Hostname: page-component-7bb8b95d7b-495rp Total loading time: 0 Render date: 2024-09-27T10:46:37.812Z Has data issue: false hasContentIssue false

A CONJECTURE OF MERCA ON CONGRUENCES MODULO POWERS OF 2 FOR PARTITIONS INTO DISTINCT PARTS

Published online by Cambridge University Press:  27 March 2023

JULIA Q. D. DU
Affiliation:
School of Mathematical Sciences, Hebei International Joint Research Center for Mathematics and Interdisciplinary Science, Hebei Normal University, Shijiazhuang 050024, PR China e-mail: [email protected]
DAZHAO TANG*
Affiliation:
School of Mathematical Sciences, Chongqing Normal University, Chongqing 401331, PR China
Rights & Permissions [Opens in a new window]

Abstract

Let $Q(n)$ denote the number of partitions of n into distinct parts. Merca [‘Ramanujan-type congruences modulo 4 for partitions into distinct parts’, An. Şt. Univ. Ovidius Constanţa 30(3) (2022), 185–199] derived some congruences modulo $4$ and $8$ for $Q(n)$ and posed a conjecture on congruences modulo powers of $2$ enjoyed by $Q(n)$. We present an approach which can be used to prove a family of internal congruence relations modulo powers of $2$ concerning $Q(n)$. As an immediate consequence, we not only prove Merca’s conjecture, but also derive many internal congruences modulo powers of $2$ satisfied by $Q(n)$. Moreover, we establish an infinite family of congruence relations modulo $4$ for $Q(n)$.

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

A partition $\pi $ of a positive integer n is a finite weakly decreasing sequence of positive integers $\pi _1\geq \pi _2\geq \cdots \geq \pi _r$ such that $\sum _{i=1}^r\pi _i=n$ . The $\pi _i$ are called the parts of the partition $\pi $ . Let $p(n)$ denote the number of partitions of n with the convention that $p(0)=1$ . The generating function of $p(n)$ , derived by Euler, is given by

$$ \begin{align*} \sum_{n=0}^{\infty} p(n)q^n=\dfrac{1}{(q;q)_{\infty}}, \end{align*} $$

where, here and throughout this paper, we always assume that q is a complex number such that $|q|<1$ and adopt the customary notation:

$$ \begin{align*} (a;q)_{\infty}=\prod_{j=0}^{\infty}(1-aq^j). \end{align*} $$

In 1919, Ramanujan discovered three celebrated congruences for the partition function $p(n)$ (see [Reference Berndt and Ono4]), which were later confirmed by Atkin [Reference Atkin2] and Watson [Reference Watson15]: for any $n\geq 0$ and $\alpha \geq 1$ ,

(1.1) $$ \begin{align} \kern3pt p(5^{\alpha} n+\delta_{5,\alpha}) &\equiv0\pmod{5^{\alpha}}, \end{align} $$
(1.2) $$ \begin{align} \kern24pt p(7^{\alpha} n+\delta_{7,\alpha}) &\equiv0\pmod{7^{\lfloor\alpha/2\rfloor+1}}, \end{align} $$
(1.3) $$ \begin{align} p(11^{\alpha} n+\delta_{11,\alpha}) &\equiv0\pmod{11^{\alpha}}, \end{align} $$

where $\delta _{p,\alpha }$ is the least positive integer satisfying $24\delta _{p,\alpha }\equiv 1\pmod {p^{\alpha }}$ with ${p\in \{5,7,11\}}$ . Since then, congruence properties for various partition functions have been a hot topic in the theory of partitions and have motivated a large amount of research.

Another ingredient of the theory of partitions is the study of partition identities. In 1748, Euler [Reference Euler7] proved the most well-known partition theorem which states that there are as many partitions of n into distinct parts as into odd parts. In terms of the generating function,

(1.4) $$ \begin{align} \sum_{n=0}^{\infty} Q(n)q^n=(-q;q)_{\infty}=\dfrac{1}{(q;q^2)_{\infty}}=\dfrac{(q^2;q^2)_{\infty}}{(q;q)_{\infty}}, \end{align} $$

where $Q(n)$ denotes the number of partitions of n into distinct parts. According to Euler’s pentagonal number theorem [Reference Andrews and Berndt1, page 17, (1.4.11)],

$$ \begin{align*} (q;q)_{\infty}=\sum_{n=-\infty}^{\infty}(-1)^nq^{n(3n-1)/2}, \end{align*} $$

we find that almost all values of $Q(n)$ are even, that is,

(1.5) $$ \begin{align} \lim_{X\rightarrow\infty}\dfrac{\#\{0\leq n\leq X\colon Q(n)\equiv0\pmod{2}\}}{X}=1. \end{align} $$

