1 Introduction

A partition of a positive integer n is a non-increasing sequence of positive integers called parts whose sum is n. For convenience, we use the following notation

$$\begin{aligned} (a;q)_\infty =\prod _{n=0}^\infty (1-aq^n), \quad \text{ and }\quad f_k=(q^k;q^k)_\infty . \end{aligned}$$

A partition \(\lambda \) of a positive integer n is said to be t-core if it has no hook numbers that are multiples of t. Let \(a_t(n)\) denote the number of t-core partitions of n. In [7, Eq. (2.1)], the generating function of \(a_t(n)\) is given by

$$\begin{aligned} \sum _{n=0}^\infty a_t(n)q^n=\frac{f_t^t}{f_1}. \end{aligned}$$
(1.1)

A k-tuple partition \((\lambda _1, \lambda _2,\ldots , \lambda _k)\) of n is a k-tuple of partitions \((\lambda _1, \lambda _2,\ldots , \lambda _k)\) such that the sum of all the parts equals n. A k-tuple partition \((\lambda _1, \lambda _2,\ldots , \lambda _k)\) of n with t-cores means that each \(\lambda _i\) is t-core for \(i=1,2,\cdots ,k\). Let \(A_{t,k}(n)\) denote the number of k-tuple partitions of n with t-cores. The generating function of \(A_{t,k}(n)\) is

$$\begin{aligned} \sum _{n=0}^\infty A_{t,k}(n)q^n=\frac{f_t^{kt}}{f_1^k}. \end{aligned}$$
(1.2)

Many authors have studied arithmetic properties of \(A_{3,2}(n)\) and obtained some Ramanujan-type congruences. Lin [11] established some infinite families of congruences modulo 4, 5, 7 and 8 for \(A_{3,2}(n)\). Based on Lin’s study, Xia [18] added proofs of several infinite families of congruences modulo 4, 8 and \(\frac{4^k-1}{3}\) \((k\ge 2)\) for \(A_{3,2}(n)\). For more results about \(A_{3,2}(n)\), see [2, 5, 17, 18, 21].

Wang [16] proved some infinite families of identities and congruences for \(A_{3,3}(n)\) by using some identities of q-series. In [17], Wang established some explicit formulas for \(A_{3,2}(n)\) and \(A_{3,3}(n)\). After that, Chern [5] extended the work of Wang [17] and studied some arithmetic identities of \(A_{t,k}(n)\) for \((t,k)=(3,4),(3,6),(4,2)\),(5, 1),(5, 2) and (7, 1) by applying the theory of modular form.

Saikia and Boruah [13] proved some infinite families of congruences modulo 2, 3 for \(A_{4,2}(n)\) and \(A_{4,3}(n)\). Dasappa [6] discovered a nice congruence modulo \(5^\alpha \) \((\alpha \ge 1)\) for \(A_{5,2}(n)\). Saikia and Boruah [14] also studied arithmetic properties of \(A_{5,2}(n)\) and proved some congruences modulo 2 and 5. In sequel, they [15] established some Ramanujan-type congruences for \(A_{t,k}(n)\) when \((t,k)=(3,4),(3,9),(4,8),(5,6),(8,4),(9,3)\) and (9, 6) by employing q-series identities.

Zou [22] proved some congruences modulo 2 for \(A_{t,2}(n)\); t is a prime such that \(7\le t\le 23\). In 2020, Naika and Nayaka [12] established some Ramanujan-type congruences modulo 5, 7 and 8 for \(A_{t,4}(n)\), \(t=3, 5, 7, 25\).

In this paper, we mainly study arithmetic properties of \(A_{3,5}(n)\). Its generating function is given by

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(n)q^n=\frac{f_3^{15}}{f_1^5}. \end{aligned}$$
(1.3)

We establish some results about congruences modulo 2, 4, 5, 8 and 10 for \(A_{3,5}(n)\).

To be specific, by using some dissection formulae, we obtain some infinite families of congruences modulo 2 for \(A_{3,5}(n)\) as follows.

Theorem 1.1

For any integer \(n\ge 0\), we have

$$\begin{aligned} A_{3,5}\left( 8n+5\right) \equiv 0\pmod {2}. \end{aligned}$$
(1.4)

and for \(\alpha \ge 1\),

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +3}n+\frac{7\cdot 2^{2\alpha +1}-5}{3}\right)&\equiv 0\pmod {2}, \end{aligned}$$
(1.5)
$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +4}n+\frac{13\cdot 2^{2\alpha +1}-5}{3}\right)&\equiv 0\pmod {2}. \end{aligned}$$
(1.6)

Theorem 1.2

For any prime \(p\ge 5\), \(\alpha \ge 1\), \(\beta \ge 0\), and \(n\ge 0\),

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}p^{2\beta }n+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^n\equiv f(-q^4)\pmod {2}. \end{aligned}$$
(1.7)

We deduce the following infinite families of congruences with two parameters \(\alpha , \beta \) modulo 2 for \(A_{3,5}(n)\).

Corollary 1.1

For any prime \(p\ge 5\), \(\alpha \ge 1\), \(\beta \ge 0\), if n cannot be represented as \(2k(3k+1)\) for some integer k, then

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +2}p^{2\beta }n+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) \equiv 0 \pmod {2}. \end{aligned}$$
(1.8)

and for any integer \(n\ge 0\),

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +2}p^{2\beta +2}n+\frac{(6j+p)\cdot 2^{ 2\alpha +1}p^{2\beta +1}-5}{3}\right)&\equiv 0\pmod {2}, \quad j\in \{1,2\cdots p-1\}. \end{aligned}$$
(1.9)
$$\begin{aligned} A_{3,5}\left( 7^2\cdot 2^{2\alpha +2}p^{2\beta }n+\frac{(42k+7)\cdot 2^{2\alpha +1}p^{2\beta }-5}{3}\right)&\equiv 0\pmod {2}, \quad k\in \{0,2,3,4,5,6\}. \end{aligned}$$
(1.10)

and

$$\begin{aligned} A_{3,5}\left( 13^2\cdot 2^{2\alpha +2} p^{2\beta }n+\frac{(78k+13)\cdot 2^{2\alpha +1} p^{2\beta }-5}{3}\right)&\equiv 0\pmod {2}, \nonumber \\ \quad k\in \{0,1,3,4,5,6,7,8,9,10,11,12\}. \end{aligned}$$
(1.11)

We find the following congruences modulo 5 and 10 for \(A_{3,5}(n)\) hold.

Theorem 1.3

For \(n\ge 0\), we have

$$\begin{aligned} A_{3,5}(4n+3)&\equiv 0 \pmod {10}, \end{aligned}$$
(1.12)
$$\begin{aligned} A_{3,5}(5n+k)&\equiv 0 \pmod {5}, \quad k\in \{1, 2, 3 ,4\}. \end{aligned}$$
(1.13)
$$\begin{aligned} A_{3,5}(20n+15)&\equiv 0 \pmod {5}, \end{aligned}$$
(1.14)

and for \(\alpha \ge 0\), we get

$$\begin{aligned} A_{3,5}\left( 5\cdot 2^{2\alpha }n+\frac{5\cdot 2^{2\alpha }-5}{3}\right)&\equiv A_{3,5}(5n) \pmod {5}, \end{aligned}$$
(1.15)
$$\begin{aligned} A_{3,5}\left( 5\cdot 2^{2\alpha +2}n+\frac{5^2\cdot 2^{2\alpha +1}-5}{3}\right)&\equiv 0 \pmod {5}. \end{aligned}$$
(1.16)

We also establish the following congruences modulo 4, 8 for \(A_{3,5}(n)\).

