Hostname: page-component-78c5997874-4rdpn Total loading time: 0 Render date: 2024-10-28T06:24:10.279Z Has data issue: false hasContentIssue false

NEW GENERALISATIONS OF VAN HAMME’S (G.2) SUPERCONGRUENCE

Published online by Cambridge University Press:  18 May 2022

NA TANG*
Affiliation:
School of Mathematics and Statistics, Huaiyin Normal University, Huai’an 223300, Jiangsu, PR China
*
Rights & Permissions [Opens in a new window]

Abstract

Swisher [‘On the supercongruence conjectures of van Hamme’, Res. Math. Sci. 2 (2015), Article no. 18] and He [‘Supercongruences on truncated hypergeometric series’, Results Math. 72 (2017), 303–317] independently proved that Van Hamme’s (G.2) supercongruence holds modulo $p^4$ for any prime $p\equiv 1\pmod {4}$ . Swisher also obtained an extension of Van Hamme’s (G.2) supercongruence for $p\equiv 3 \pmod 4$ and $p>3$ . In this note, we give new one-parameter generalisations of Van Hamme’s (G.2) supercongruence modulo $p^3$ for any odd prime p. Our proof uses the method of ‘creative microscoping’ introduced by Guo and Zudilin [‘A q-microscope for supercongruences’, Adv. Math. 346 (2019), 329–358].

Type
Research Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

In his first letter to Hardy in 1913, Ramanujan mentioned the following formula (see [Reference Berndt and Rankin1, page 25, (2)]):

(1.1) $$ \begin{align} \sum_{k=0}^\infty(8k+1)\frac{(\frac{1}{4})_k^4}{k!^4} =\frac{2\sqrt{2}}{\sqrt{\pi}\,\Gamma(\frac 34)^2}. \end{align} $$

Here $(a)_n=a(a+1)\cdots (a+n-1)$ is the rising factorial and $\Gamma (x) $ is the Gamma function. Ramanujan did not give a proof of (1.1) and the first proof was given by Hardy [Reference Hardy7]. In 1997, Van Hamme [Reference Van Hamme12] conjectured 13 p-adic analogues of Ramanujan’s series, including

(1.2) $$ \begin{align} &\sum_{k=0}^{(p-1)/4}(8k+1)\frac{(\frac{1}{4})_k^4}{k!^4} \equiv p\frac{\Gamma_p(\frac 12)\Gamma_p(\frac 14)}{\Gamma_p(\frac 34)} \pmod{p^3} \quad\text{for}\ p\equiv 1 \pmod 4 \end{align} $$

(marked (G.2) in Van Hamme’s list). Here and throughout the paper, p always denotes an odd prime and $\Gamma _p(x)$ stands for Morita’s p-adic Gamma function [Reference Morita10]. Swisher [Reference Swisher11] and He [Reference He8] independently showed that (1.2) holds modulo the stronger power $p^4$ .

We shall give a generalisation of (1.2): for $p\equiv 1 \pmod 4$ and $0\leqslant s\leqslant (p-1)/4$ ,

(1.3) $$ \begin{align} \sum_{k=s}^{(p-1)/4} (8k+1)\frac{(\frac{1}{4})_{k-s}(\frac{1}{4})_{k+s}(\frac{1}{4})_k^2} {(k-s)!(k+s)!k!^2} \equiv (p+4s)\frac{(\frac{1}{4})_s^2(\frac{1}{4})_{(p-1)/4+s} (\frac{1}{2})_{(p-1)/4-s}} {s!^2(1)_{(p-1)/4+s}(\frac{1}{4})_{(p-1)/4-s}} \pmod{p^3}. \end{align} $$

When $s=0$ , the right-hand side of (1.3) reduces to $p(\tfrac 12)_{(p-1)/4}/(1)_{(p-1)/4}$ , which is congruent to the right-hand side of (1.2) modulo $p^3$ (see [Reference Liu and Wang9]). Thus, the supercongruence (1.3) is indeed a generalisation of (1.2). A similar extension of the (A.2) supercongruence of Van Hamme was recently given by Guo [Reference Guo3].