Indeed, $Q(n)$ is odd if and only if n is a generalised pentagonal number. Motivated by (1.1)–(1.5), many scholars subsequently investigated congruence properties and arithmetic density properties of $Q(n)$ . For instance, in 1997, Gordon and Ono [Reference Gordon and Ono8] proved the striking result that for any positive integer m, $Q(n)$ is divisible by $2^m$ for almost all nonnegative integers n, that is,

(1.6) $$ \begin{align} \lim_{X\rightarrow\infty}\dfrac{\#\{0\leq n\leq X\colon Q(n)\equiv0\pmod{2^m}\}}{X}=1. \end{align} $$

The identity (1.6) is a powerful result on the arithmetic properties of $Q(n)$ . However, it is not a constructive result and the theory of modular forms used in the proof of (1.6) cannot be applied to derive the explicit congruences enjoyed by $Q(n)$ . Therefore, it is still of interest to find explicit congruences for $Q(n)$ .

In a recent paper, Merca [Reference Merca9] derived some congruences modulo $4$ and $8$ for $Q(n)$ by using Smoot’s Mathematica implementation [Reference Smoot13] of Radu’s algorithm [Reference Radu12] on Ramanujan–Kolberg identities for partition functions. At the end of his paper, Merca posed the following conjecture on congruences modulo powers of 2 for $Q(n)$ .

Conjecture 1.1 (Merca [Reference Merca9], Conjecture).

Let $(p,k)\in S$ . For any $n\not \equiv 0\pmod {p}$ ,

$$ \begin{align*} Q{\bigg(pn+\dfrac{p^2-1}{24}\bigg)}\equiv0\pmod{2^k}, \end{align*} $$

where

(1.7) $$ \begin{align} S\in\{ &(11,5),(13,6),(17,8),(19,9),(23,11),(31,3),(37,6),\nonumber\\ &(41,8),(43,9),(47,11),(59,6),(61,6),(67,10),(71,13),\nonumber\\ &(79,3),(83,5),(89,9),(103,3),(107,6),(109,6),(113,9)\}. \end{align} $$

In this paper, we prove the following result.

Theorem 1.2. Let S be defined as in (1.7). Then for any $(p,k)\in S$ ,

(1.8) $$ \begin{align} \sum_{n=0}^{\infty} Q{\bigg(pn+\dfrac{p^2-1}{24}\bigg)}q^n\equiv c_p\sum_{n=0}^{\infty} Q(n)q^{pn}\pmod{2^k}, \end{align} $$

where $c_p$ is given in Table 1.

Table 1 A table of values of $c_p$ .

As an immediate consequence of (1.8), we obtain the following congruences and internal congruences enjoyed by $Q(n)$ , which confirms Conjecture 1.1.

Corollary 1.3. Let S be defined as in (1.7). Then for any $(p,k)\in S$ and $1\leq i\leq p-1$ ,

$$ \begin{align*} Q\kern1pt{\bigg(p^2n+\dfrac{(24i+p)p-1}{24}\bigg)}\equiv0\pmod{2^k}. \end{align*} $$

Moreover, for any $n\geq 0$ ,

$$ \begin{align*} Q{\bigg(p^2n+\dfrac{p^2-1}{24}\bigg)}\equiv c_pQ(n)\pmod{2^k}, \end{align*} $$

where $c_p$ is given in Table 1.

The following theorem shows that there are an infinite family of congruence relations of the form (1.8) satisfied by $Q(n)$ .

Theorem 1.4. Let $p\geq 5$ be a prime number. If $\big(\frac{-24}{p}\big)=-1$ , then

(1.9) $$ \begin{align} \sum_{n=0}^{\infty} Q{\bigg(pn+\dfrac{p^2-1}{24}\bigg)}q^n\equiv(-1)^{(\pm p-1)/6}\sum_{n=0}^{\infty} Q(n)q^{pn}\pmod{4}, \end{align} $$

where $\big (\frac {\cdot }{p}\big )$ is the Legendre symbol and

(1.10) $$ \begin{align} (\pm p-1)/6= \begin{cases} (p-1)/6 &\textrm{if } p\equiv1\pmod{6},\\ (-p-1)/6 &\textrm{if } p\equiv5\pmod{6}. \end{cases} \end{align} $$

The rest of this paper is organised as follows. In Section 2, we collect some notation and terminology on modular forms. The proof of Theorem 1.2 is presented in Section 3 and that of Theorem 1.4 in Section 4. Finally, we pose a conjecture on congruence relations for $Q(n)$ modulo $4$ which strengthens both (1.9) and a result of Merca.