Theorem 1.4

For \(\alpha \ge 2\) and \(n\ge 0\), we have

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +1}n+\frac{5\cdot 2^{2\alpha }-5}{3}\right) \equiv 0\pmod {4}. \end{aligned}$$
(1.17)

Let \(p\ge 3\) be a prime and a be an integer. The Legendre symbol is defined by

$$\begin{aligned} \left( \frac{a}{p}\right) ={\left\{ \begin{array}{ll} 1,&{}\ \text{ if } a ~\text{ is } \text{ a } \text{ quadratic } \text{ residue } \text{ modulo } p~ \text{ and }~ a\not \equiv 0\pmod {p},\\ 0,&{}\ \text{ if }~ a\equiv 0\pmod {p},\\ -1,&{}\ \text{ if }~ a ~\text{ is } \text{ a } \text{ quadratic } \text{ non-residue } \text{ modulo } p. \end{array}\right. } \end{aligned}$$

Theorem 1.5

For any prime \(p\ge 5\) such that \(\left( \frac{-6}{p}\right) =-1\), \(\alpha \ge 0\), and \(n\ge 0\),

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{14\cdot p^{2\alpha }-5}{3}\right) q^n \equiv 2f(-q)\psi (q^2)\pmod {4}. \end{aligned}$$
(1.18)

Corollary 1.2

For any prime \(p\ge 5\) such that \(\left( \frac{-6}{p}\right) =-1\), \(\alpha \ge 0\), if n cannot be represented as the sum of a pentagonal number and twice a triangular number, then

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{14\cdot p^{2\alpha }-5}{3}\right) \equiv 0 \pmod {4}. \end{aligned}$$
(1.19)

and for any integer \(n\ge 0\),

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha +2}n+\frac{(48j+14p)\cdot p^{2\alpha +1}-5}{3}\right) \equiv 0\pmod {4}, \end{aligned}$$
(1.20)

where \(1\le j\le p-1.\)

Theorem 1.6

For any prime \(p\ge 5\) such that \(\left( \frac{-18}{p}\right) =-1\), \(\alpha \ge 0\), and \(n\ge 0\),

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{38\cdot p^{2\alpha }-5}{3}\right) q^n\equiv 2f(-q)\psi (q^6)\pmod {4}. \end{aligned}$$
(1.21)

Corollary 1.3

For any prime \(p\ge 5\) such that \(\left( \frac{-18}{p}\right) =-1\), \(\alpha \ge 0\), if n cannot be represented as the sum of a pentagonal number and six times a triangular number, then

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{38\cdot p^{2\alpha }-5}{3}\right) \equiv 0 \pmod {4}. \end{aligned}$$
(1.22)

and for any integer \(n\ge 0\),

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha +2}n+\frac{(48j+38p)\cdot p^{2\alpha +1}-5}{3}\right) \equiv 0\pmod {4}, \end{aligned}$$
(1.23)

where \(1\le j\le p-1\).

Theorem 1.7

For \(n\ge 0\), we have

$$\begin{aligned} A_{3,5}(4n+2)&\equiv 0\pmod {4}, \end{aligned}$$
(1.24)
$$\begin{aligned} A_{3,5}(16n+13)&\equiv 0 \pmod {8}. \end{aligned}$$
(1.25)

Theorem 1.8

For any prime \(p\ge 5\) such that \(\left( \frac{-9}{p}\right) =-1\), \(\alpha \ge 0\), and \(n\ge 0\),

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2\cdot p^{2\alpha }n+\frac{5\cdot p^{2\alpha }-5}{3}\right) q^n \equiv (-1)^{\alpha (\frac{\pm p-1}{6})}f(-q^2)\psi (q^6)\pmod {4}. \end{aligned}$$
(1.26)

Corollary 1.4

For any prime \(p\ge 5\) such that \(\left( \frac{-9}{p}\right) =-1\), \(\alpha \ge 0\), if n cannot be represented as the sum of a pentagonal number and three times a triangular number, then

$$\begin{aligned} A_{3,5}\left( 4\cdot p^{2\alpha }n+\frac{5\cdot p^{2\alpha }-5}{3}\right) \equiv 0 \pmod {4}. \end{aligned}$$
(1.27)

and for any integer \(n\ge 0\),

$$\begin{aligned} A_{3,5}\left( 2\cdot p^{2\alpha +2}n+\frac{(6j+5p)\cdot p^{2\alpha +1}-5}{3}\right) \equiv 0\pmod {4}, \end{aligned}$$
(1.28)

where \(1\le j\le p-1\).

Theorem 1.9

For any prime \(p\ge 5\) such that \(\left( \frac{-9}{p}\right) =-1\), \(\alpha \ge 0\) and \(n\ge 0\),

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 8\cdot p^{2\alpha }n+\frac{20\cdot p^{2\alpha }-5}{3}\right) q^n \equiv 6(-1)^{\alpha (\frac{\pm p-1}{6})}f(-q^2)\psi (q^6)\pmod {8}. \end{aligned}$$
(1.29)

Corollary 1.5

For any prime \(p\ge 5\) such that \(\left( \frac{-9}{p}\right) =-1\), \(\alpha \ge 0\), if n cannot be represented as the sum of a pentagonal number and three times a triangular number, then

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{20\cdot p^{2\alpha }-5}{3}\right) \equiv 0 \pmod {8}. \end{aligned}$$
(1.30)

and for any integer \(n\ge 0\),

$$\begin{aligned} A_{3,5}\left( 8\cdot p^{2\alpha +2}n+\frac{(24j+20p)\cdot p^{2\alpha +1}-5}{3}\right) \equiv 0\pmod {8}, \end{aligned}$$
(1.31)

where \(1\le j\le p-1\).

This paper is organized as follows. In Sect. 2, we shall prove some theorems about congruences for \(A_{3,5}(n)\) modulo 2, 5 and 10. In Sect. 3, we give the proofs of remaining theorems about congruences modulo 4, 8 for \(A_{3,5}(n)\).

2 Proofs of Theorems 1.11.3 and Corollary 1.1

Proof of Theorem 1.1

In order to prove Theorem 1.1, we first prove the following lemma.