We shall prove (1.3) by establishing the following q-supercongruence.

Theorem 1.1. Let $n\equiv 1\pmod 4$ be an integer greater than $1$ and let $0\leqslant s\leqslant(n-1)/4$ . Then

(1.4) $$ \begin{align} &\sum_{k=s}^{(n-1)/4} [8k+1]\frac{(q;q^4)_{k-s}(q;q^4)_{k+s}(q;q^4)_k^2} {(q^4;q^4)_{k-s}(q^4;q^4)_{k+s}(q^4;q^4)_k^2} q^{2k} \notag\\[6pt] &\quad\equiv [n+4s] \frac{(q;q^4)_s^2(q;q^4)_{(n-1)/4+s} (q^2;q^4)_{(n-1)/4-s}} {(q^4;q^4)_s^2(q^4;q^4)_{(n-1)/4+s}(q;q^4)_{(n-1)/4-s}} q^{3s+(1-n)/4} \pmod{\Phi_n(q)^3}. \end{align} $$

Here we need to be familiar with the standard q-notation. The q-shifted factorial is defined by $(a;q)_0=1$ and $(a;q)_n=(1-a)(1-aq)\cdots (1-aq^{n-1})$ for any positive integer n. The q-integer is defined as $[n]=(1-q^n)/(1-q)$ and $\Phi _n(q)$ denotes the nth cyclotomic polynomial, which can be written as

$$ \begin{align*} \Phi_n(q)=\prod_{\substack{1\leqslant k\leqslant n\\ \gcd(k,n)=1}}(q-\zeta^k), \end{align*} $$

where $\zeta $ is a primitive nth root of unity.

It is easy to see that (1.3) follows from (1.4) by taking $n=p$ and $q\to 1$ . The $s=0$ case of (1.4) was given by Liu and Wang [Reference Liu and Wang9] and can also be deduced from [Reference Guo and Schlosser5, Theorem 4.3].

Swisher [Reference Swisher11, (3)] gave the following extension of Van Hamme’s (G.2) supercongruence: for $p\equiv 3 \pmod 4$ and $p>3$ ,

(1.5) $$ \begin{align} &\sum_{k=0}^{(3p-1)/4}(8k+1)\frac{(\frac{1}{4})_k^4}{k!^4} \equiv -\frac{3p^2\Gamma_p(\frac 12)\Gamma_p(\frac 14)}{2\Gamma_p(\frac 34)} \pmod{p^4}. \end{align} $$

(The negative sign was missing in Swisher’s original supercongruence.)

We shall give a new generalisation of (1.5) modulo $p^3$ as follows: for $p\equiv 3\pmod {4}$ and $0\leqslant s\leqslant (p-3)/4$ ,

(1.6) $$ \begin{align} \sum_{k=s}^{(3p-1)/4} (8k+1)\frac{(\frac{1}{4})_{k-s}(\frac{1}{4})_{k+s}(\frac{1}{4})_k^2} {(k-s)!(k+s)!k!^2} \equiv (3p+4s)\frac{(\frac{1}{4})_s^2(\frac{1}{4})_{(3p-1)/4+s} (\frac{1}{2})_{(3p-1)/4-s}} {s!^2(1)_{(3p-1)/4+s}(\frac{1}{4})_{(3p-1)/4-s}} \pmod{p^3}. \end{align} $$

When $s=0$ , the right-hand side of (1.6) reduces to $3p(\tfrac 12)_{(3p-1)/4}/(1)_{(3p-1)/4}$ , which is easily seen to be congruent to the right-hand side of (1.5) modulo $p^3$ . Thus, the supercongruence (1.6) can be deemed a generalisation of the modulo $p^3$ case of (1.5). A result of Guo and Schlosser [Reference Guo and Schlosser5, Corollary 1.2 with $d = 4$ and $q\to 1$ ] implies that (1.6) is even true modulo $p^4$ for $s=0$ . However, numerical evaluation indicates that (1.6) is not true modulo $p^4$ for general s.