2 Preliminaries

We first recall some terminology from the theory of modular forms. The full modular group is given by

$$ \begin{align*} \Gamma=\bigg\{\begin{pmatrix}a &b\\ c &d\end{pmatrix}\colon a, b, c, d\in\mathbb{Z}, \text{ and } ad-bc=1\bigg\}, \end{align*} $$

and for a positive integer N, the congruence subgroup $\Gamma _0(N)$ is defined by

$$ \begin{align*} \Gamma_0(N)=\bigg\{\begin{pmatrix}a &b\\ c &d\end{pmatrix}\in\Gamma \colon c\equiv 0\pmod{N}\bigg\}. \end{align*} $$

Let $\gamma $ be the matrix $(\begin {smallmatrix}a &b\\ c &d\end {smallmatrix})$ from now on. Then $\gamma $ acts on $\tau \in \mathbb {C}$ by the linear fractional transformation

$$ \begin{align*} \gamma\tau = \frac{a\tau+b}{c\tau+d} \quad \text{and} \quad \gamma\infty =\lim_{\tau\rightarrow \infty} \gamma\tau. \end{align*} $$

Let N, k be positive integers and $\mathbb {H}=\{\tau \in \mathbb {C}\colon \mathrm {Im}(\tau )>0\}$ . A holomorphic function $f\colon \mathbb {H}\rightarrow \mathbb {C}$ is called a modular function of weight k for $\Gamma _0(N)$ if it satisfies the following two conditions:

  1. (1) for all $\gamma \in \Gamma _0(N)$ , $f(\gamma \tau )=(c\tau +d)^kf(\tau )$ ;

  2. (2) for any $\gamma \in \Gamma $ , $(c\tau +d)^{-k}f(\gamma \tau )$ has a Fourier expansion of the form

    $$ \begin{align*} (c\tau+d)^{-k}f(\gamma\tau)=\sum_{n=n_{\gamma}}^{\infty} a(n)q_{w_{\gamma}}^n, \end{align*} $$

    where $a(n_{\gamma })\neq 0$ , $q_{w_{\gamma }}=e^{2\pi i\tau /w_{\gamma }}$ and $w_{\gamma }={N}/{\gcd (c^2, N)}$ .

In particular, if $n_{\gamma } \geq 0$ for all $\gamma \in \Gamma $ , then we call f a modular form of weight k for $\Gamma _0(N)$ . A modular function with weight 0 for $\Gamma _0(N)$ is referred to as a modular function for $\Gamma _0(N)$ . For a modular function $f(\tau )$ of weight k with respect to $\Gamma _0(N)$ , the order of $f(\tau )$ at the cusp $a/c\in \mathbb {Q}\cup \{\infty \}$ is defined by

$$ \begin{align*} \mathrm{ord}_{a/c}(f)=n_{\gamma} \end{align*} $$

for some $\gamma \in \Gamma $ such that $\gamma \infty =a/c$ ; $\mathrm {ord}_{a/c}(f)$ is well defined (see [Reference Diamond and Shurman6, page 72]).

Radu [Reference Radu12] developed the Ramanujan–Kolberg algorithm to derive the Ramanujan– Kolberg identities on a class of partition functions defined in terms of eta-quotients using modular functions for $\Gamma _0(N)$ (see [Reference Paule, Radu, Beveridge, Griggs, Hogben, Musiker and Tetali11]). Smoot [Reference Smoot13] developed a Mathematica package $\mathtt {RaduRK}$ to implement Radu’s algorithm.