Lemma 2.1

For \(\alpha \ge 1\) and \(n\ge 0\), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha }n+\frac{2^{2\alpha +1}-5}{3}\right) q^n\equiv \frac{f_3^6}{f_1^2}\pmod {2}. \end{aligned}$$
(2.1)

Proof

Hirschhorn, Garvan and Borwein [9] proved that

$$\begin{aligned} \frac{f_3^3}{f_1}=\frac{f_4^3f_6^2}{f_2^2f_{12}}+q\frac{f_{12}^3}{f_4}. \end{aligned}$$
(2.2)

By the binomial theorem, for any positive integer k and any prime p,

$$\begin{aligned} f_1^{p^k}\equiv f_p^{p^{k-1}} \pmod {p^k}. \end{aligned}$$
(2.3)

Substituting (2.2) into (1.3) and employing (2.3), we obtain that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(n)q^n&=\frac{f_3^{12}}{f_1^4}\frac{f_3^3}{f_1} \equiv \frac{f_6^{6}}{f_2^2}\left( \frac{f_4^3f_6^2}{f_2^2f_{12}}+q\frac{f_{12}^3}{f_4}\right) \nonumber \\&\equiv \frac{f_4^3f_6^8}{f_2^4f_{12}}+q\frac{f_6^6f_{12}^3}{f_2^2f_4}\pmod {2}. \end{aligned}$$
(2.4)

Extracting the terms involving \(q^{2n+1}\) from both sides of (2.4) , dividing by q and replacing \(q^2\) by q, then employing (2.3), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(2n+1)q^n\equiv \frac{f_3^6f_6^3}{f_1^2f_2}\equiv \frac{f_6^6}{f_2^2}\pmod {2}. \end{aligned}$$
(2.5)

Hence,

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(4n+1)q^n \equiv \frac{f_3^6}{f_1^2} \pmod {2}, \end{aligned}$$
(2.6)

which is the case \(\alpha =1\) of (2.1).

Suppose that (2.1) holds for \(\alpha \ge 1\). Utilizing (2.2), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha }n+\frac{2^{2\alpha +1}-5}{3}\right) q^n&\equiv \frac{f_3^6}{f_1^2}\equiv \left( \frac{f_4^3f_6^2}{f_2^2f_{12}}+q\frac{f_{12}^3}{f_4}\right) ^2\nonumber \\&\equiv \frac{f_4^6f_6^4}{f_2^4f_{12}^2}+q^2\frac{f_{12}^6}{f_4^2}\pmod {2}. \end{aligned}$$
(2.7)

Hence,

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +1}n+\frac{2^{2\alpha +1}-5}{3}\right) q^n \equiv \frac{f_2^6f_3^4}{f_1^4f_{6}^2}+q\frac{f_{6}^6}{f_2^2}\equiv f_8+q\frac{f_6^6}{f_2^2}\pmod {2}. \end{aligned}$$
(2.8)

Extracting the terms involving \(q^{2n+1}\) from both sides of (2.8), dividing by q and replacing \(q^2\) by q, we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}n+\frac{2^{2\alpha +3}-5}{3}\right) q^n \equiv \frac{f_3^6}{f_1^2}\pmod {2}, \end{aligned}$$
(2.9)

which is the case \(\alpha +1\) of (2.1). The proof of Lemma 2.1 is completed by induction on \(\alpha \).

Now, we turn to prove Theorem 1.1. It follows from (2.3) that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(4n+1)q^n \equiv \frac{f_3^6}{f_1^2}\equiv \frac{f_6^3}{f_2}\pmod {2}, \end{aligned}$$
(2.10)

which implies (1.4).

Extracting the terms involving \(q^{2n}\) from both sides of (2.8) and replacing \(q^2\) by q, we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}n+\frac{2^{2\alpha +1}-5}{3}\right) q^n\equiv f_4 \pmod {2}. \end{aligned}$$
(2.11)

which implies that

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +2}(2n+1)+\frac{2^{2\alpha +1}-5}{3}\right)&\equiv 0\pmod {2}, \end{aligned}$$
(2.12)
$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +2}(4n+2)+\frac{2^{2\alpha +1}-5}{3}\right)&\equiv 0 \pmod {2}. \end{aligned}$$
(2.13)

We thus obtain (1.5) and (1.6). This completes the proof of Theorem 1.1. \(\square \)

Proof of Theorem 1.2 and Corollary 1.1.

We first recall that Ramanujan’s general theta function f(ab) is defined by

$$\begin{aligned} f(a,b)=\sum _{n=-\infty }^\infty a^{\frac{n(n+1)}{2}}b^{\frac{n(n-1)}{2}},\quad |ab|<1. \end{aligned}$$
(2.14)

Two important cases of f(ab) are the theta functions \(\psi (q)\) and \(f(-q)\) [1, p.36. Entry 22], which are given by

$$\begin{aligned} \psi (q)&:=f(q,q^3)=\sum _{n=0}^{\infty }q^{\frac{n(n+1)}{2}}=\frac{f_2^2}{f_1}, \end{aligned}$$
(2.15)
$$\begin{aligned} f(-q)&:=f(-q,-q^2)=\sum _{n=-\infty }^{\infty }(-1)^nq^{\frac{n(3n+1)}{2}}=f_1. \end{aligned}$$
(2.16)

Cui and Gu [4] proved the following p-dissection identities for \(\psi (q)\) and \(f(-q)\).

Lemma 2.2

[4, Theorem 2.1] For any odd prime p,

$$\begin{aligned} \psi (q)=\sum _{k=0}^{\frac{p-3}{2}}q^{\frac{k^2+k}{2}}f(q^{\frac{p^2+(2k+1)p}{2}},q^{\frac{p^2-(2k+1)p}{2}})+ q^{\frac{p^{2}-1}{8}}\psi (q^{p^2}), \end{aligned}$$
(2.17)

Furthermore, for \(0\le k\le \frac{p-3}{2},\)

$$\begin{aligned} \frac{k^2+k}{2}\not \equiv \frac{p^2-1}{8} \pmod {p}. \end{aligned}$$

Lemma 2.3

[4, Theorem 2.2] For any prime \(p\ge 5\),

$$\begin{aligned} f(-q)= & {} \sum _{k=-\frac{p-1}{2},k\ne \frac{\pm p-1}{6}}^{\frac{p-1}{2}}(-1)^kq^{\frac{3k^2+k}{2}}f(-q^{\frac{3p^2+(6k+1)p}{2}},-q^{\frac{3p^2-(6k+1)p}{2}})\nonumber \\&\quad +(-1)^{\frac{\pm p-1}{6}}q^{\frac{p^2-1}{24}}f(-q^{p^2}), \end{aligned}$$
(2.18)

where

$$\begin{aligned} \frac{\pm p-1}{6}:=\left\{ \begin{array}{ll} \frac{p-1}{6}, &{}\quad \hbox {if}\ p\equiv 1\pmod {6}, \\ \frac{-p-1}{6}, &{}\quad \hbox {if}\ p\equiv -1\pmod {6}. \end{array} \right. \end{aligned}$$

Furthermore, for \(-\frac{p-1}{2}\le k\le \frac{p-1}{2}\) and \(k\ne \frac{\pm p-1}{6}\),

$$\begin{aligned} \frac{3k^2+k}{2}\not \equiv \frac{p^2-1}{24} \pmod {p}. \end{aligned}$$

It follows from (2.11) and (2.16) that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}n+\frac{2^{2\alpha +1}-5}{3}\right) q^n\equiv f(-q^4) \pmod {2}, \end{aligned}$$
(2.19)

which is the case \(\beta =0\) of (1.7).