In the same way as before, we shall prove (1.6) via the following q-supercongruence.

Theorem 1.2. Let $n\equiv 3\pmod 4$ be a positive integer and let $0\leqslant s\leqslant (n-3)/4$ . Then

(1.7) $$ \begin{align} &\sum_{k=s}^{(3n-1)/4} [8k+1]\frac{(q;q^4)_{k-s}(q;q^4)_{k+s}(q;q^4)_k^2} {(q^4;q^4)_{k-s}(q^4;q^4)_{k+s}(q^4;q^4)_k^2} q^{2k} \notag\\[6pt] &\quad\equiv [3n+4s] \frac{(q;q^4)_s^2(q;q^4)_{(3n-1)/4+s} (q^2;q^4)_{(3n-1)/4-s}} {(q^4;q^4)_s^2(q^4;q^4)_{(3n-1)/4+s}(q;q^4)_{(3n-1)/4-s}} q^{3s+(1-3n)/4} \pmod{\Phi_n(q)^3}. \end{align} $$

Our proof of Theorems 1.1 and 1.2 will use the powerful method of ‘creative microscoping’, which was devised by Guo and Zudilin [Reference Guo and Zudilin6].

2 Proof of Theorem 1.1

We require the following easily proved q-congruence, which was first given by Guo and Schlosser [Reference Guo and Schlosser4, Lemma 3].

Lemma 2.1. Let d, m and n be positive integers with $m\leqslant n-1$ and $dm\equiv -1\pmod {n}$ . Then, for $0\leqslant k\leqslant m$ ,

$$ \begin{align*} \frac{(aq;q^d)_{m-k}}{(q^d/a;q^d)_{m-k}} \equiv (-a)^{m-2k}\frac{(aq;q^d)_k}{(q^d/a;q^d)_k} q^{m(dm-d+2)/2+(d-1)k} \pmod{\Phi_n(q)}. \end{align*} $$

Following Gasper and Rahman’s monograph [Reference Gasper and Rahman2], the basic hypergeometric series $_{r+1}\phi _r$ is defined as

$$ \begin{align*}_{r+1}\phi_{r}\bigg[\begin{array}{@{}c@{}} a_1,a_2,\ldots,a_{r+1}\\[6pt] b_1,b_2,\ldots,b_{r} \end{array};q,\, z \bigg] =\sum_{k=0}^{\infty}\frac{(a_1;q)_k(a_2;q)_k\cdots(a_{r+1};q)_k } {(q;q)_k(b_1;q)_k\cdots(b_{r};q)_k}z^k. \end{align*} $$

Jackson’s ${}_6\phi _5$ summation (see [Reference Gasper and Rahman2, Appendix (II.21)]) can be stated as follows:

(2.1) $$ \begin{align} _6\phi_5 \bigg[\begin{array}{@{}c@{}} a,\, qa^{1/2},\, -qa^{1/2},\, b,\, c,\, q^{-n} \\[6pt] a^{1/2},\, -a^{1/2},\, aq/b,\, aq/c,\, aq^{n+1}\end{array};q,\,\frac{aq^{n+1}}{bc} \bigg] =\frac{(aq;q)_n (aq/bc;q)_n}{(aq/b;q)_n(aq/c;q)_n}. \end{align} $$

To prove Theorem 1.1, we first establish the following result.