Let the partition function $a(n)$ be defined by

(2.1) $$ \begin{align} \sum_{n=0}^{\infty} a(n)q^n=\prod_{\delta|M} (q^{\delta}, q^{\delta})_{\infty}^{r_{\delta}}, \end{align} $$

where M, $\delta $ are positive integers and $r_{\delta }$ are integers. For any $m\geq 1$ and $0\leq t\leq m-1$ , Radu [Reference Radu12] defined

$$ \begin{align*} g_{m,t}(\tau) = q^{{(t+\ell)}/{m}}\sum\limits_{n=0}^{\infty} a(mn+t)q^n, \end{align*} $$

where

$$ \begin{align*} \ell=\frac{1}{24}\sum\limits_{\delta|M}\delta r_{\delta}, \end{align*} $$

and gave a criterion for a function involving $g_{m,t}(\tau )$ to be a modular function with respect to $\Gamma _0(N)$ , where N satisfies the following conditions, with $\kappa =\gcd (1-m^2, 24)$ :

  1. (1) for every prime p, $p \mid m$ implies $p \mid N$ ;

  2. (2) for every $\delta $ dividing M with $r_{\delta }\neq 0$ , $\delta \mid M$ implies $\delta \mid mN$ ;

  3. (3) $\kappa mN^2\sum _{\delta |M}{r_{\delta }}/{\delta }\equiv 0\pmod {24}$ ;

  4. (4) $\kappa N \sum _{\delta |M} r_{\delta } \equiv 0\pmod {8}$ ;

  5. (5) ${24 m}/{\gcd (\kappa (-24 t-\sum _{\delta |M}\delta r_{\delta }), 24 m)} \mid N ;$

  6. (6) if $2 \mid m$ , then $\kappa N\equiv 0\pmod {4}$ and $8 \mid Ns$ , or $2 \mid s$ and $8 \mid N(1-j)$ , where $\prod _{\delta |M} \delta ^{|r_{\delta }|}=2^sj$ and $j, s\in \mathbb {Z}$ with j odd.

Given a positive integer n and an integer x, we denote by $[x]_n$ the residue class of x modulo n. Let

$$ \begin{align*} \mathbb{Z}_n^*=\{[x]_n\in\mathbb{Z}_n : \gcd(x,n)=1\} \quad \text{and} \quad \mathbb{S}_n=\{y^2: y\in\mathbb{Z}_n^*\}. \end{align*} $$

Define the set

$$ \begin{align*} P_m(t)=\bigg\{\bigg[ts+\frac{s-1}{24}\sum_{\delta|M}\delta r_{\delta}\bigg]_m : s\in\mathbb{S}_{24m}\bigg\}. \end{align*} $$

Recall that the Dedekind eta-function $\eta (\tau )$ is defined by

$$ \begin{align*} \eta(\tau)=q^{{1}/{24}}\prod_{n=1}^{\infty}(1-q^n), \end{align*} $$

where $q=e^{2\pi i \tau }$ and $\tau \in \mathbb {H}$ .

Theorem 2.1 [Reference Radu12, Theorem 45].

For a partition function $a(n)$ defined as in (2.1), and integers $m\geq 1, 0\leq t\leq m-1$ , suppose that N is a positive integer satisfying the conditions (1)(6). Let

$$ \begin{align*} F(\tau)=\prod_{\delta|N} \eta^{s_{\delta}}(\delta\tau)\prod_{t'\in P_{m}(t)}g_{m,t'}(\tau), \end{align*} $$

where $s_{\delta }$ are integers. Then $F(\tau )$ is a modular function for $\Gamma _0(N)$ if and only if the $s_{\delta }$ satisfy the following conditions:

  1. (1) $|P_m(t)| \sum _{\delta |M}r_{\delta }+\sum _{\delta |N} s_{\delta } =0;$

  2. (2) $\sum _{t'\in P_m(t)}{(1-m^2)(24t'+\sum _{\delta |M}\delta r_{\delta })}/{m} +|P_m(t)|m\sum _{\delta |M}\delta r_{\delta }+\sum _{\delta |N}\delta s_{\delta }\equiv 0\pmod {24};$

  3. (3) $|P_m(t)|mN\sum _{\delta |M}{r_{\delta }}/{\delta }+\sum _{\delta |N} ({N}/{\delta })s_{\delta } \equiv 0\pmod {24};$

  4. (4) ${(\prod _{\delta |M}(m\delta )^{|r_{\delta }|})}^{|P_{m}(t)|}\prod _{\delta |N}\delta ^{|s_{\delta }|}$ is a square.

Radu [Reference Radu12, Theorem 47] also gave lower bounds for the orders of $F(\tau )$ at cusps of $\Gamma _0(N)$ .

