1 Introduction

A partition \(\lambda \) of a positive integer n is a finite non-increasing sequence of positive integers \(\lambda =(\lambda _1,\ldots , \lambda _l)\) such that \(|\lambda |=\sum _{i=1}^{l}\lambda _i=n\). The Ferrers graph of a partition \(\lambda \) is a set of coordinates in the bottom right quadrant of the plane where the i-th row contains \(\lambda _i\) dots. We denote by \(\lambda '\) the conjugate of \(\lambda \), which is the partition whose graph is obtained by reflecting the Ferrers graph of \(\lambda \) about the main diagonal. For example, we give \(\lambda =(4,4,2,2,1)\) and its conjugate partition \(\lambda '=(5,4,2,2)\) in Fig. 1.

Fig. 1
figure 1

Ferrers graph of partitions \(\lambda =(4,4,2,2,1)\) and \(\lambda '=(5,4,2,2)\)

Let p(n) denote the ordinary partition function. The partition statistic crank defined by Andrews and Garvan [2, 6] can be used to provide combinatorial interpretations for Ramanujan’s famous congruences

$$\begin{aligned} p(5n+4)&\equiv 0 \pmod {5}, \nonumber \\ p(7n+5)&\equiv 0 \pmod {7},\nonumber \\ p(11n+6)&\equiv 0\pmod {11}. \end{aligned}$$
(1.1)

The crank of a partition \(\lambda \ne (1)\) is defined as follows:

$$\begin{aligned} c(\lambda )=\left\{ \begin{array}{ll} \lambda _1,\ \ if\ \ n_1(\lambda )=0,\\ \mu (\lambda )-n_1(\lambda ), \ \ if \ n_1(\lambda )>0, \end{array}\right. \end{aligned}$$

where \(n_1(\lambda )\) denotes the number of parts equal to one in \(\lambda \) and \(\mu (\lambda )\) denotes the number of parts in \(\lambda \) larger than \(n_1(\lambda )\). Let M(mn) enumerate partitions of n with crank m. It should be pointed out that when \(\lambda =(1)\),

$$\begin{aligned} M(0,1)=-1, \ \ M(-1,1)=M(1,1)=1. \end{aligned}$$

Andrews and Garvan [2, 6] established the generating function of M(mn) as given by

$$\begin{aligned} \sum _{n=0}^{\infty }\sum _{m=-\infty }^{\infty } M(m,n)z^mq^n =\frac{(q;q)_\infty }{(zq;q)_\infty (z^{-1}q;q)_\infty }. \end{aligned}$$
(1.2)

Here and throughout this paper, \((a;q)_\infty \) stands for the q-shifted factorial

$$\begin{aligned} (a;q)_{\infty }=\prod _{n=1}^{\infty }(1-aq^{n-1}),\,\, |q|<1, \end{aligned}$$

and for any positive integer k,

$$\begin{aligned} f_k=(q^k;q^k)_{\infty }. \end{aligned}$$

Let M(mtn) denote the number of partitions of n with crank congruent to m modulo t. In 1990, Garvan [7] presented a graceful refinement of the congruence (1.1)

$$\begin{aligned} M(m,2,5n+4)\equiv 0 \pmod {5},\ \ m=0,1, \end{aligned}$$

together with the combinatorial interpretation

$$\begin{aligned} M(m+2k,10,5n+4)=\frac{M(m,2,5n+4)}{5},\ \ 0\le k\le 4,m=0,1. \end{aligned}$$

Let ped(n) be the function that enumerates partitions of n with even parts distinct. Obviously,

$$\begin{aligned} \sum _{n=0}^{\infty }ped(n)q^n =\frac{(-q^2;q^2)_{\infty }}{(q;q^2)_{\infty }}. \end{aligned}$$

The sequence \({ped(n)}_{n\ge 0}\) is well known and can be seen in [11, A001935], as well as other combinatorial interpretations. In 2010, Andrews, Hirschhorn, and Sellers [3] proved the following congruences.

Theorem 1.1

For \(\alpha ,n \ge 0\),

$$\begin{aligned} ped\left( 3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&\equiv 0\pmod {2},\\ ped\left( 3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&\equiv 0\pmod {2}. \end{aligned}$$

Theorem 1.2

For \(n \ge 0\),

$$\begin{aligned} ped(9n+4)&\equiv 0\pmod {4},\end{aligned}$$
(1.3)
$$\begin{aligned} ped(9n+7)&\equiv 0\pmod {12}. \end{aligned}$$
(1.4)

In 2017, Merca [9] provided a simple criterion for deciding the parity of ped(n).

Theorem 1.3

The number of partitions of n with distinct even parts is odd if and only if n is a triangular number.

In this paper, we aim at introducing a partition statistic which we call ped-crank to study the partition function ped(n). Let \(M_{ped}(m,n)\) denote the number of partitions of n with even parts distinct with ped-crank m, and let

$$\begin{aligned} M_{ped}(m,t,n)=\sum _{k\equiv m\pmod {t}} M_{ped}(k,n). \end{aligned}$$
(1.5)

The main results of this paper are summarized below.

Theorem 1.4

For \(\alpha ,n \ge 0\),

$$\begin{aligned} M_{ped}\left( 0,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&= M_{ped}\left( 3,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) ,\\ M_{ped}\left( 1,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&= M_{ped}\left( 2,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) ,\\ M_{ped}\left( 0,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&= M_{ped}\left( 3,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) ,\\ M_{ped}\left( 1,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&= M_{ped}\left( 2,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) . \end{aligned}$$

Any of the following three corollaries deduced from Theorem 1.4 provides a combinatorial interpretation or a refinement of Theorem 1.1. When \(\alpha =0\), combining Corollary 1.5 and Corollary 1.6 refines (1.3). Meanwhile, Corollary 1.7 combinatorially interprets (1.3).