Theorem 2.2. Let $n\equiv 1\pmod 4$ be an integer greater than $1$ . Let $0\leqslant s\leqslant (n-1)/4$ and let a be an indeterminate. Then, modulo $\Phi _n(q)(1-aq^n)(a-q^n)$ ,

(2.2) $$ \begin{align} &\sum_{k=s}^{(n-1)/4} [8k+1]\frac{(aq;q^4)_k(q/a;q^4)_k (q;q^4)_{k-s}(q;q^4)_{k+s} } {(aq^4;q^4)_k(q^4/a;q^4)_k(q^4;q^4)_{k-s}(q^4;q^4)_{k+s}}q^{2k} \notag\\[6pt] &\quad\equiv [n+4s] \frac{(aq;q^4)_{s}(q/a;q^4)_{s}(q;q^4)_{(n-1)/4+s} (q^2;q^4)_{(n-1)/4-s}} {(aq^4;q^4)_{s}(q^4/a;q^4)_{s}(q^4;q^4)_{(n-1)/4+s}(q;q^4)_{(n-1)/4-s}} q^{3s+(1-n)/4}. \end{align} $$

Proof. For $a=q^{-n}$ or $a=q^{n}$ , the left-hand side of (2.2) is equal to

(2.3) $$ \begin{align} &\sum_{k=s}^{(n-1)/4} [8k+1]\frac{(q^{1-n};q^4)_k(q^{1+n};q^4)_k (q;q^4)_{k-s}(q;q^4)_{k+s} } {(q^{4-n};q^4)_k(q^{4+n};q^4)_k(q^4;q^4)_{k-s}(q^4;q^4)_{k+s}}q^{2k} \notag\\[6pt] &\quad=\sum_{k=0}^{(n-1)/4-s} [8k+8s+1]\frac{(q^{1-n};q^4)_{k+s}(q^{1+n};q^4)_{k+s} (q;q^4)_{k}(q;q^4)_{k+2s} } {(q^{4-n};q^4)_{k+s}(q^{4+n};q^4)_{k+s}(q^4;q^4)_{k}(q^4;q^4)_{k+2s}}q^{2k+2s} \notag\\[6pt] &\quad=[8s+1] \frac{(q^{1-n};q^4)_{s}(q^{1+n};q^4)_{s} (q;q^4)_{2s} } {(q^{4-n};q^4)_{s}(q^{4+n};q^4)_{s}(q^4;q^4)_{2s}} q^{2s}\notag\\[6pt] &\quad\quad\times {}_{6}\phi_{5}\!\left[\begin{array}{@{}cccccccc@{}} q^{1+8s},&\!\! q^{\frac{9}{2}+4s},&\!\! -q^{\frac{9}{2}+4s}, &\!\! q, &\!\! q^{1+n+4s}, &\!\! q^{1-n+4s} \\[8pt] &\!\! q^{\frac{1}{2}+4s}, &\!\! -q^{\frac{1}{2}+4s}, &\!\! q^{4+8s}, &\!\! q^{4-n+4s}, &\!\! q^{4+n+4s} \end{array}\!\!;q^4, q^2 \right]. \end{align} $$

Letting $q\mapsto q^4$ , $a=q^{1+8s}$ , $b=q$ , $c= q^{1+n+4s}$ and $n\mapsto (n-1)/4-s$ in (2.1), one sees that the right-hand side of (2.3) can be simplified as

$$ \begin{align*} &q^{2s}[8s+1] \frac{(q^{1-n};q^4)_{s}(q^{1+n};q^4)_{s} (q;q^4)_{2s} } {(q^{4-n};q^4)_{s}(q^{4+n};q^4)_{s}(q^4;q^4)_{2s}} \frac{(q^{5+8s};q^4)_{(n-1)/4-s} (q^{3-n+4s};q^4)_{(n-1)/4-s}}{(q^{4+8s};q^4)_{(n-1)/4-s}(q^{4-n+4s};q^4)_{(n-1)/4-s}} \\[6pt] &\quad=[n+4s]\frac{(q^{1-n};q^4)_{s}(q^{1+n};q^4)_{s} } {(q^{4-n};q^4)_{s}(q^{4+n};q^4)_{s}} \frac{(q;q^4)_{(n-1)/4+s} (q^{3-n+4s};q^4)_{(n-1)/4-s}}{(q^4;q^4)_{(n-1)/4+s}(q^{4-n+4s};q^4)_{(n-1)/4-s}} q^{2s}\\[6pt] &\quad=[n+4s]\frac{(q^{1-n};q^4)_{s}(q^{1+n};q^4)_{s} } {(q^{4-n};q^4)_{s}(q^{4+n};q^4)_{s}} \frac{(q;q^4)_{(n-1)/4+s} (q^2;q^4)_{(n-1)/4-s}}{(q^4;q^4)_{(n-1)/4+s}(q;q^4)_{(n-1)/4-s}} q^{3s+(1-n)/4}. \end{align*} $$

Thus, we have proved that (2.2) is true modulo $1-aq^n$ and $a-q^n$ .

