1}{2}\varepsilon _1. \end{aligned}$$
(2.29)

Let

$$\begin{aligned} Q_{k}(x_0)=\sum \limits _{\lambda \in \Lambda _k}|\widehat{\mu } (\lambda +x_0)|^2,\quad \ k=1,2,\cdots . \end{aligned}$$

According to [2, (2.2)] and (2.29), for \(k\ge k_0\) we have

$$\begin{aligned} \begin{array}{rl} &{}Q_{k+1}(x_0)-Q_{k}(x_0)\\ &{}\quad = \sum \limits _{\lambda \in \Lambda _{k+1}\setminus \Lambda _k}\prod _{n=1}^\infty |m_{D_n}((p_1\cdots p_n)^{-1}(\lambda +x_0))|^2\\ &{}\quad = \sum \limits _{\lambda \in \Lambda _{k+1}\setminus \Lambda _k}\prod _{n=1}^{n_{k+1}} |m_{D_n}((p_1\cdots p_n)^{-1}(\lambda +x_0)|^2|\widehat{\mu _{>n_{k+1}}}((p_1\cdots p_{n_{k+1}})^{-1}(\lambda +x_0))|^2\\ &{}\quad \ge \frac{1}{4}\varepsilon ^2_1\sum \limits _{\lambda \in \Lambda _{k+1}\setminus \Lambda _k}\prod _{n=1}^{n_{k+1}} |m_{D_n}((p_1\cdots p_n)^{-1}(\lambda +x_0))|^2. \end{array} \end{aligned}$$
(2.30)

The above claim shows

$$\begin{aligned} \sum \limits _{\lambda \in \Lambda _{k+1}}\prod _{n=1}^{n_{k+1}} \left| m_{D_n}(p_1^{-1}\cdots p_n^{-1}(\lambda +x_0))\right| ^2=1,\quad k=1,\ 2,\cdots . \end{aligned}$$

Thus (2.30) implies that for any \(k\ge k_0\), we have

$$\begin{aligned} Q_{k+1}(x_0)-Q_{k}(x_0) \ge \frac{1}{4}\varepsilon ^2_1\left( 1-\sum \limits _{\lambda \in \Lambda _k}\prod _{n=1}^{n_{k+1}}\left| m_{D_n}((p_1\cdots p_n)^{-1}(\lambda +x_0))\right| ^2\right) . \end{aligned}$$

On the other hand, by the definition of \(\beta _k\) in (2.28) for \(k\ge 1\), we have

$$\begin{aligned} \begin{array}{rl} Q_{k}(x_0)=&{}\sum \limits _{\lambda \in \Lambda _k}\prod _{n=1}^\infty |m_{D_n}((p_1\cdots p_n)^{-1}(\lambda +x_0)|^2\\ \ge &{} \beta ^2_k\sum \limits _{\lambda \in \Lambda _k}\prod _{n=1}^{n_{k+1}}\left| m_{D_n}((p_1\cdots p_n)^{-1}(\lambda +x_0))\right| ^2. \end{array} \end{aligned}$$

By the above inequality, we have

$$\begin{aligned} Q_{k+1}(x_0)-Q_{k}(x_0) \ge \frac{1}{4}\varepsilon ^2_1\left( 1-\beta ^{-2}_kQ_k(x_0)\right) ,\quad \forall \ k\ge k_0. \end{aligned}$$

Therefore, the limit property in (2.28) shows

$$\begin{aligned}\begin{array}{rl} \liminf \limits _{k\rightarrow \infty }(Q_{k+1}(x_0)-Q_{k}(x_0))\ge &{}\frac{1}{4}\varepsilon ^2_1\left( 1-\lim \limits _{k\rightarrow \infty }\beta _k^{-2}Q_{k}(x_0)\right) \\ =&{}\frac{1}{4}\varepsilon ^2_1\left( 1-Q_{\mu ,\Gamma }(x_0)\right) >0. \end{array} \end{aligned}$$

Together with (2.25), the above inequalities imply

$$\begin{aligned} 1>Q_{\mu ,\Gamma }(x_0)=\lim \limits _{k\rightarrow \infty }Q_{k}(x_0)\ge \sum \limits _{k=1}^\infty \left( Q_{k+1}(x_0)-Q_{k}(x_0)\right) =+\infty , \end{aligned}$$

which is impossible. Hence, \(\Gamma \) is a spectrum of \(\mu \). The sufficiency of [2, Theorem 1.1] is proven. \(\square \)