Corollary 1.5

For \(m=0,1,2\) and \(\alpha ,n \ge 0\),

$$\begin{aligned} M_{ped}\left( m,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&=\frac{M_{ped}\left( m,3,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) }{2},\\ M_{ped}\left( m,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&=\frac{M_{ped}\left( m,3,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) }{2}. \end{aligned}$$

Corollary 1.6

For \(\alpha ,n \ge 0\),

$$\begin{aligned} M_{ped}\left( 3,12,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) =\,&\frac{M_{ped}\left( 0,3,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) }{4},\\ M_{ped}\left( 3,12,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) =\,&\frac{M_{ped}\left( 0,3,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) }{4}. \end{aligned}$$

Corollary 1.7

For \(\alpha ,n \ge 0\),

$$\begin{aligned} M_{ped}\left( 1,4,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&=\frac{ped\left( 3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) }{4}, \\ M_{ped}\left( 1,4,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&=\frac{ped\left( 3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) }{4}. \end{aligned}$$

Theorem 1.8

If n cannot be written as a sum of a triangular number and a square of even integer, we have

$$\begin{aligned} M_{ped}(0,4,n)&=M_{ped}(2,4,n). \end{aligned}$$

Moreover, \(M_{ped}(0,n)\) is odd if and only if n is a triangular number.

Theorem 1.9

For \(m=0,1,2,3,4,5\) and \(n\ge 0\),

$$\begin{aligned} M_{ped}(m,6,9n+7)=\,&\frac{ped(9n+7)}{6}, \nonumber \\ M_{ped}(3,12,9n+7) =\,&\frac{ped(9n+7)}{12}. \end{aligned}$$
(1.6)

It is worth mentioning that Theorem 1.8 not only combinatorially interprets but also refines Theorem 1.3, and (1.6) provides a combinatorial interpretation for (1.4).

2 Definition of the ped-crank

In this section, we shall define the ped-crank of partitions with even parts distinct based on Glaisher’s bijection and a modified version \(\varphi \) of the Wright map established by Seo and Yee [10].

We first give a quick overview of Glaisher’s bijection and the Frobenius symbol. Let \(D_n\) denote the set of distinct partitions, and let \(O_n\) denote the set of odd partitions of n respectively. Glaisher’s bijection \(\phi \): \(O_n\rightarrow D_n\) is defined as follows. Let \(\lambda =(1^{m_1}3^{m_3}\ldots )\in O_n\) be an odd partition. For every odd i, let \(\phi (\lambda )\) contain part \(i\cdot 2^r\), if and only if the integer \(m_i\) written in binary has 1 at the r-th position. In the other direction, let \(\psi \): \(D_n\rightarrow O_n\) be defined by an iterative procedure. Start with \(\lambda =(\lambda _1,\lambda _2,\ldots )\in D_n\). Substitute every even part \(\lambda _i\) with two parts \(\lambda _i/2\). Repeat until the resulting partition has no even parts.

The Frobenius symbol of n is a two-rowed array [1, 14]

$$\begin{aligned} F=\left( \begin{array}{ccccc} \alpha _1 &{} \quad \alpha _2 &{} \cdots &{} \alpha _\ell \\ \beta _1 &{} \quad \beta _2 &{} \cdots &{} \beta _\ell \end{array}\right) , \end{aligned}$$

where \(\alpha _1>\alpha _2>\ldots >\alpha _{\ell }\ge 0\), \(\beta _1>\beta _2>\ldots >\beta _{\ell }\ge 0\) and \(n=|\alpha |+|\beta |+\ell \). If we express an ordinary partition by Ferrers graph, it is easy to see that \(\alpha _i\) form rows to the right of the diagonal and \(\beta _i\) form columns below the diagonal. Thus the Frobenius symbol is another representation of an ordinary partition. For instance, the Frobenius symbol for (8, 7, 4, 3, 1) is

$$\begin{aligned} \left( \begin{array}{ccccc} 7 &{}\quad 5 &{} \quad 1 \\ 4 &{}\quad 2 &{}\quad 1 \end{array}\right) . \end{aligned}$$

Giving a real number c, we define \(c\lambda \) as the partition whose parts are c times each part of \(\lambda \). For example, let \(\lambda =(4,2,2)\). We have \(4\lambda =(16,8,8)\) and \(\frac{1}{2}\lambda =(2,1,1)\). Suppose \(\mu \) and \(\nu \) are two partitions. Let \(\mu \cup \nu \) denote the partition consisting of all the parts of \(\mu \) and \(\nu \). The definition of ped-crank is given based on the following theorem.

Theorem 2.1

For integer \(k_1\ge -1\), \(k_2\ge 1\), there is a bijection \(\Delta \) between the set of partitions of n with even parts distinct and the set of vector partitions \((\alpha ,\beta ,\gamma )\) with \(|\alpha |+|\beta |+|\gamma |\) equal to n. Here \(\alpha \) is an even partition, \(\beta \) is a partition of the form \((4k_1+1,\ldots ,9,5,1)\) or \((4k_2-1,\ldots ,11,7,3)\) and \(\gamma \) is a distinct even partition.

Proof

The bijection \(\Delta \) can be decomposed into six weight preserving steps.

  1. Step 1.

    \(\lambda \rightarrow (\omega ,\gamma )\): Start with a partition \(\lambda \) of n with even parts distinct. Split \(\lambda \) into a pair of partitions \((\omega ,\gamma )\) according to the parts odd or even. It is clear that \(\omega \) is an odd partition and \(\gamma \) is a distinct even partition.

  2. Step 2.

    \((\omega ,\gamma )\rightarrow (\xi ,\gamma )\): By Glaisher’s bijection, let \(\phi (\omega )=\xi \). One can see that \(\xi \) is a distinct partition.

  3. Step 3.

    \((\xi ,\gamma )\rightarrow (\mu ^1,\mu ^2,\pi ,\gamma )\): Divide \(\xi \) into a triple of partitions \((\mu ^1,\mu ^2,\pi )\) according to the remainder of the parts mod 4. Here \(\mu ^1\ (\mu ^2)\) consist of all the parts congruent to 1(3) mod 4 and \(\pi \) consists of all the even parts of \(\xi \).