Theorem 2.2. For a partition function $a(n)$ defined as in (2.1) and integers $m\geq 1$ , $0\leq t\leq m-1$ , let

$$ \begin{align*} F(\tau)=\prod_{\delta|N} \eta^{s_{\delta}}(\delta\tau)\prod_{t'\in P_{m}(t)}g_{m,t'}(\tau) \end{align*} $$

be a modular function for $\Gamma _0(N)$ , where $s_{\delta }$ are integers and N satisfies the conditions (1)(6). Let $\{s_1, s_2,\ldots ,s_{\epsilon }\}$ be a complete set of inequivalent cusps of $\Gamma _0(N)$ and, for $1\leq i \leq \epsilon $ , let $\gamma _i\in \Gamma $ be such that $\gamma _i\infty = s_i$ . Then

$$ \begin{align*} \mathrm{ord}_{s_i}(F(\tau)) \geq \frac{N}{\gcd(c^2, N)}(|P_m(t)|p(\gamma_i)+p^*(\gamma_i)), \end{align*} $$

where

$$ \begin{align*} p(\gamma_i)=\min_{\lambda\in\{0, 1, \ldots, m-1\}}\frac{1}{24}\sum_{\delta|M} r_{\delta}\frac{\gcd^2(\delta(a+\kappa\lambda c), mc)}{\delta m} \end{align*} $$

and

$$ \begin{align*} p^*(\gamma_i)=\frac{1}{24}\sum_{\delta|N}s_{\delta}\frac{\gcd^2(\delta, c)}{\delta}. \end{align*} $$

The following theorem of Sturm [Reference Sturm14, Theorem 1] plays an important role in proving congruences using the theory of modular forms.

Theorem 2.3. Let k be an integer and $g(\tau )=\sum _{n=0}^{\infty } c(n)q^n$ a modular form of weight k for $\Gamma _0(N)$ . For any given positive integer u, if $c(n)\equiv 0\pmod {u}$ holds for all $n\leq ({kN}/{12})\prod _{p|N, \, p\mathrm {~prime}}{(1+{1}/{p})}$ , then $c(n)\equiv 0\pmod {u}$ holds for any $n\geq 0$ .

3 Proof of Theorem 1.2

This section is devoted to the proof of Theorem 1.2. The following lemma plays a vital role in the proof of Theorem 1.2.

Lemma 3.1. Let $p$ be a prime with $p\ge5$ and define $k_1 = \lceil(p^2-1)/48p\rceil$ and ${k_2 = \lceil(p^2-1)/48p^2\rceil}$ . Then for any constant c,

$$ \begin{align*} \frac{\eta^{24k_1}(\tau)\eta^{16k_2}(2p\tau)}{\eta^{8k_2}(p\tau)} {\bigg(q^{{p}/{24}}\dfrac{\eta(p\tau)}{\eta(2p\tau)}\sum_{n=0}^{\infty} Q{\bigg(pn+\frac{p^2-1}{24}\bigg)}q^n-c\bigg)} \end{align*} $$

is a modular form of weight $12k_1+4k_2$ for $\Gamma _0(2p)$ .

Proof. Recall that the generating function of $Q(n)$ is

$$ \begin{align*} \sum_{n=0}^{\infty} Q(n)q^n=\frac{(q^2; q^2)_{\infty}}{(q; q)_{\infty}}. \end{align*} $$

Taking $M=2$ , $(r_1, r_2)=(-1,1)$ , $m=p$ , $t=(p^2-1)/24$ in Theorem 2.1, one can find that $N=2p$ satisfies the conditions (1)(6), and for $(s_1, s_2, s_p, s_{2p})=(0,0,1, -1)$ ,

$$ \begin{align*} F(\tau)=q^{{p}/{24}}\frac{\eta(p\tau)}{\eta(2p\tau)} \sum_{n=0}^{\infty} Q{\bigg(pn+\frac{p^2-1}{24}\bigg)}q^n \end{align*} $$

is a modular function for $\Gamma _0(2p)$ .

By Theorem 2.2, we derive lower bounds for the orders of $F(\tau )$ at the cusps of $\Gamma _0(2p)$ :

$$ \begin{align*} \mathrm{ord}_0(F(\tau)) \geq-\frac{p^2-1}{24}, &\quad \mathrm{ord}_{1/2}(F(\tau))\geq -\frac{1}{24p},\\ \mathrm{ord}_{1/p}(F(\tau)) \geq \frac{2p^2-1}{24p}, &\quad \mathrm{ord}_{\infty}(F(\tau))\geq -\frac{p^2-1}{24p}, \end{align*} $$

which implies that

$$ \begin{align*} \mathrm{ord}_0(F(\tau)-c) \geq -\frac{p^2-1}{24}, &\quad \mathrm{ord}_{1/2}(F(\tau)-c) \geq 0,\\ \mathrm{ord}_{1/p}(F(\tau)-c) \geq 0, &\quad \mathrm{ord}_{\infty}(F(\tau)-c) \geq -\frac{p^2-1}{24p}. \end{align*} $$