Since $n\equiv 1\pmod {4}$ , letting $d=4$ and $m=(n-1)/4$ in Lemma 2.1, we obtain

(2.4) $$ \begin{align} \frac{(aq;q^4)_{m-k}}{(q^4/a;q^d)_{m-k}} \equiv (-a)^{m-2k}\frac{(aq;q^4)_k}{(q^4/a;q^d)_k} q^{m(2m-1)+3k} \pmod{\Phi_n(q)} \end{align} $$

for $0\leqslant k\leqslant m$ . Using this q-congruence, we can easily verify the following congruence, for $m=(n-1)/4$ and $s\leqslant k\leqslant m-s$ ,

(2.5) $$ \begin{align} &[8(m-k)+1]\frac{(aq;q^4)_{m-k}(q/a;q^4)_{m-k} (q;q^4)_{m-k-s}(q;q^4)_{m-k+s} } {(aq^4;q^4)_{m-k}(q^4/a;q^4)_{m-k}(q^4;q^4)_{m-k-s}(q^4;q^4)_{m-k+s}} q^{2m-2k} \notag\\[6pt] &\quad\equiv -[8k+1]\frac{(aq;q^4)_k(q/a;q^4)_k (q;q^4)_{k-s}(q;q^4)_{k+s} } {(aq^4;q^4)_k(q^4/a;q^4)_k(q^4;q^4)_{k-s}(q^4;q^4)_{k+s} } q^{2k} \pmod{\Phi_n(q)}. \end{align} $$

Moreover, for $(n-1)/4-s<k\leqslant (n-1)/4$ , the summand indexed k on the left-hand side of (2.2) is congruent to $0$ modulo $\Phi _n(q)$ because $k+s>(n-1)/4$ and $(q;q^4)_{k+s}$ in the numerator incorporates the factor $1-q^n$ . This means that the left-hand side of (2.2) is congruent to $0$ modulo $\Phi _n(q)$ . Since

$$ \begin{align*}[n+4s](q;q^4)_{(n-1)/4+s}=[n](q;q^4)_{(n-1)/4}(q^{n+4};q^4)_s\equiv 0\pmod{\Phi_n(q)}\end{align*} $$

for $n>1$ , we conclude that (2.2) is also true modulo $\Phi _n(q)$ . Noting that the polynomials $1-aq^n$ , $a-q^n$ and $\Phi _n(q)$ are pairwise relatively prime, we complete the proof of the theorem.

Proof of Theorem 1.1.

For $a=1$ , the denominators on both sides of (2.2) are relatively prime to $\Phi _n(q)$ . Moreover, when $a=1$ the polynomial $(1-aq^n)(a-q^n)=(1-q^n)^2$ contains the factor $\Phi _n(q)^2$ . Therefore, putting $a=1$ in (2.2), we obtain the desired q-supercongruence (1.4).

3 Proof of Theorem 1.2

The proof is similar to that of Theorem 1.1. We first establish the following parametric generalisation of Theorem 1.2.