  4. Step 4.

    \((\mu ^1,\mu ^2,\pi ,\gamma )\rightarrow (\mu ^1,\mu ^2,\zeta ,\gamma )\): Let \(\zeta =2\psi (\frac{1}{2}\pi )\) by applying Glaisher’s bijection. Since \(\frac{1}{2}\pi \) is a distinct partition, we can say that \(\zeta \) is a partition with all parts congruent to 2 mod 4.

  5. Step 5.

    \((\mu ^1,\mu ^2,\zeta ,\gamma )\rightarrow (\eta ,\beta ,\zeta ,\gamma )\): Write \(\mu ^1\) and \(\mu ^2\) as

    $$\begin{aligned} \mu ^1&=(4a_1+1,4a_2+1,\ldots ,4a_{s+m}+1),\\ \mu ^2&=(4b_1+3,4b_2+3,\ldots ,4b_{s}+3), \end{aligned}$$

    where \(a_1>a_2>\cdots >a_{s+m}\ge 0\) and \(b_1>b_2>\cdots >b_s\ge 0\).

    1. Case 1.

      \(m\ge 0\). Using the bijection \(\varphi \) established by Seo and Yee [10], a Frobenius symbol

      $$\begin{aligned} \mu =\left( \begin{array}{llll} a_{1+m} &{}\quad a_{2+m} &{}\cdots &{}a_{s+m} \\ b_1 &{} \quad b_2 &{} \cdots &{} b_s \end{array}\right) \end{aligned}$$

      and a partition \(\nu =(a_1-m+1,a_2-m+2,\ldots ,a_m)\) can be constructed. Let \(\varphi (\mu ^1,\mu ^2)=(\eta ,\beta )\), where \(\eta =4(\mu \cup \nu )\) and \(\beta =(4(m-1)+1,4(m-2)+1,\ldots ,5,1)\).

    2. Case 2.

      \(m<0\). Correspondingly, a Frobenius symbol

      $$\begin{aligned} \mu =\left( \begin{array}{llll} b_{1-m} &{}\quad b_{2-m} &{} \cdots &{} b_s \\ a_{1} &{} \quad a_{2} &{}\cdots &{}a_{s+m} \end{array}\right) \end{aligned}$$

      and a partition \(\nu =(b_1+m+1,b_2+m+2,\ldots ,b_{-m})\) can be constructed. Let \(\varphi (\mu ^1,\mu ^2)=(\eta ,\beta )\), where \(\eta =4(\mu \cup \nu )'\) and \(\beta =(4(-m-1)+3,4(-m-2)+3,\ldots ,7,3)\).

  6. Step 6.

    \((\eta ,\beta ,\zeta ,\gamma )\rightarrow (\alpha ,\beta ,\gamma )\): Ultimately, let \(\alpha =\eta \cup \zeta \) and define \(\Delta (\lambda )=(\alpha ,\beta ,\gamma )\).

Furthermore, one sees that the above construction can be reversed. This completes the proof. \(\square \)

An example of the bijection \(\Delta \) is given below.

Example 2.2

[4]

$$\begin{aligned} \lambda =(30,25,18,13,11,9,6,&5,5,5,5,3,3,3,3,3,3,3,1,1,1)\\&\Updownarrow _{\text {Step 1.}} \\ (\omega ,\gamma )=\big ((25,13,11,9,5,5,5,5,&3,3,3,3,3,3,3,1,1,1),(30,18,6)\big )\\&\Updownarrow _{\text {Step 2.}} \\ (\xi ,\gamma )=\big ((25,20,13,12,&11,9,6,3,2,1),(30,18,6)\big )\\&\Updownarrow _{\text {Step 3.}} \\ (\mu ^1,\mu ^2,\pi ,\gamma )=\big ((25,13,9,1),&(11,3),(20,12,6,2),(30,18,6)\big )\\&\Updownarrow _{\text {Step 4.}} \\ (\mu ^1,\mu ^2,\zeta ,\gamma )=\big ((25,13,9,1),(11&,3),(10,10,6,6,6,2),(30,18,6)\big )\\&\Updownarrow _{\text {Step 5.}} \\ (\eta ,\beta ,\zeta ,\gamma )=\big ((20,12,12,8,4),(&5,1),(10,10,6,6,6,2),(30,18,6)\big )\\&\Updownarrow _{\text {Step 6.}} \\ (\alpha ,\beta ,\gamma )=\big ((20,12,12,10,10,&8,6,6,6,4,2),(5,1),(30,18,6)\big )\\ \end{aligned}$$

Now we are ready to give the definition of the ped-crank of a partition with even parts distinct under the bijection \(\Delta \).

Definition 2.3

Let \(\lambda \) be a partition with even parts distinct and \(\Delta (\lambda )=(\alpha , \beta , \gamma )\). The ped-crank of \(\lambda \), denoted by \(c_{ped}(\lambda )\), is defined as the crank of \(\frac{1}{2}\alpha \).

3 Generating function of \(M_{ped}(m,n)\)

This section focuses on the generating function of \(M_{ped}(m,n)\).

According to the bijection \(\Delta \), the generating function of \(\beta \) can be derived by using Jacobi’s triple product identity.

$$\begin{aligned} \sum _{n=0}^{\infty }q^{4{n \atopwithdelims ()2}+n}+ \sum _{n=-\infty }^{-1}q^{4{n \atopwithdelims ()2}+n} =(-q;q^4)_\infty (-q^3;q^4)_\infty (q^4;q^4)_\infty . \end{aligned}$$

Moreover, it is trivial that the generating function of \(\gamma \) is

$$\begin{aligned} (-q^2;q^2)_{\infty }. \end{aligned}$$

Since the ped-crank only relies on the even partition \(\alpha \), by (1.2), the generating function of \(M_{ped}(m,n)\) can be given as

(3.1)

By considering the transformation that interchanges z and \(z^{-1}\) in (3.1), we have

$$\begin{aligned} M_{ped}(m,n)=M_{ped}(-m,n). \end{aligned}$$

Thus, for any positive integer t,

$$\begin{aligned} M_{ped}(m,t,n)=M_{ped}(-m,t,n). \end{aligned}$$

In other words,

$$\begin{aligned} M_{ped}(m,t,n)=M_{ped}(t-m,t,n). \end{aligned}$$
(3.2)

4 Preliminaries

In this section we present some results that will be used in Section 5.