By [Reference Ono10, Theorems 1.64 and 1.65], one easily shows

$$ \begin{align*} F_1(\tau)=\eta^{24}(\tau) \quad \text{and} \quad F_2(\tau)=\frac{\eta^{16}(2p\tau)}{\eta^8(p\tau)} \end{align*} $$

are modular forms with weight 12 and 4 for $\Gamma _0(2p)$ , respectively, and the orders at the cusps of $\Gamma _0(2p)$ are

$$ \begin{align*} \mathrm{ord}_0(F_1(\tau)) =2p, \quad \mathrm{ord}_{1/2}(F_1(\tau)) =p, &\quad \mathrm{ord}_{1/p}(F_1(\tau)) =2, \quad \mathrm{ord}_{\infty}(F_1(\tau)) =1,\\ \mathrm{ord}_0(F_2(\tau)) =0, \quad \mathrm{ord}_{1/2}(F_2(\tau)) =1, &\quad \mathrm{ord}_{1/p}(F_2(\tau)) =0, \quad \mathrm{ord}_{\infty}(F_2(\tau)) =p. \end{align*} $$

Therefore, the orders of $F^{k_1}_1(\tau )F^{k_2}_2(\tau )F(\tau )$ at all cusps of $\Gamma _0(2p)$ are nonnegative, and so $F^{k_1}_1(\tau )F^{k_2}_2(\tau )F(\tau )$ is a modular form with weight $12k_2+4k_2$ for $\Gamma _0(2p)$ . This completes the proof.

Proof of Theorem 1.2.

Fix $k\ge 1$ . By Lemma 3.1 and Sturm’s theorem, to prove

$$ \begin{align*} \dfrac{(q^p;q^p)_{\infty}}{(q^{2p};q^{2p})_{\infty}} \sum_{n=0}^{\infty} Q{\bigg(pn+\frac{p^2-1}{24}\bigg)}q^n-c_p\equiv 0\pmod{2^k}, \end{align*} $$

we only need to check that the coefficients of the first $l_p=(p+1)(3k_1+k_2)$ terms of the expansion of

$$ \begin{align*} \frac{\eta^{24k_1}(\tau)\eta^{16k_2}(2p\tau)}{\eta^{8k_2}(p\tau)} {\bigg(q^{{p}/{24}}\dfrac{\eta(p\tau)}{\eta(2p\tau)} \sum_{n=0}^{\infty} Q{\bigg(pn+\frac{p^2-1}{24}\bigg)}q^n-c_p\bigg)} \end{align*} $$

are congruent to 0 modulo $2^k$ . Here, $k_1$ and $k_2$ are defined in Lemma 3.1 and the corresponding $l_p$ are displayed in Table 2. This information allows us to do the computations to complete the proof of Theorem 1.2.

Table 2 A table of values of $l_p$ .

4 Proof of Theorem 1.4

In this section, we give a proof of Theorem 1.4. Before starting the proof, we need to introduce Ramanujan’s theta function, given by

(4.1) $$ \begin{align} f(a,b)=\sum_{n=0}^{\infty} a^{n(n+1)/2}b^{n(n-1)/2}=(-a,-b,ab;ab)_{\infty},\quad|ab|<1, \end{align} $$

where the last identity in (4.1) is the celebrated Jacobi triple product [Reference Andrews and Berndt1, page 17, (1.4.8)]. Two important cases of $f(a,b)$ are

(4.2) $$ \begin{align} \varphi(q) &:=f(q,q)=\sum_{n=0}^{\infty} q^{n^2}=\dfrac{(q^2;q^2)_{\infty}^5}{(q;q)_{\infty}^2(q^4;q^4)_{\infty}^2},\\ f(-q) &:=f(-q,-q^2)=\sum_{n=-\infty}^{\infty} q^{n(3n+1)/2}=(q;q)_{\infty}.\nonumber \end{align} $$

Replacing q by $-q$ in (4.2) yields

$$ \begin{align*} \varphi(-q)=\dfrac{(q;q)_{\infty}^2}{(q^2;q^2)_{\infty}}. \end{align*} $$

The following p-dissections for $\varphi (-q)$ and $f(-q)$ play an important role in the proof of Theorem 1.4.