Now, suppose that (1.7) holds for \(\beta \ge 0\). Invoking Lemma 2.3, we get that for any prime \(p\ge 5\),

$$\begin{aligned}&\sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}p^{2\beta }(pn+\frac{p^2-1}{6})+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^n\nonumber \\&\quad = \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}p^{2\beta +1}n+\frac{2^{2\alpha +1}p^{2\beta +2}-5}{3}\right) q^n \equiv f(-q^{4p})\pmod {2}, \end{aligned}$$
(2.20)

which implies that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}p^{2\beta +2}n+\frac{2^{2\alpha +1}p^{2\beta +2}-5}{3}\right) q^n\equiv f(-q^4)\pmod {2}, \end{aligned}$$
(2.21)

which is the case \(\beta +1\) of (1.7). The proof of Theorem 1.2 is completed by induction on \(\beta \).

Combining (1.7) and (2.16), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}p^{2\beta }n+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^n \equiv \sum _{k=-\infty }^{\infty }(-1)^kq^{2k(3k+1)} \pmod {2}, \end{aligned}$$
(2.22)

which implies (1.8).

Applying (2.20) yields that for \(j=1,2,\ldots ,p-1\),

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha +2}p^{2\beta +1}(pn+j)+\frac{2^{2\alpha +1}p^{2\beta +2}-5}{3}\right) \equiv 0\pmod {2}, \end{aligned}$$

which implies (1.9).

Following Hirschhorn [10], for a positive integer k and a power series \(\sum _{n=0}^\infty a(n)q^n\), we introduce an operator \(H_k\), which acts on series of (positive and negative) powers of a single variable and picks out those terms in which the power is congruent to 0 modulo k. That is

$$\begin{aligned} H_k\left( \sum _{n=0}^\infty a(n)q^n\right) :=\sum _{n=0}^\infty a(kn)q^{kn}. \end{aligned}$$
(2.23)

Garvan [8] proved the following results. Let

$$\begin{aligned} \xi =\frac{f(-q)}{q^2f(-q^{49})} \quad \mathrm{and} \quad T=\frac{f(-q^7)^4}{q^7f(-q^{49})^4}. \end{aligned}$$
(2.24)

We have

$$\begin{aligned}&H_7(\xi )=-1, \quad \quad \quad \quad H_7(\xi ^2)=1, \quad \quad \quad \quad H_7(\xi ^3)=-7,\nonumber \\&H_7(\xi ^4)=-4T-7, \quad H_7(\xi ^5)=10T+49, \quad H_7(\xi ^6)=49. \end{aligned}$$
(2.25)

Applying (2.3) and (2.24) in Theorem 1.2, for any prime \(p\ge 5\), \(\alpha \ge 1\) and \(\beta \ge 0\), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2} p^{2\beta }n+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^n\equiv f(-q)^4 \equiv q^8f(-q^{49})^4\xi ^4\pmod {2}. \end{aligned}$$
(2.26)

Then, we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2} p^{2\beta }(7n+1)+\frac{2^{2\alpha +1} p^{2\beta }-5}{3}\right) q^{7n+1}&\equiv q^8f(-q^{49})^4H_7(\xi ^4)\nonumber \\&\equiv q^8f(-q^{49})^4\pmod {2}. \end{aligned}$$
(2.27)

      Furthermore,

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 7\cdot 2^{2\alpha +2} p^{2\beta }n+\frac{7\cdot 2^{2\alpha +1} p^{2\beta }-5}{3}\right) q^n\equiv qf(-q^{7})^4\pmod {2}. \end{aligned}$$
(2.28)

Therefore, for \(k\in \{0,2,3,4,5,6\}\), we get

$$\begin{aligned} A_{3,5}\left( 7\cdot 2^{2\alpha +2} p^{2\beta }(7n+k)+\frac{7\cdot 2^{2\alpha +1} p^{2\beta }-5}{3}\right) \equiv 0\pmod {2}, \end{aligned}$$
(2.29)

which implies (1.10).

Hirschhorn [11] stated the following results. Denote

$$\begin{aligned} \zeta =\frac{f(-q)}{q^7f(-q^{169})} \quad and \quad S=\frac{f(-q^{13})^2}{q^{13}f(-q^{169})^2}. \end{aligned}$$
(2.30)

We have

$$\begin{aligned}&H_{13}(\zeta )=1,\quad \quad \quad H_{13}(\zeta ^2)=-2S-1, \nonumber \\&H_{13}(\zeta ^3)=13, \quad \quad H_{13}(\zeta ^4)=2S^2-13. \end{aligned}$$
(2.31)

Applying (2.3) and (2.30) in Theorem 1.2, for any prime \(p\ge 5\), \(\alpha \ge 1\) and \(\beta \ge 0\), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2} p^{2\beta }n+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^n\equiv f(-q)^4 \equiv q^{28}f(-q^{169})^4\zeta ^4\pmod {2}. \end{aligned}$$
(2.32)

Then,

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}p^{2\beta }(13n+2)+\frac{2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^{13n+2}&\equiv q^{28}f(-q^{169})^4H_{13}(\zeta ^4)\nonumber \\&\equiv q^{28}f(-q^{169})^4\pmod {2}. \end{aligned}$$
(2.33)

      Furthermore, we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 13\cdot 2^{2\alpha +2}p^{2\beta }n+\frac{13\cdot 2^{2\alpha +1}p^{2\beta }-5}{3}\right) q^n\equiv q^2f(-q^{13})^4\pmod {2}. \end{aligned}$$
(2.34)

Therefore, for \(k\in \{0,1,3,4,5,6,7,8,9,10,11,12\}\),

$$\begin{aligned} A_{3,5}\left( 13\cdot 2^{2\alpha +2}p^{2\beta }(13n+k)+\frac{13\cdot 2^{2\alpha +1}p^{2\beta }-5}{3}\right) \equiv 0\pmod {2}, \end{aligned}$$
(2.35)

which implies (1.11). This completes the proof of Theorem 1.2 and Corollary 1.1.

Proof of Theorem 1.3

Using (2.2) in (1.3), we obtain

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(n)q^n&=\frac{f_3^{15}}{f_1^5}=\left( \frac{f_3^3}{f_1}\right) ^5 =\left( \frac{f_4^3f_6^2}{f_2^2f_{12}}+q\frac{f_{12}^3}{f_4}\right) ^5\nonumber \\&\equiv \frac{f_4^{15}f_6^{10}}{f_2^{10}f_{12}^5}+5q\frac{f_4^{11}f_6^8}{f_2^8f_{12}} +5q^4\frac{f_6^2f_{12}^{11}}{f_2^2f_4}+q^5\frac{f_{12}^{15}}{f_4^5} \pmod {10}. \end{aligned}$$
(2.36)

Extracting the terms involving \(q^{2n+1}\) from both sides of (2.36), dividing by q and replacing \(q^2\) by q, then employing (2.3), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(2n+1)q^n\equiv 5\frac{f_2^{11}f_3^8}{f_1^8f_{6}}+q^2\frac{f_{6}^{15}}{f_2^5} \equiv 5f_2^7f_6^3+q^2\frac{f_{6}^{15}}{f_2^5}\pmod {10}, \end{aligned}$$
(2.37)

which implies (1.12).

Utilizing (2.3) in (1.3), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(n)q^n=\frac{f_3^{15}}{f_1^5}\equiv \frac{f_{15}^3}{f_5} \pmod {5}. \end{aligned}$$
(2.38)

This yields that for \(k\in \{1,2,3,4\}\), \(A_{3,5}(5n+k)\equiv 0 \pmod {5}\), which is (1.13).