Lemma 4.1

$$\begin{aligned} (-q;-q)_\infty =\frac{f^3_2}{f_1f_4}. \end{aligned}$$

Proof

Replacing q by \(-q\) in \((q;q)_\infty \), we have

$$\begin{aligned} (-q;-q)_\infty =\,&(-q;q^2)_\infty (q^2;q^2)_\infty \\ = \,&\frac{(-q;q)_\infty (q^2;q^2)_\infty }{(-q^2;q^2)_\infty }\\ =\,&\frac{(q;q)_\infty (-q;q)_\infty (q^2;q^2)^2_\infty }{(q;q)_\infty (-q^2;q^2)_\infty (q^2;q^2)_\infty }\\ =\,&\frac{(q^2;q^2)^3_\infty }{(q;q)_\infty (q^4;q^4)_\infty }. \end{aligned}$$

\(\square \)

Lemma 4.2

( [4, Entry 22, p. 36])

$$\begin{aligned} \phi (q)=\sum _{n=-\infty }^{\infty }q^{n^2} =\frac{f_2^5}{f_1^2f_4^2}. \end{aligned}$$

Replacing q by \(-q\) in the above equation, we get

$$\begin{aligned} \phi (-q)=2\sum _{n=0}^{\infty }(-1)^nq^{n^2}-1=\frac{f_1^2}{f_2}. \end{aligned}$$

Lemma 4.3

( [4, p. 49])

$$\begin{aligned} \psi (q)=\sum _{n=0}^{\infty }q^{n+1 \atopwithdelims ()2}=\frac{f_2^2}{f_1}=\frac{f_6f_9^2}{f_3f_{18}}+q\frac{f_{18}^2}{f_9}. \end{aligned}$$

Theorem 4.4

( [5, Lemma 2.2])

$$\begin{aligned} \frac{1}{\varphi (-q)}&=\frac{\varphi ^3(-q^9)}{\varphi ^4(-q^3)}\left( 1+2q\omega (q^3)+4q^2\omega ^2(q^3)\right) ,\\ \frac{1}{\psi (q)}&=\frac{\psi ^3(q^9)}{\psi ^4(q^3)}\left( \frac{1}{\omega ^2(q^3)}-q\frac{1}{\omega (q^3)}+q^2\right) , \end{aligned}$$

where

$$\begin{aligned} \omega (q)=\frac{f_1f^3_6}{f_2f^3_3}. \end{aligned}$$

Lemma 4.5

( [13, Lemma 2.5])

$$\begin{aligned} \frac{f_1}{f^3_3}&=\frac{f_2f^2_4f^2_{12}}{f^7_{6}}-q\frac{f^3_2f^6_{12}}{f^2_4f^9_6}. \end{aligned}$$

Lemma 4.6

( [8, p. 5])

$$\begin{aligned} f_1f_2=\frac{f_6f^4_9}{f_3f^2_{18}}-qf_9f_{18} -2q^2\frac{f_3f^4_{18}}{f_6f^2_9}. \end{aligned}$$

Lemma 4.7

( [3, Theorem 3.1])

$$\begin{aligned} \frac{f_4}{f_1}&=\frac{f_{12}f^4_{18}}{f^3_3f^2_{36}}+q\frac{f^2_6f^3_9f_{36}}{f^4_3f^2_{18}}+2q^2\frac{f_6f_{18}f_{36}}{f^3_3}. \end{aligned}$$

The following two theorems are crucial for establishing combinatorial interpretations.

Theorem 4.8

For any fixed n, if

$$\begin{aligned} M_{ped}(0,2,n)&=M_{ped}(1,2,n),\end{aligned}$$
(4.1)
$$\begin{aligned} M_{ped}(0,3,n)&=M_{ped}(1,3,n),\end{aligned}$$
(4.2)
$$\begin{aligned} M_{ped}(0,6,n)+M_{ped}(1,6,n)&=M_{ped}(2,6,n)+M_{ped}(3,6,n), \end{aligned}$$
(4.3)

then

$$\begin{aligned} M_{ped}(m,6,n)=\frac{ped(n)}{6},\ \ m=0,1,2,3,4,5,\ \ \textrm{and}\ \ M_{ped}(3,12,n)=\frac{ped(n)}{12}. \end{aligned}$$

Proof

By (3.2), we have

$$\begin{aligned} M_{ped}(0,2,n)&=M_{ped}(0,6,n)+ 2M_{ped}(2,6,n), \end{aligned}$$
(4.4)
$$\begin{aligned} M_{ped}(1,2,n)&=2M_{ped}(1,6,n)+M_{ped}(3,6,n),\end{aligned}$$
(4.5)
$$\begin{aligned} M_{ped}(0,3,n)&=M_{ped}(0,6,n)+M_{ped}(3,6,n),\end{aligned}$$
(4.6)
$$\begin{aligned} M_{ped}(1,3,n)&=M_{ped}(1,6,n)+M_{ped}(2,6,n). \end{aligned}$$
(4.7)

Substituting (4.4)–(4.7) into (4.1)–(4.2), we have

$$\begin{aligned} M_{ped}(0,6,n)-2M_{ped}(1,6,n)+&2M_{ped}(2,6,n) -M_{ped}(3,6,n)=0,\end{aligned}$$
(4.8)
$$\begin{aligned} M_{ped}(0,6,n)-M_{ped}(1,6,n)-&M_{ped}(2,6,n) +M_{ped}(3,6,n)=0. \end{aligned}$$
(4.9)

Solving system of linear homogeneous equations (4.3), (4.8) and (4.9), we get

$$\begin{aligned} M_{ped}(m,6,n)=\frac{ped(n)}{6},\ \ m=0,1,2,3,4,5. \end{aligned}$$

Moreover,

$$\begin{aligned} M_{ped}(3,6,n)=M_{ped}(3,12,n)+M_{ped}(9,12,n) =2M_{ped}(3,12,n), \end{aligned}$$
(4.10)

which implies

$$\begin{aligned} M_{ped}(3,12,n)=\frac{M_{ped}(3,6,n)}{2}=\frac{ped(n)}{12}. \end{aligned}$$

This completes the proof. \(\square \)

The following theorem can be checked similarly.