Theorem 3.1. Let $n\equiv 3\pmod 4$ be a positive integer. Let $0\leqslant s\leqslant (n-3)/4$ and let a be an indeterminate. Then, modulo $\Phi _n(q)(1-aq^{3n})(a-q^{3n})$ ,

(3.1) $$ \begin{align} &\sum_{k=s}^{(3n-1)/4} [8k+1]\frac{(aq;q^4)_k(q/a;q^4)_k (q;q^4)_{k-s}(q;q^4)_{k+s} } {(aq^4;q^4)_k(q^4/a;q^4)_k(q^4;q^4)_{k-s}(q^4;q^4)_{k+s}}q^{2k} \notag\\[6pt] &\quad\equiv [n+4s] \frac{(aq;q^4)_{s}(q/a;q^4)_{s}(q;q^4)_{(3n-1)/4+s} (q^2;q^4)_{(3n-1)/4-s}} {(aq^4;q^4)_{s}(q^4/a;q^4)_{s}(q^4;q^4)_{(3n-1)/4+s}(q;q^4)_{(3n-1)/4-s}} q^{3s+(1-3n)/4}. \end{align} $$

Proof. For $a=q^{-3n}$ or $a=q^{3n}$ , the left-hand side of (3.1) is equal to

(3.2) $$ \begin{align} &\sum_{k=s}^{(3n-1)/4} [8k+1]\frac{(q^{1-3n};q^4)_k(q^{1+3n};q^4)_k (q;q^4)_{k-s}(q;q^4)_{k+s} } {(q^{4-3n};q^4)_k(q^{4+3n};q^4)_k(q^4;q^4)_{k-s}(q^4;q^4)_{k+s}}q^{2k} \notag\\[6pt] &\quad=[8s+1] \frac{(q^{1-3n};q^4)_{s}(q^{1+3n};q^4)_{s} (q;q^4)_{2s} } {(q^{4-3n};q^4)_{s}(q^{4+3n};q^4)_{s}(q^4;q^4)_{2s}} q^{2s}\notag\\[6pt] &\quad\quad\times {}_{6}\phi_{5}\!\left[\begin{array}{@{}cccccccc@{}} q^{1+8s},&\!\! q^{\frac{9}{2}+4s},&\!\! -q^{\frac{9}{2}+4s}, &\!\! q, &\!\! q^{1+3n+4s}, &\!\! q^{1-3n+4s} \\[6pt] &\!\! q^{\frac{1}{2}+4s}, &\!\! -q^{\frac{1}{2}+4s}, &\!\! q^{4+8s}, &\!\! q^{4-3n+4s}, &\!\! q^{4+3n+4s} \end{array}\!\!;q^4, q^2 \right]. \end{align} $$

Letting $q\mapsto q^4$ , $a=q^{1+8s}$ , $b=q$ , $c= q^{1+3n+4s}$ and $n\mapsto (3n-1)/4-s$ in (2.1), one sees that the right-hand side of (3.2) may be written as

$$ \begin{align*} &q^{2s}[8s+1] \frac{(q^{1-3n};q^4)_{s}(q^{1+3n};q^4)_{s} (q;q^4)_{2s} } {(q^{4-3n};q^4)_{s}(q^{4+3n};q^4)_{s}(q^4;q^4)_{2s}} \frac{(q^{5+8s};q^4)_{(3n-1)/4-s} (q^{3-3n+4s};q^4)_{(3n-1)/4-s}}{(q^{4+8s};q^4)_{(3n-1)/4-s}(q^{4-3n+4s};q^4)_{(3n-1)/4-s}} \\[6pt] &=[3n+4s]\frac{(q^{1-3n};q^4)_{s}(q^{1+3n};q^4)_{s} } {(q^{4-3n};q^4)_{s}(q^{4+3n};q^4)_{s}} \frac{(q;q^4)_{(3n-1)/4+s} (q^2;q^4)_{(3n-1)/4-s}}{(q^4;q^4)_{(3n-1)/4+s}(q;q^4)_{(3n-1)/4-s}} q^{3s+(1-3n)/4}. \end{align*} $$

This proves that (2.2) holds modulo $1-aq^{3n}$ and $a-q^{3n}$ .