Combining (2.2) and (2.38), we have that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(5n)q^n\equiv \frac{f_{3}^3}{f_1}=\frac{f_4^3f_6^2}{f_2^2f_{12}}+q\frac{f_{12}^3}{f_4} \pmod {5}. \end{aligned}$$
(2.39)

Extracting the terms involving \(q^{2n+1}\) from both sides of (2.39), dividing by q and replacing \(q^2\) by q, we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(10n+5)q^n\equiv \frac{f_{6}^3}{f_2} \pmod {5}. \end{aligned}$$
(2.40)

Then, we have

$$\begin{aligned} A_{3,5}(20n+15)\equiv 0 \pmod {5}, \end{aligned}$$
(2.41)

which is (1.14). And we obtain

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(20n+5)q^n\equiv \frac{f_{3}^3}{f_1} \pmod {5}. \end{aligned}$$
(2.42)

In view of (2.39) and (2.42), we have

$$\begin{aligned} A_{3,5}(20n+5)\equiv A_{3,5}(5n) \pmod {5}. \end{aligned}$$
(2.43)

      Utilizing (2.43) and by mathematical induction on \(\alpha \), we get (1.15). Combining (2.41) and (1.15), we thus arrive at (1.16). This completes the proof of Theorem 1.3. \(\square \)

3 Proofs of Theorems 1.41.9 and Corollaries 1.21.5

In this section, we prove the remaining theorems and corollaries about congruences modulo 4, 8 for \(A_{3,5}(n)\).

Proof of Theorem 1.4

In order to prove Theorem 1.4, we first prove the following lemma.

Lemma 3.1

For \(\alpha \ge 2\) and \(n\ge 0\), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha }n+\frac{2^{2\alpha +1}-5}{3}\right) q^n\equiv -\frac{f_6^3}{f_2}\pmod {4}. \end{aligned}$$
(3.1)

Proof

Baruah and Ojah [3] established that

$$\begin{aligned} \frac{1}{f_1f_3}&=\frac{f_8^2f_{12}^5}{f_2^2f_4f_6^4f_{24}^2}+q\frac{f_4^5f_{24}^2}{f_2^4f_6^2f_8^2f_{12}}. \end{aligned}$$
(3.2)

      Using (3.2) in (1.3), we obtain

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(n)q^n&=\frac{f_3^{15}}{f_1^5}=\frac{f_3^{16}}{f_1^4}\frac{1}{f_1f_3}\nonumber \\&\equiv \frac{f_6^8}{f_2^2}\left( \frac{f_8^2f_{12}^5}{f_2^2f_4f_6^4f_{24}^2}+q\frac{f_4^5f_{24}^2}{f_2^4f_6^2f_8^2f_{12}}\right) \nonumber \\&\equiv \frac{f_6^4f_8^2f_{12}^5}{f_2^4f_4f_{24}^2}+q\frac{f_4^5f_6^6f_{24}^2}{f_2^6f_8^2f_{12}}\pmod {4}. \end{aligned}$$
(3.3)

Xia and Yao [19] proved the following identity

$$\begin{aligned} \frac{f_3^2}{f_1^2}=\frac{f_4^4f_6f_{12}^2}{f_2^5f_8f_{24}}+2q\frac{f_4f_6^2f_8f_{24}}{f_2^4f_{12}}. \end{aligned}$$
(3.4)

Extracting the terms involving \(q^{2n+1}\) from both sides of (3.3), dividing by q and replacing \(q^2\) by q, then employing (3.4), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(2n+1)q^n&\equiv \frac{f_2^5f_3^6f_{12}^2}{f_1^6f_4^2f_6} \equiv \frac{f_2^5f_{12}^2}{f_4^2f_6}\left( \frac{f_{3}^2}{f_1^2}\right) ^3\nonumber \\&\equiv \frac{f_2^5f_{12}^2}{f_4^2f_6}\left( \frac{f_4^4f_6f_{12}^2}{f_2^5f_8f_{24}}+2q\frac{f_4f_6^2f_8f_{24}}{f_2^4f_{12}}\right) ^3\nonumber \\&\equiv \frac{f_4^{10}f_6^2f_{12}^8}{f_2^{10}f_8^3f_{24}^3} +2q\frac{f_4^7f_6^3f_{12}^5}{f_2^9f_8f_{24}}\pmod {4}. \end{aligned}$$
(3.5)

Extracting the terms involving \(q^{2n}\) from both sides of (3.5), replacing \(q^2\) by q and using (3.4), we obtain

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(4n+1)q^n&\equiv \frac{f_2^{10}f_3^2f_{6}^8}{f_1^{10}f_4^3f_{12}^3} \equiv \frac{f_2^6f_6^8}{f_4^3f_{12}^3}\frac{f_3^2}{f_1^2}\nonumber \\&\equiv \frac{f_2^6f_6^8}{f_4^3f_{12}^3}\left( \frac{f_4^4f_6f_{12}^2}{f_2^5f_8f_{24}}+2q\frac{f_4f_6^2f_8f_{24}}{f_2^4f_{12}}\right) \nonumber \\&\equiv \frac{f_2f_4f_6^9}{f_8f_{12}f_{24}}+2q\frac{f_2^2f_6^{10}f_8f_{24}}{f_4^2f_{12}^4}\pmod {4}. \end{aligned}$$
(3.6)

Xia and Yao [20] proved the following 2-dissection formula.

$$\begin{aligned} f_1f_3=\frac{f_2f_8^2f_{12}^4}{f_4^2f_6f_{24}^2}-q\frac{f_4^4f_6f_{24}^2}{f_2f_8^2f_{12}^2}. \end{aligned}$$
(3.7)

Extracting the terms involving \(q^{2n}\) from both sides of (3.6), replacing \(q^2\) by q and using (3.7), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(8n+1)q^n&\equiv \frac{f_1f_2f_3^9}{f_4f_{6}f_{12}}\equiv \frac{f_2f_6^3}{f_4f_{12}}\cdot f_1f_3\nonumber \\&\equiv \frac{f_2f_6^3}{f_4f_{12}}\cdot \left( \frac{f_2f_8^2f_{12}^4}{f_4^2f_6f_{24}^2}-q\frac{f_4^4f_6f_{24}^2}{f_2f_8^2f_{12}^2}\right) \nonumber \\&\equiv \frac{f_2^2f_6^2f_8^2f_{12}^3}{f_4^3f_{24}^2}-q\frac{f_4^3f_6^4f_{24}^2}{f_8^2f_{12}^3}\pmod {4}. \end{aligned}$$
(3.8)

Extracting the terms involving \(q^{2n+1}\) from both sides of (3.8), dividing by q and replacing \(q^2\) by q, then using (2.3), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(16n+9)q^n\equiv - \frac{f_2^3f_3^4f_{12}^2}{f_4^2f_{6}^3} \equiv -\frac{f_6^3}{f_2}\pmod {4}, \end{aligned}$$
(3.9)

which is the case \(\alpha =2\) of (3.1).