Theorem 4.9

For any fixed n, if

$$\begin{aligned} M_{ped}(0,2,n)&=M_{ped}(1,2,n),\\ M_{ped}(0,6,n)+M_{ped}(1,6,n)&=M_{ped}(2,6,n)+M_{ped}(3,6,n), \end{aligned}$$

then

$$\begin{aligned} M_{ped}(0,6,n)&=M_{ped}(3,6,n),\\ M_{ped}(1,6,n)&=M_{ped}(2,6,n). \end{aligned}$$

5 Proofs of main results

In this section, we give proofs of our main results. Hereafter we always assume \(\alpha , n \ge 0\) unless specified otherwise.

Proof of Theorem 1.4

Setting \(z=e^{\pi i}=-1\) in (3.1), we get

$$\begin{aligned} \sum _{n=0}^{\infty } \sum _{m=-\infty }^{\infty }M_{ped}(m,n)(-1)^mq^n =\sum _{n=0}^{\infty }\left( M_{ped}(0,2,n)-M_{ped}(1,2,n)\right) q^n =\frac{f_2^4}{f_1f_4}. \end{aligned}$$
(5.1)

Let g(q) be a polynomial of q, observing that the coefficient of \(q^n\) in g(q) is zero implies the coefficient of \(q^n\) in \(g(-q)\) is zero and vice versa. Hence we consider the following equation. By Lemma 4.1, replacing q by \(-q\) in (5.1), we have

$$\begin{aligned} \sum _{n=0}^{\infty }\left( M_{ped}(0,2,n)-M_{ped}(1,2,n)\right) (-q)^n=f_1f_2. \end{aligned}$$
(5.2)

According to Lemma 4.6, we find that

$$\begin{aligned} \sum _{n=0}^{\infty }\left( M_{ped}(0,2,n)-M_{ped}(1,2,n)\right) (-q)^n= \frac{f_6f^4_9}{f_3f^2_{18}}-qf_9f_{18} -2q^2\frac{f_3f^4_{18}}{f_6f^2_9}.\nonumber \\ \end{aligned}$$
(5.3)

Extracting those terms associated with powers \(q^{3n+1}\) on both sides of (5.3), then dividing by q and replacing \(q^3\) by q, we arrive at

$$\begin{aligned} \sum _{n=0}^{\infty }\left( M_{ped}(0,2,3n+1)- M_{ped}(1,2,3n+1)\right) (-1)^{3n+1}q^n =-f_3f_6. \end{aligned}$$
(5.4)

Since the coefficients of \(q^{3n+1}\) and \(q^{3n+2}\) in (5.4) are both zero, we can conclude that the coefficients of \(q^{9n+4}\) and \(q^{9n+7}\) in (5.2) are both zero. This yields

$$\begin{aligned} M_{ped}(0,2,9n+4)=M_{ped}(1,2,9n+4),\end{aligned}$$
(5.5)
$$\begin{aligned} M_{ped}(0,2,9n+7)=M_{ped}(1,2,9n+7). \end{aligned}$$
(5.6)

Extracting the terms involving \(q^{3n}\) in (5.4) and substituting \(q^3\) by q gives

$$\begin{aligned} \sum _{n=0}^{\infty }\left( M_{ped}(0,2,9n+1)-M_{ped}(1,2,9n+1)\right) (-1)^{9n+1}q^n =-f_1f_2. \end{aligned}$$
(5.7)

Since 9n has the same parity as n, (5.7) becomes

$$\begin{aligned} \sum _{n=0}^{\infty }\left( M_{ped}(0,2,9n+1)- M_{ped}(1,2,9n+1)\right) (-q)^n =f_1f_2. \end{aligned}$$
(5.8)

From (5.2), (5.8) and mathematical induction, it follows that

$$\begin{aligned} \sum _{n=0}^{\infty }\left( M_{ped}\left( 0,2,3^{2\alpha }n+\frac{3^{2\alpha }-1}{8}\right) - M_{ped}\left( 1,2,3^{2\alpha }n+\frac{3^{2\alpha }-1}{8}\right) \right) (-q)^n =f_1f_2. \end{aligned}$$
(5.9)

Comparing (5.9) with (5.2), the following equations can be proved by similar arguments for (5.5)–(5.6), and hence the proof is omitted.

$$\begin{aligned} M_{ped}\left( 0,2,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&=M_{ped}\left( 1,2,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) , \end{aligned}$$
(5.10)
$$\begin{aligned} M_{ped}\left( 0,2,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&=M_{ped}\left( 1,2,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) . \end{aligned}$$
(5.11)

Substituting \(z=e^{\frac{\pi i}{3}}\) into (3.1), by (3.2) and \(e^{\frac{\pi i}{3}}+e^{\frac{5\pi i}{3}}=-(e^{\frac{2\pi i}{3}}+e^{\frac{4\pi i}{3}})=-e^{\pi i}=1\), we see that