Lemma 4.1. Let $p\geq 5$ be a prime number. Then

(4.3) $$ \begin{align} \varphi(-q) &=\varphi({-}q^{p^2})+2\sum_{j=1}^{(p-1)/2}q^{{\kern1.2pt}j^2} f({-}q^{p^2+2pj},-q^{p^2-2pj}), \end{align} $$
(4.4) $$ \begin{align} f(-q) &=\sum_{\substack{k=-(p-1)/2\\ k\neq(\pm p-1)/6}}^{(p-1)/2}(-1)^kq^{k(3k+1)/2} f({-}q^{(3p^2+(6k+1)p)/2},-q^{(3p^2-(6k+1)p)/2})\nonumber\\ &\quad+(-1)^{(\pm p-1)/6}q^{(p^2-1)/24}f({-}q^{p^2}), \end{align} $$

where $(\pm p-1)/6$ is defined as in (1.10). Further, for $-(p-1)/2\leq k\leq (p-1)/2$ and $k\neq (\pm p-1)/6$ ,

$$ \begin{align*} \dfrac{3k^2+k}{2}\not\equiv\dfrac{p^2-1}{24}\pmod{p}. \end{align*} $$

Proof. The identity (4.3) follows immediately from [Reference Berndt3, page 49]. The identity (4.4) appears in [Reference Cui and Gu5, Theorem 2.2].

Proof of Theorem 1.4.

From (1.4), we find that

(4.5) $$ \begin{align} \sum_{n=0}^{\infty} Q(n)q^n=\dfrac{(q^2;q^2)_{\infty}}{(q;q)_{\infty}}=\dfrac{(q^2;q^2)_{\infty}^2}{(q;q)_{\infty}^4}\cdot \dfrac{(q;q)_{\infty}^3}{(q^2;q^2)_{\infty}}\equiv\varphi(-q)\cdot f(-q)\pmod{4}. \end{align} $$

For a prime $p\geq 5$ , $0\leq j\leq (p-1)/2$ , $-(p-1)/2\leq k\leq (p-1)/2$ , assume that

$$ \begin{align*} j^2+\dfrac{3k^2+k}{2}\equiv\dfrac{p^2-1}{24}\pmod{p}, \end{align*} $$

which implies that

$$ \begin{align*} 24j^2+(6k+1)^2\equiv0\pmod{p}. \end{align*} $$

Since $\big(\frac{-24}{p}\big)=-1$ , we get $j=0$ and $k=(\pm p-1)/6$ . Substituting (4.3) and (4.4) into (4.5), we find that

$$ \begin{align*} \sum_{n=0}^{\infty} Q{\bigg(pn+\dfrac{p^2-1}{24}\bigg)}q^n &\equiv(-1)^{(\pm p-1)/6}\varphi({-}q^p)f({-}q^p)\nonumber\\ &\equiv(-1)^{(\pm p-1)/6}\sum_{n=0}^{\infty} Q(n)q^{pn}\pmod{4}, \end{align*} $$

where we have used (4.5) in the last congruence. The congruence (1.9) follows. This completes the proof of Theorem 1.4.

5 Concluding remarks

One can use Lemma 3.1 to establish congruence relations satisfied by $Q(n)$ similar to (1.8) for other primes p. For example,

$$ \begin{align*} \sum_{n=0}^{\infty} Q(127n+672)q^n &\equiv\sum_{n=0}^{\infty} Q(n)q^{127n}\pmod{2^3},\\ \sum_{n=0}^{\infty} Q(131n+715)q^n &\equiv43\sum_{n=0}^{\infty} Q(n)q^{131n}\pmod{2^7},\\ \sum_{n=0}^{\infty} Q(137n+782)q^n &\equiv71\sum_{n=0}^{\infty} Q(n)q^{137n}\pmod{2^8},\\ \sum_{n=0}^{\infty} Q(139n+805)q^n &\equiv803\sum_{n=0}^{\infty} Q(n)q^{139n}\pmod{2^{10}}. \end{align*} $$

However, the corresponding bound $l_p$ will become much larger as p increases.

Merca [Reference Merca9] proved the following infinite family of congruences modulo $4$ for $Q(n)$ .

Theorem 5.1. Let $p\geq 5$ be a prime number such that $p\not \equiv 1\pmod {24}$ . Then for any $n\not \equiv 0\pmod {p}$ ,

(5.1) $$ \begin{align} Q{\bigg(pn+\dfrac{p^2-1}{24}\bigg)}\equiv0\pmod{4}. \end{align} $$

The congruence (1.8) together with numerical evidence suggests the following conjecture, which contains (1.9) and (5.1) as special cases.