Suppose that (3.1) is true for \(\alpha \ge 2\). According to (3.1) and employing (2.2), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +1}n+\frac{2^{2\alpha +1}-5}{3}\right) q^n \equiv -\frac{f_3^3}{f_1}\equiv -\frac{f_4^3f_6^2}{f_2^2f_{12}}-q\frac{f_{12}^3}{f_4}\pmod {4}. \end{aligned}$$
(3.10)

Extracting the terms involving \(q^{2n+1}\) from both sides of (3.10), dividing by q and replacing \(q^2\) by q, we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 2^{2\alpha +2}n+\frac{2^{2\alpha +3}-5}{3}\right) q^n \equiv -\frac{f_6^3}{f_2}\pmod {4}, \end{aligned}$$
(3.11)

which is the case \(\alpha +1\) of (3.1). The proof of Lemma 3.1 is completed by induction on \(\alpha \).

Now, we turn to prove Theorem 1.4. According to (3.1), we have

$$\begin{aligned} A_{3,5}\left( 2^{2\alpha }(2n+1)+\frac{2^{2\alpha +1}-5}{3}\right) \equiv 0 \pmod {4}, \end{aligned}$$
(3.12)

which is (1.17). This completes the proof of Theorem 1.4. \(\square \)

Proof of Theorem 1.5 and Corollary 1.2

Hirschhorn, Garvan and Borwein [9] proved that

$$\begin{aligned} \frac{f_3}{f_1^3}=\frac{f_4^6f_6^3}{f_2^9f_{12}^2}+3q\frac{f_4^2f_6f_{12}^2}{f_2^7}. \end{aligned}$$
(3.13)

Extracting the terms involving \(q^{2n+1}\) from both sides of (3.5), dividing by q and replacing \(q^2\) by q, then using (3.13), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(4n+3)q^n&\equiv 2\frac{f_2^7f_6^5}{f_4f_{12}}\frac{f_3^3}{f_1^9} \equiv 2\frac{f_2^7f_6^5}{f_4f_{12}}\left( \frac{f_4^6f_6^3}{f_2^9f_{12}^2}+3q\frac{f_4^2f_6f_{12}^2}{f_2^7}\right) ^3 \nonumber \\&\equiv 2\frac{f_4^{17}f_6^{14}}{f_2^{20}f_{12}^7}+2q\frac{f_4^{13}f_6^{12}}{f_2^{18}f_{12}^3} +2q^2\frac{f_4^9f_6^{10}f_{12}}{f_2^{16}}+2q^3\frac{f_4^5f_6^8f_{12}^5}{f_2^{14}} \pmod {4}. \end{aligned}$$
(3.14)

Extracting the terms involving \(q^{2n}\) from both sides of (3.14), replacing \(q^2\) by q, we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(8n+3)q^n\equiv 2\frac{f_2^{17}f_3^{14}}{f_1^{20}f_{6}^7}+2q\frac{f_2^9f_3^{10}f_6}{f_1^{16}}\pmod {4}. \end{aligned}$$
(3.15)

By (2.3), we see that

$$\begin{aligned} \frac{f_2^{17}f_3^{14}}{f_1^{20}f_{6}^7}\equiv f_2f_4^3 \pmod {2}, \end{aligned}$$
(3.16)
$$\begin{aligned} \frac{f_2^9f_3^{10}f_6}{f_1^{16}}\equiv f_2f_{12}^3 \pmod {2}. \end{aligned}$$
(3.17)

Combining (3.15), (3.16) and (3.17), we deduce that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(8n+3)q^n\equiv 2f_2f_4^3+2qf_2f_{12}^3 \pmod {4}, \end{aligned}$$
(3.18)

which yields

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(16n+3)q^n&\equiv 2f_1f_2^3\pmod {4}, \end{aligned}$$
(3.19)
$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(16n+11)q^n&\equiv 2 f_1f_{6}^3\pmod {4}. \end{aligned}$$
(3.20)

Using (2.3) in (2.15), we obtain

$$\begin{aligned} \psi (q)\equiv f_1^3 \pmod {4}. \end{aligned}$$
(3.21)

Furthermore, in view of (2.16) and (3.21), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(16n+3)q^n&\equiv 2 f(-q)\psi (q^2)\pmod {4}, \end{aligned}$$
(3.22)
$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(16n+11)q^n&\equiv 2f(-q)\psi (q^6)\pmod {4}. \end{aligned}$$
(3.23)

By (3.22), we find (1.18) holds for \(\alpha =0\). Assume that (1.18) holds for \(\alpha \ge 0\). Employing Lemmas 2.2 and 2.3 in (1.18), we consider the congruence

$$\begin{aligned} \frac{3k^2+k}{2}+2\cdot \frac{m^2+m}{2}\equiv \frac{7p^2-7}{24} \pmod {p}, \end{aligned}$$
(3.24)

where \(-\frac{p-1}{2}\le k\le \frac{p-1}{2}\) and \(0\le m\le \frac{p-3}{2}\).

The congruence (3.24) is equivalent to

$$\begin{aligned} (6k+1)^2+6(2m+1)^2\equiv 0 \pmod {p}. \end{aligned}$$
(3.25)

For \(\left( \frac{-6}{p}\right) =-1\), the congruence (3.25) holds if and only if \(k=\frac{\pm p-1}{6}\) and \(m=\frac{p-1}{2}.\)

Extracting the terms containing \(q^{pn+\frac{7p^2-7}{24}}\) from both sides of (1.18), dividing by \(q^{\frac{7p^2-7}{24}}\) and replacing \(q^p\) by q, we arrive at

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha +1}n+\frac{14\cdot p^{2\alpha +2}-5}{3}\right) q^n\equiv 2f(-q^p)\psi (q^{2p})\pmod {4}. \end{aligned}$$
(3.26)

This implies that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha +2}n+\frac{14\cdot p^{2\alpha +2}-5}{3}\right) q^n\equiv 2f(-q)\psi (q^{2})\pmod {4}, \end{aligned}$$
(3.27)

which is the case \(\alpha +1\) of (1.18). The proof of Theorem 1.5 is completed by induction on \(\alpha \).

Combining (1.18), (2.15) and (2.16), we obtain

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{14\cdot p^{2\alpha }-5}{3}\right) q^n \equiv 2\sum _{n=-\infty }^\infty \sum _{k=0}^\infty q^{\frac{n(3n+1)}{2}+k(k+1)}\pmod {4}, \end{aligned}$$
(3.28)

which implies (1.19).

Applying (3.26) yields that for \(j=1, 2, \ldots ,p-1\),

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha +1}(pn+j)+\frac{14\cdot p^{2\alpha +2}-5}{3}\right) q^n\equiv 0\pmod {4}, \end{aligned}$$
(3.29)

which implies (1.20). This completes the proof of Theorem 1.5 and Corollary 1.2. \(\square \)

Proof of Theorem 1.6 and Corollary 1.3

It follows from (3.23) that (1.21) holds for \(\alpha =0\). Suppose that (1.21) holds for \(\alpha \ge 0\). Employing Lemma 2.2 and Lemma 2.3 in (1.21), we consider the congruence

$$\begin{aligned} \frac{3k^2+k}{2}+6\cdot \frac{m^2+m}{2}\equiv \frac{19p^2-19}{24} \pmod {p}, \end{aligned}$$
(3.30)

where \(-\frac{p-1}{2}\le k\le \frac{p-1}{2}\) and \(0\le m\le \frac{p-3}{2}\).