$$\begin{aligned}&\sum \limits _{{n = 0}}^{\infty } {\sum \limits _{{m = 0}}^{5} {M_{{ped}} } } (m,6,n)e^{{\frac{{m\pi i}}{3}}} q^{n} \\&\quad = \sum \limits _{{n = 0}}^{\infty } {\left( {M_{{ped}} (0,6,n)} \right. } n + (e^{{\frac{{\pi i}}{3}}} + e^{{\frac{{5\pi i}}{3}}} )M_{{ped}} (1,6,n)n \\&\qquad \left. { + (e^{{\frac{{2\pi i}}{3}}} + e^{{\frac{{4\pi i}}{3}}} )M_{{ped}} (2,6,n) + e^{{\pi i}} M_{{ped}} (3,6,n)} \right) q^{n} n \\&\quad = \sum \limits _{{n = 0}}^{\infty } {\left( {M_{{ped}} (0,6,n) + M_{{ped}} (1,6,n)} \right. } n \left. { - M_{{ped}} (2,6,n) - M_{{ped}} (3,6,n)} \right) q^{n} n \\&\quad = \frac{{f_{2}^{2} f_{4} }}{{f_{1} }}\prod \limits _{{n = 0}}^{\infty } {\frac{1}{{1 - q^{{2n}} + q^{{4n}} }}} n \\&\quad = \frac{{f_{2}^{2} f_{4} }}{{f_{1} }}\prod \limits _{{n = 0}}^{\infty } {\frac{{1 + q^{{2n}} }}{{1 + q^{{6n}} }}} n \\&\quad = \frac{{f_{2}^{2} }}{{f_{1} }}\frac{{f_{4}^{2} }}{{f_{2} }}\frac{{f_{6} }}{{f_{{12}} }} . \end{aligned}$$

By Lemma 4.3, we have

$$\begin{aligned}&\sum _{n=0}^{\infty }\left( M_{ped}(0,6,n)+M_{ped}(1,6,n)- M_{ped}(2,6,n)-M_{ped}(3,6,n)\right) q^n \nonumber \\&\qquad =\left( \frac{f_6f_9^2}{f_3f_{18}}+q\frac{f_{18}^2}{f_9}\right) \left( \frac{f_{12}f_{18}^2}{f_6f_{36}}+q^2\frac{f_{36}^2}{f_{18}}\right) \frac{f_6}{f_{12}}. \end{aligned}$$
(5.12)

Extracting the terms involving \(q^{3n+1}\) in (5.12), then dividing by q and replacing \(q^3\) by q, we find that

$$\begin{aligned}&\sum _{n=0}^{\infty }\left( M_{ped}(0,6,3n+1)+M_{ped}(1,6,3n+1)- M_{ped}(2,6,3n+1)\right. \nonumber \\&\qquad \left. -M_{ped}(3,6,3n+1)\right) q^n =\frac{f_6^4}{f_3f_{12}}. \end{aligned}$$
(5.13)

Obviously, the coefficients of \(q^{3n+1}\) and \(q^{3n+2}\) in (5.13) are both zero, which gives

$$\begin{aligned} M_{ped}(0,6,9n+4)+M_{ped}(1,6,9n+4)=\,&M_{ped}(2,6,9n+4)+M_{ped}(3,6,9n+4),\end{aligned}$$
(5.14)
$$\begin{aligned} M_{ped}(0,6,9n+7)+M_{ped}(1,6,9n+7)=\,&M_{ped}(2,6,9n+7)+M_{ped}(3,6,9n+7). \end{aligned}$$
(5.15)

Considering the terms involving \(q^{3n}\) in (5.13), after simplification, we get

$$\begin{aligned}&\sum _{n=0}^{\infty }\left( M_{ped}(0,6,9n+1)+M_{ped}(1,6,9n+1)- M_{ped}(2,6,9n+1)\right. \nonumber \\&\quad \left. -M_{ped}(3,6,9n+1)\right) q^n =\frac{f_2^4}{f_1f_4}. \end{aligned}$$
(5.16)

Comparing (5.16) with (5.1), according to (5.14)–(5.15) and the proofs of (5.10)–(5.11), a simple deduction shows that

$$\begin{aligned}&M_{ped}\left( 0,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) +M_{ped}\left( 1,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) \nonumber \\&\quad =M_{ped}\left( 2,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) +M_{ped}\left( 3,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) , \end{aligned}$$
(5.17)
$$\begin{aligned}&M_{ped}\left( 0,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) +M_{ped}\left( 1,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) \nonumber \\&\quad =M_{ped}\left( 2,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) +M_{ped}\left( 3,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) . \end{aligned}$$
(5.18)

Combining (5.10)–(5.11), (5.17)–(5.18) and Theorem 4.9, Theorem 1.4 follows immediately. \(\square \)

Proof of Corollary 1.5

Corollary 1.5 can be checked easily by (4.6)–(4.7) and Theorem 1.4, hence we omitted the details. \(\square \)

Proof of Corollary 1.6

By (4.6), (4.10) and Theorem 1.4, one can see that

$$\begin{aligned} M_{ped}\left( 3,12,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right)&= \frac{M_{ped}\left( 3,6,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) }{2}\\&=\frac{M_{ped}\left( 0,3,3^{2\alpha +2}n+\frac{11\cdot 3^{2\alpha +1}-1}{8}\right) }{4},\\ M_{ped}\left( 3,12,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right)&= \frac{M_{ped}\left( 3,6,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) }{2}\\&=\frac{M_{ped}\left( 0,3,3^{2\alpha +2}n+\frac{19\cdot 3^{2\alpha +1}-1}{8}\right) }{4}. \end{aligned}$$

Hence Corollary 1.6 holds.

Proof of Corollary 1.7

By (3.2), we have

$$\begin{aligned} M_{ped}(1,2,n)=M_{ped}(1,4,n)+M_{ped}(3,4,n)=2M_{ped}(1,4,n). \end{aligned}$$
(5.19)

Then Corollary 1.7 follows immediately according to (5.10)–(5.11) and the fact that \(ped(n)=M_{ped}(0,2,n)+M_{ped}(1,2,n)\). \(\square \)