Conjecture 5.2. Let $p\geq 5$ be a prime number such that $p\not \equiv 1\pmod {24}$ . Then

$$ \begin{align*} \sum_{n=0}^{\infty} Q{\bigg(pn+\dfrac{p^2-1}{24}\bigg)}q^n\equiv c_p\sum_{n=0}^{\infty} Q(n)q^{pn}\pmod{4}, \end{align*} $$

where $c_p=-1$ or $1$ .

Footnotes

The first author was partially supported by the National Natural Science Foundation of China (No. 12201177), the Natural Science Foundation of Hebei Province (No. A2021205018), the Science and Technology Research Project of Colleges and Universities of Hebei Province (No. BJK2023092), the Doctor Foundation of Hebei Normal University (No. L2021B02), the Program for Foreign Experts of Hebei Province and the Program for 100 Foreign Experts Plan of Hebei Province. The second author was partially supported by the National Natural Science Foundation of China (No. 12201093), the Natural Science Foundation Project of Chongqing CSTB (No. CSTB2022NSCQ–MSX0387), the Science and Technology Research Program of Chongqing Municipal Education Commission (No. KJQN202200509) and the Doctoral Start-up Research Foundation (No. 21XLB038) of Chongqing Normal University.

References

Andrews, G. E. and Berndt, B. C., Ramanujan’s Lost Notebook, Part II (Springer-Verlag, New York, 2005).CrossRefGoogle Scholar
Atkin, A. O. L., ‘Proof of a conjecture of Ramanujan’, Glasg. Math. J. 8 (1967), 1432.CrossRefGoogle Scholar
Berndt, B. C., Ramanujan’s Notebook, Part III (Springer-Verlag, New York, 1991).CrossRefGoogle Scholar
Berndt, B. C. and Ono, K., ‘Ramanujan’s unpublished manuscipt on the partition and tau functions with proofs and commentary’, Sém. Lothar. Combin. 42 (1999), Article no. B42c.Google Scholar
Cui, S.-P. and Gu, N. S. S., ‘Arithmetic properties of $\ell$ -regular partitions’, Adv. in Appl. Math. 51(4) (2013), 507523.CrossRefGoogle Scholar
Diamond, F. and Shurman, J., A First Course in Modular Forms, Graduate Texts in Mathematics, 228 (Springer-Verlag, New York, 2005).Google Scholar
Euler, L., Introductio in Analysin Infinitorum, Vol. 2 (MM Bousquet Hanson, Lausanne, 1748).Google Scholar
Gordon, B. and Ono, K., ‘Divisibility of certain partition functions by powers of primes’, Ramanujan J. 1(1) (1997), 2534.CrossRefGoogle Scholar
Merca, M., ‘Ramanujan-type congruences modulo 4 for partitions into distinct parts’, An. Şt. Univ. Ovidius Constanţa 30(3) (2022), 185199.Google Scholar
Ono, K., The Web of Modularity: Arithmetic of the Coefficients of Modular Forms and $q$ -Series , CBMS Regional Conference Series in Mathematics, 102 (American Mathematical Society, Providence, RI, 2004).Google Scholar
Paule, P. and Radu, S., ‘Partition analysis, modular functions, and computer algebra’, in: Recent Trends in Combinatorics, IMA Volumes in Mathematics and its Applications, 159 (eds. Beveridge, A., Griggs, J. R., Hogben, L., Musiker, G. and Tetali, P.) (Springer, Cham, 2016), 511543.CrossRefGoogle Scholar
Radu, C.-S., ‘An algorithmic approach to Ramanujan–Kolberg identities’, J. Symbolic Comput. 68(1) (2015), 225253.CrossRefGoogle Scholar
Smoot, N. A., ‘On the computation of identities relating partition numbers in arithmetic progressions with eta quotients: an implementation of Radu’s algorithm’, J. Symbolic Comput. 104 (2021), 276311.CrossRefGoogle Scholar
Sturm, J., ‘On the congruence of modular forms’, in: Number Theory (eds. D. V. Chudnovsky, G. V. Chudnovsky, H. Cohn and M. B. Nathanson) (Springer, Berlin, 1987) 275280.CrossRefGoogle Scholar
Watson, G. N., ‘Ramanujan’s vermutung über zerfällungsanzahlen’, J. reine angew. Math. 179 (1938), 97128.CrossRefGoogle Scholar
Figure 0

Table 1 A table of values of $c_p$.

Figure 1

Table 2 A table of values of $l_p$.