The congruence (3.30) is equivalent to

$$\begin{aligned} (6k+1)^2+18(2m+1)^2\equiv 0 \pmod {p}. \end{aligned}$$
(3.31)

For \(\left( \frac{-18}{p}\right) =-1\), the congruence (3.31) holds if and only if \(k=\frac{\pm p-1}{6}\) and \(m=\frac{p-1}{2}.\)

Extracting the terms containing \(q^{pn+\frac{19p^2-19}{24}}\) from both sides of (1.21), dividing by \(q^{\frac{19p^2-19}{24}}\) and replacing \(q^p\) by q, we arrive at

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha +1}n+\frac{38\cdot p^{2\alpha +2}-5}{3}\right) q^n\equiv 2f(-q^p)\psi (q^{6p})\pmod {4}. \end{aligned}$$
(3.32)

This implies that

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha +2}n+\frac{38\cdot p^{2\alpha +2}-5}{3}\right) q^n\equiv 2f(-q)\psi (q^{6})\pmod {4}, \end{aligned}$$
(3.33)

which is the case \(\alpha +1\) of (1.21). The proof of Theorem 1.6 is completed by induction on \(\alpha \).

Combining (1.21), (2.15) and (2.16), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( 16\cdot p^{2\alpha }n+\frac{38\cdot p^{2\alpha }-5}{3}\right) q^n \equiv 2\sum _{n=-\infty }^\infty \sum _{k=0}^\infty q^{\frac{n(3n+1)}{2}+3k(k+1)}\pmod {4}, \end{aligned}$$
(3.34)

which implies (1.22).

Applying (3.32) yields that for \(1\le j\le p-1\),

$$\begin{aligned} A_{3,5}\left( 16\cdot p^{2\alpha +1}(pn+j)+\frac{38\cdot p^{2\alpha +2}-5}{3}\right) \equiv 0\pmod {4}, \end{aligned}$$
(3.35)

which implies (1.23). This completes the proof of Theorem 1.6 and Corollary 1.3. \(\square \)

Proof of Theorem 1.7

Xia and Yao [19] proved the following 2-dissection formulae.

$$\begin{aligned} \frac{f_1^3}{f_3}&=\frac{f_4^3}{f_{12}}-3q\frac{f_2^2f_{12}^3}{f_4f_6^2}, \end{aligned}$$
(3.36)
$$\begin{aligned} \frac{1}{f_1^2f_3^2}&=\frac{f_8^5f_{24}^5}{f_2^5f_6^5f_{16}^2f_{48}^2}+2q\frac{f_4^4f_{12}^4}{f_2^6f_6^6} +4q^4\frac{f_4^2f_{12}^2f_{16}^4f_{48}^2}{f_2^5f_6^5f_8f_{24}}. \end{aligned}$$
(3.37)

In view of (3.36), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(n)q^n&=\frac{f_3^{16}}{f_1^8}\frac{f_1^3}{f_3}\equiv \frac{f_6^8}{f_2^4}\left( \frac{f_4^3}{f_{12}}-3q\frac{f_2^2f_{12}^3}{f_4f_6^2}\right) \nonumber \\&\equiv \frac{f_6^8f_4^3}{f_2^4f_{12}}-3q\frac{f_6^6f_{12}^3}{f_2^2f_4}\pmod {8}. \end{aligned}$$
(3.38)

      Utilizing (2.3), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(2n)q^n\equiv \frac{f_3^8f_2^3}{f_1^4f_{6}} \equiv f_2f_6^3\pmod {4}, \end{aligned}$$
(3.39)

which implies (1.24).

Extracting the terms involving \(q^{2n+1}\) from both sides of (3.38), dividing by q and replacing \(q^2\) by q, then using (3.37), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(2n+1)q^n&\equiv -3\frac{f_3^6f_{6}^3}{f_1^2f_2}\equiv 5\frac{f_3^6f_{6}^3}{f_1^2f_2}\equiv 5\frac{f_3^8f_{6}^3}{f_2f_1^2f_3^2} \equiv 5\frac{f_6^7}{f_2}\frac{1}{f_1^2f_3^2}\nonumber \\&\equiv 5\frac{f_6^7}{f_2}\left( \frac{f_8^5f_{24}^5}{f_2^5f_6^5f_{16}^2f_{48}^2}+2q\frac{f_4^4f_{12}^4}{f_2^6f_6^6} +4q^4\frac{f_4^2f_{12}^2f_{16}^4f_{48}^2}{f_2^5f_6^5f_8f_{24}}\right) \nonumber \\&\equiv 5\frac{f_6^2f_8^5f_{24}^5}{f_2^6f_{16}^2f_{48}^2}+2q\frac{f_4^4f_6f_{12}^4}{f_2^7} +4q^4\frac{f_4^2f_6^2f_{12}^2f_{16}^4f_{48}^2}{f_2^6f_8f_{24}}\pmod {8}. \end{aligned}$$
(3.40)

In view of (2.3), we obtain

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(4n+1)q^n&\equiv 5\frac{f_3^2f_4^5f_{12}^5}{f_1^6f_{8}^2f_{24}^2}+4q^2\frac{f_2^2f_3^2f_{6}^2f_{8}^4f_{24}^2}{f_1^6f_4f_{12}} \nonumber \\&\equiv 5\frac{f_3^2f_4^5f_{12}^5}{f_1^6f_{8}^2f_{24}^2}+4q^2\frac{f_6f_{8}^4f_{24}^2}{f_2f_4}\pmod {8}. \end{aligned}$$
(3.41)

Using (3.13), we see that

$$\begin{aligned} 5\frac{f_3^2f_4^5f_{12}^5}{f_1^6f_{8}^2f_{24}^2}&\equiv 5\frac{f_4^5f_{12}^5}{f_8^2f_{24}^2}\left( \frac{f_3}{f_1^3}\right) ^2 \equiv 5\frac{f_4^5f_{12}^5}{f_8^2f_{24}^2}\left( \frac{f_4^6f_6^3}{f_2^9f_{12}^2}+3q\frac{f_4^2f_6f_{12}^2}{f_2^7}\right) ^2\nonumber \\&\equiv 5\frac{f_4^5f_{12}^5}{f_8^2f_{24}^2} \left( \frac{f_4^{12}f_6^6}{f_2^{18}f_{12}^4}+6q\frac{f_4^8f_6^4}{f_2^{16}}+9q^2\frac{f_4^4f_6^2f_{12}^4}{f_2^{14}}\right) \nonumber \\&\equiv 5\frac{f_4^{17}f_6^6f_{12}}{f_2^{18}f_8^2f_{24}^2} +6q\frac{f_4^{13}f_6^4f_{12}^5}{f_2^{16}f_8^2f_{24}^2}+5q^2\frac{f_4^9f_6^2f_{12}^9}{f_2^{14}f_8^2f_{24}^2}\pmod {8}. \end{aligned}$$
(3.42)

Combining (3.41) and (3.42), extracting the terms involving \(q^{2n+1}\) from both sides of (3.41), dividing by q and replacing \(q^2\) by q, we find

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(8n+5)q^n&\equiv 6\frac{f_2^{13}f_3^4f_{6}^5}{f_1^{16}f_4^2f_{12}^2}\equiv 6f_2f_6^3\pmod {8}, \end{aligned}$$
(3.43)

which implies (1.25). This completes the proof of Theorem 1.7. \(\square \)

Proof of Theorem 1.8 and Corollary 1.4

By (3.39), using (2.16) and (3.21), we get

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(2n)q^n\equiv f(-q^2)\psi (q^6) \pmod {4}, \end{aligned}$$
(3.44)

which is the case \(\alpha =0\) of (1.26).