Proof of Theorem 1.8

Substituting \(z=e^\frac{\pi i}{2}=i\) into (3.1), by (3.2) and Lemmas 4.24.3, we find that

$$\begin{aligned} \sum _{n=0}^{\infty }\sum _{m=-\infty }^{\infty }M_{ped}(m,n)i^mq^n&=\sum _{n=0}^{\infty }\sum _{m=0}^{3}M_{ped}(m,4,n)i^mq^n \nonumber \\&=\sum _{n=0}^{\infty }\left( M_{ped}(0,4,n)+(i+i^3)M _{ped}(1,4,n)\right. \nonumber \\&\quad \left. +i^2M_{ped}(2,4,n)\right) q^n\nonumber \\&=\sum _{n=0}^{\infty }\left( M_{ped}(0,4,n)- M_{ped}(2,4,n)\right) q^n\nonumber \\&=\frac{f_2^2}{f_1}\frac{f_4^2}{f_8}\nonumber \\&=\sum _{n=0}^{\infty }q^{n+1 \atopwithdelims ()2}\left( 2\sum _{n=0}^{\infty }(-1)^nq^{4n^2}-1\right) . \end{aligned}$$
(5.20)

By (3.2), one can see that

$$\begin{aligned} ped(n)=\sum _{m=0}^{3}M_{ped}(m,4,n)= M_{ped}(0,4,n)+2M_{ped}(1,4,n)+2M_{ped}(2,8,n). \end{aligned}$$

In light of (5.20) and the fact that \(M_{ped}(0,n)\) has the same parity as \(M_{ped}(0,4,n)\), Theorem 1.8 holds. \(\square \)

We next aim to prove Theorem 1.9.

Proof of Theorem 1.9

Substituting \(z=e^{\frac{2\pi i}{3}}\) into (3.1), we obtain

$$\begin{aligned} \sum _{n=0}^{\infty }\sum _{m=0}^{2}M_{ped}(m,3,n)e^{\frac{2m\pi i}{3}}q^n=\frac{(q^2;q^2)_{\infty }}{(\zeta q^2;q^2)_{\infty }(\zeta ^{-1}q^2;q^2)_{\infty }} \frac{(q^4;q^4)_\infty }{(q;q^2)_\infty } =\frac{f_2^2}{f_1}\frac{f_2f_4}{f_6}. \end{aligned}$$
(5.21)

Using Lemmas 4.34.6, by (3.2) and the fact that \(1+e^{\frac{2\pi i}{3}}+e^{\frac{4\pi i}{3}}=0\), we get

$$\begin{aligned}&\sum \limits _{{n = 0}}^{\infty } {(M_{{ped}} (} 0,3,n) - M_{{ped}} (1,3,n))q^{n} = \left( {\frac{{f_{6} f_{9}^{2} }}{{f_{3} f_{{18}} }} + q\frac{{f_{{18}}^{2} }}{{f_{9} }}} \right) \\&\quad \left( {\frac{{f_{{12}} f_{{18}}^{4} }}{{f_{6} f_{{36}}^{2} }} - q^{2} f_{{18}} f_{{36}} - 2q^{4} \frac{{f_{6} f_{{36}}^{4} }}{{f_{{12}} f_{{18}}^{2} }}} \right) \frac{1}{{f_{6} }} . \end{aligned}$$

Extracting those terms associated with powers \(q^{3n+1}\) on both sides of the above equation, then dividing by q and replacing \(q^3\) by q, one can see that

$$\begin{aligned} \sum _{n=0}^{\infty }(M_{ped}(0,3,3n+1)-M_{ped}(1,3,3n+1))q^n =\frac{f_4}{f_2^2}\frac{f_6^6}{f_3f_{12}^2}- 2q\frac{f_2}{f_1f_4}\frac{f_3^2f_{12}^4}{f_6^3}. \end{aligned}$$
(5.22)

Since

$$\begin{aligned} \frac{f_2}{f_1f_4}=\frac{1}{\psi (-q)}, \end{aligned}$$

using Lemma 4.4,

$$\begin{aligned} \begin{gathered} \sum \limits _{{n = 0}}^{\infty } {(M_{{ped}} (} 0,3,3n + 1) - M_{{ped}} (1,3,3n + 1))q^{n} \\ \quad \quad \quad = \frac{{f_{{12}}^{2} f_{{18}}^{6} }}{{f_{3} f_{6}^{2} f_{{36}}^{3} }} - 2q^{3} \frac{{f_{6} f_{9}^{3} f_{{36}}^{3} }}{{f_{3}^{2} f_{{18}}^{3} }} - 2q\frac{{f_{{12}}^{2} f_{{18}}^{9} }}{{f_{6}^{3} f_{9}^{3} f_{{36}}^{3} }} + 4q^{4} \frac{{f_{{36}}^{3} }}{{f_{3} }} \\ \end{gathered} \end{aligned}$$
(5.23)

after simplification.

Clearly, the coefficient of \(q^{3n+2}\) in (5.23) is zero. We can conclude that

$$\begin{aligned} M_{ped}(0,3,9n+7)=M_{ped}(1,3,9n+7). \end{aligned}$$
(5.24)

Combining (5.11), (5.18), (5.24) and Theorem 4.8, we complete the proof of Theorem 1.9. \(\square \)

Remark 1

Andrews, Hirschhorn, and Sellers [3] presented an interesting infinite family of congruences modulo 3 as given by

$$\begin{aligned} ped\left( 3^{2\alpha +1}n+\frac{17\cdot 3^{2\alpha }-1}{8}\right)&\equiv 0\pmod {3}, \ \ \alpha \ge 1, \end{aligned}$$
(5.25)

and deduced that

$$\begin{aligned} ped\left( 3^{2\alpha +1}n+\frac{17\cdot 3^{2\alpha }-1}{8}\right) \equiv 0\pmod {6}, \ \ \alpha \ge 1. \end{aligned}$$
(5.26)

Actually, based on a substantial amount of numerical evidence, we conjecture that the ped-crank can be used to provide a combinatorial interpretation of (5.25), namely

$$\begin{aligned} M_{ped}{} & {} \left( 0,3,3^{2\alpha +1}n+\frac{17\cdot 3^{2\alpha }-1}{8}\right) =M_{ped}\left( 1,3,3^{2\alpha +1}n+\frac{17\cdot 3^{2\alpha }-1}{8}\right) , \\ {}{} & {} \quad \alpha \ge 1. \end{aligned}$$

Here, we only prove the case for \(\alpha =1\), and for any \(\alpha >1\), we are not able to provide an elementary proof of this conjecture.