Suppose that (1.26) holds for \(\alpha \ge 0\). Employing Lemma 2.2 and Lemma 2.3 in (1.26), we consider the congruence

$$\begin{aligned} 2\cdot \frac{3k^2+k}{2}+6\cdot \frac{m^2+m}{2}\equiv \frac{20p^2-20}{24} \pmod {p}, \end{aligned}$$
(3.45)

where \(-\frac{p-1}{2}\le k\le \frac{p-1}{2}\) and \(0\le m\le \frac{p-3}{2}\).

The congruence (3.45) is equivalent to

$$\begin{aligned} (6k+1)^2+9(2m+1)^2\equiv 0 \pmod {p}. \end{aligned}$$
(3.46)

For \(\left( \frac{-9}{p}\right) =-1\), the congruence (3.46) holds if and only if \(k=\frac{\pm p-1}{6}\) and \(m=\frac{p-1}{2}.\)

Extracting the terms containing \(q^{pn+\frac{5p^2-5}{6}}\) from both sides of (1.26), dividing by \(q^{\frac{5p^2-5}{6}}\) and replacing \(q^p\) by q, we arrive at

$$\begin{aligned}&\sum _{n=0}^\infty A_{3,5}\left( 2\cdot p^{2\alpha +1}n+\frac{5\cdot p^{2\alpha +2}-5}{3}\right) q^n\nonumber \\&\quad \equiv (-1)^{(\alpha +1)(\frac{\pm p-1}{6})}f(-q^{2p})\psi (q^{6p})\pmod {4}. \end{aligned}$$
(3.47)

This implies that

$$\begin{aligned}&\sum _{n=0}^\infty A_{3,5}\left( 2\cdot p^{2\alpha +2}n+\frac{5\cdot p^{2\alpha +2}-5}{3}\right) q^n\nonumber \\&\quad \equiv (-1)^{(\alpha +1)(\frac{\pm p-1}{6})}f(-q^2)\psi (q^6)\pmod {4}, \end{aligned}$$
(3.48)

which is the case \(\alpha +1\) of (1.26). The proof of Theorem 1.8 is completed by induction on \(\alpha \).

Combining (1.26), (2.15) and (2.16), we obtain

$$\begin{aligned}&\sum _{n=0}^\infty A_{3,5}\left( 4\cdot p^{2\alpha }n+\frac{5\cdot p^{2\alpha }-5}{3}\right) q^n\equiv \sum _{n=-\infty }^\infty \sum _{k=0}^\infty (-1)^{\alpha (\frac{\pm p-1}{6})+n}q^{\frac{n(3n+1)}{2}+3\frac{k(k+1)}{2}}\pmod {4}, \end{aligned}$$

which implies (1.27).

Applying (3.47) yields that for \(j=1,2, \cdots , p-1\),

$$\begin{aligned} A_{3,5}\left( 2\cdot p^{2\alpha +1}(pn+j)+\frac{5\cdot p^{2\alpha +2}-5}{3}\right) \equiv 0\pmod {4}. \end{aligned}$$

which implies (1.28). This finishes the proof of Theorem 1.8 and Corollary 1.4. \(\square \)

Proof of Theorem 1.9 and Corollary 1.5

By (3.43), according to (2.16) and (3.21), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}(8n+5)q^n\equiv 6f(-q^2)\psi (q^6)\pmod {8}, \end{aligned}$$
(3.49)

which is the case \(\alpha =0\) of (1.29).

Suppose that (1.29) holds for \(\alpha \ge 0\). Employing Lemma 2.2 and Lemma 2.3 in (1.29), we consider the congruence

$$\begin{aligned} 2\cdot \frac{3k^2+k}{2}+6\cdot \frac{m^2+m}{2}\equiv \frac{20p^2-20}{24} \pmod {p}, \end{aligned}$$
(3.50)

where \(-\frac{p-1}{2}\le k\le \frac{p-1}{2}\) and \(0\le m\le \frac{p-3}{2}\).

The congruence (3.50) is equivalent to

$$\begin{aligned} (6k+1)^2+9(2m+1)^2\equiv 0 \pmod {p}. \end{aligned}$$
(3.51)

For \(\left( \frac{-9}{p}\right) =-1\), the congruence (3.51) holds if and only if \(k=\frac{\pm p-1}{6}\) and \(m=\frac{p-1}{2}.\)

Extracting the terms containing \(q^{pn+\frac{5p^2-5}{6}}\) from both sides of (1.29), dividing by \(q^{\frac{5p^2-5}{6}}\) and replacing \(q^p\) by q, we arrive at

$$\begin{aligned}&\sum _{n=0}^\infty A_{3,5}\left( 8\cdot p^{2\alpha +1}n+\frac{20\cdot p^{2\alpha +2}-5}{3}\right) q^n\nonumber \\&\quad \equiv 6(-1)^{(\alpha +1) (\frac{\pm p-1}{6})}f(-q^{2p})\psi (q^{6p})\pmod {8}. \end{aligned}$$
(3.52)

This implies that

$$\begin{aligned}&\sum _{n=0}^\infty A_{3,5}\left( 8\cdot p^{2\alpha +2}n+\frac{20\cdot p^{2\alpha +2}-5}{3}\right) q^n\nonumber \\&\quad \equiv 6(-1)^{(\alpha +1) (\frac{\pm p-1}{6})}f(-q^{2})\psi (q^{6})\pmod {8}, \end{aligned}$$
(3.53)

which is the case \(\alpha +1\) of (1.29). The proof of Theorem 1.9 is completed by induction on \(\alpha \).

Combining (1.29), (2.16) and (3.21), we have

$$\begin{aligned} \sum _{n=0}^\infty A_{3,5}\left( \!16\cdot p^{2\alpha }n+\frac{20\cdot p^{2\alpha }-5}{3}\!\right) q^n \equiv 6 \sum _{n=-\infty }^\infty \sum _{k=0}^\infty (-1)^{\alpha (\frac{\pm p-1}{6})+n}q^{\frac{n(3n+1)}{2}+3\frac{k(k+1)}{2}}\pmod {8}, \end{aligned}$$

which implies (1.30).

Applying (3.52) yields that for \(1\le j\le p-1\),

$$\begin{aligned} A_{3,5}\left( 8\cdot p^{2\alpha +1}(pn+j)+\frac{20\cdot p^{2\alpha +2}-5}{3}\right) \equiv 0\pmod {8}, \end{aligned}$$
(3.54)

which implies (1.31). This completes the proof of Theorem 1.9 and Corollary 1.5. \(\square \)