Proof

Extracting the terms involving \(q^{3n}\) in (5.23) and substituting \(q^3\) by q, we obtain

$$\begin{aligned} \sum _{n=0}^{\infty }(M_{ped}(0,3,9n+1)-M_{ped}(1,3,9n+1))q^n =\frac{f_4}{f_1}\frac{f_4}{f_2^2}\frac{f_6^6}{f_{12}^3}- 2q\frac{f_2f_3^3f_{12}^3}{f_1^2f_{6}^3}.\nonumber \\ \end{aligned}$$
(5.27)

From Lemmas 4.44.54.7, considering the terms involving \(q^{3n+2}\) in (5.27) leads to

$$\begin{aligned}&\sum _{n=0}^{\infty }(M_{ped}(0,3,27n+19)-M_{ped}(1,3,27n+19))q^n\\&\quad =\frac{f_4f_6^7}{f_1^3f_2f_{12}^2}-\frac{f_3^3f_4^3}{f_1^4} +q\frac{f_2^2f_3^3f_{12}^4}{f_1^4f_4f_6^2}\\&\quad =\frac{f_3^3}{f_1^4}\left( \frac{f_1}{f_3^3} \frac{f_4f_6^7}{f_2f_{12}^2}-f_4^3 +q\frac{f_2^2f_{12}^4}{f_4f_6^2}\right) \\&\quad =\frac{f_3^3}{f_1^4}\left( \left( \frac{f_2f^2_4f^2_{12}}{f^7_{6}}-q\frac{f^3_2f^6_{12}}{f^2_4f^9_6}\right) \frac{f_4f_6^7}{f_2f_{12}^2}-f_4^3 +q\frac{f_2^2f_{12}^4}{f_4f_6^2}\right) \\&\quad =0. \end{aligned}$$

That means

$$\begin{aligned} M_{ped}(0,3,27n+19)=M_{ped}(1,3,27n+19). \end{aligned}$$

\(\square \)

Unfortunately, the ped-crank cannot be employed to interpret (5.26) even for \(\alpha =1\). Hence, it will be interesting to introduce another partition statistic that could combinatorially interpret (5.26).

Remark 2

For ordinary partitions, recall that we define \(M(0,1)=-1,\ \ M(-1,1)=M(1,1)=1\). From the definition of ped-crank, one can see that a similar problem will arise when \(\Delta (\lambda )=((2), \beta , \gamma )\). So we make the following adjustment to the definition of ped-crank. Let

$$\begin{aligned} \gamma&=(\gamma _1,\gamma _2,\cdots ,\gamma _k),\\ A_n&=\{\lambda | \Delta (\lambda )=((2),\beta ,\gamma ),\ \ |\lambda |=n\},\\ B_n&=\{\lambda | \Delta (\lambda )=((0),\beta ,\gamma ),\ \ |\lambda |=n,\ \ \gamma =(2) \ \ \text {or} \ \ \gamma _1-\gamma _2 \ge 4\}. \end{aligned}$$

Definition 5.1

Let \(\lambda \) be a partition of n with even parts distinct. The \(c_{mped}(\lambda )\) is given by

$$\begin{aligned} c_{mped}(\lambda )=\left\{ \begin{array}{ll} 1 &{} if\ \ \lambda \in A_n, \\ -1 &{} if\ \ \lambda \in B_n, \\ c_{ped}(\lambda ) &{} otherwise, \\ \end{array}\right. \end{aligned}$$

where \(c_{ped}(\lambda )\) is the ped-crank of \(\lambda \).

When \(\lambda \in A_n\), an injection from \(A_n\) to \(B_n\) can be constructed by changing \(\alpha \) to \(\emptyset \) and adding 2 to the largest part of \(\gamma \). Another direction is obvious. Hence for any non-negative integer n, there is a bijection between \(A_n\) and \(B_n\). Let \(M_{mped}(m,n)\) denote the number of partitions of n with even parts distinct with \(c_{mped}(\lambda )=m\). By the definition of \(c_{mped}(\lambda )\), one can check \(M_{mped}(m,n)=M_{ped}(m,n)\) for any integer m and non-negative integer n.

For example, if \(\lambda =(6,5,4,1,1,1),\Delta (\lambda )=((2),(5,1),(6,4))\), then \(c_{mped}(\lambda )=1\) and if \(\lambda =(8,5,4,1),\Delta (\lambda )=((0),(5,1),(8,4))\), then \(c_{mped}(\lambda )=-1\).

Table 1 The case for \(n=7\)

Table 1 gives the 12 partitions of 7 with even parts distinct. It is easy to check that these partitions are divided into six equinumerous subsets by ped-crank. Moreover,

$$\begin{aligned} M_{ped}(1,4,7)=3=\frac{ped(7)}{4},\\ M_{ped}(3,12,7)=1=\frac{ped(7)}{12}. \end{aligned}$$

6 Closing remarks

In 2014, Xia [12] proved the following congruence modulo 4 for ped(n).

Theorem 6.1

[12, Equation (9), Theorem 1] For \(\alpha ,n\ge 0\),

$$\begin{aligned} ped\left( 3^{2\alpha }n+\frac{3^{2\alpha }-1}{8}\right) \equiv ped(n) \pmod {4}. \end{aligned}$$

Note that comparing (5.4) with (5.13), a simple deduction gives

$$\begin{aligned} M_{ped}(1,6,3n+1)=M_{ped}(2,6,3n+1). \end{aligned}$$

Thus

$$\begin{aligned} ped(3n+1)=M_{ped}(0,3,3n+1)+4M_{ped}(1,6,3n+1). \end{aligned}$$

Since for all \(\alpha >0,n\ge 0\), \(3^{2\alpha }n+\frac{3^{2\alpha }-1}{8}\equiv 1\pmod {3}\), a refinement of Theorem 6.1 can be given as

$$\begin{aligned} M_{ped}\left( 0,3,3^{2\alpha }n+\frac{3^{2\alpha }-1}{8}\right) \equiv ped(n) \pmod {4}. \end{aligned}$$