Since $n\equiv 3\pmod {4}$ , letting $d=4$ and $m=(3n-1)/4$ in Lemma 2.1, we again obtain (2.4) for $0\leqslant k\leqslant m$ . Applying this q-congruence, we can check (2.5) for $m=(3n-1)/4$ and $s\leqslant k\leqslant m-s$ . For $(3n-1)/4-s<k\leqslant (3n-1)/4$ , the summand indexed k on the left-hand side of (3.1) is congruent to $0$ modulo $\Phi _n(q)$ because $k+s>(3n-1)/4$ and $(q;q^4)_{k+s}$ has the factor $1-q^{3n}$ . This implies that the left-hand side of (3.1) is congruent to $0$ modulo $\Phi _n(q)$ . Since

$$ \begin{align*}[3n+4s](q;q^4)_{(3n-1)/4+s}=[3n](q;q^4)_{(3n-1)/4}(q^{3n+4};q^4)_s\equiv 0\pmod{\Phi_n(q)},\end{align*} $$

we conclude that (3.1) is also true modulo $\Phi _n(q)$ .

Proof of Theorem 1.1.

When $a=1$ , the polynomial $(1-aq^{3n})(a-q^{3n})=(1-q^{3n})^2$ contains the factor $\Phi _n(q)^2$ . Thus, letting $a=1$ in (2.2), we get the q-supercongruence (1.7).

4 An open problem

We believe that the following conjecture is true.

Conjecture 4.1. The q-supercongruences (1.4) and (1.7) are also true modulo $[n]\Phi _n(q)^2$ .

The above conjecture is clearly true for $s=0$ (see [Reference Guo and Schlosser5, Reference Liu and Wang9]). Numerical computation indicates that both sides of (1.4) (or (1.7)) should be congruent to $0$ modulo $[n]$ . However, it seems difficult to confirm this. The technique of proving a q-congruence modulo $[n]$ introduced in [Reference Guo and Zudilin6] does not work here, because of the additional parameter s.

References

Berndt, B. C. and Rankin, R. A., Ramanujan, Letters and Commentary, History of Mathematics, 9 (American Mathematical Society, Providence, RI, 1995).10.1090/hmath/009CrossRefGoogle Scholar
Gasper, G. and Rahman, M., Basic Hypergeometric Series, 2nd edn, Encyclopedia of Mathematics and Its Applications, 96 (Cambridge University Press, Cambridge, 2004).10.1017/CBO9780511526251CrossRefGoogle Scholar
Guo, V. J. W., ‘A new extension of the (A.2) supercongruence of Van Hamme’, Results Math. 77 (2022), Article no. 96.CrossRefGoogle Scholar
Guo, V. J. W. and Schlosser, M. J., ‘A family of $q$ -hypergeometric congruences modulo the fourth power of a cyclotomic polynomial’, Israel J. Math. 240 (2020), 821835.CrossRefGoogle Scholar
Guo, V. J. W. and Schlosser, M. J., ‘A new family of $q$ -supercongruences modulo the fourth power of a cyclotomic polynomial’, Results Math. 75 (2020), Article no. 155.Google ScholarPubMed
Guo, V. J. W. and Zudilin, W., ‘A $q$ -microscope for supercongruences’, Adv. Math. 346 (2019), 329358.CrossRefGoogle Scholar
Hardy, G. H., ‘A chapter from Ramanujan’s note-book’, Math. Proc. Cambridge Philos. Soc. 21(2) (1923), 492503.Google Scholar
He, B., ‘Supercongruences on truncated hypergeometric series’, Results Math. 72 (2017), 303317.10.1007/s00025-016-0635-7CrossRefGoogle Scholar
Liu, Y. and Wang, X., ‘ $q$ -Analogues of the (G.2) supercongruence of Van Hamme’, Rocky Mountain J. Math. 51 (2021), 13291340.10.1216/rmj.2021.51.1329CrossRefGoogle Scholar
Morita, Y., ‘A $p$ -adic supercongruence of the $\varGamma$ function’, J. Fac. Sci. Univ. Tokyo 22 (1975), 255266.Google Scholar
Swisher, H., ‘On the supercongruence conjectures of van Hamme’, Res. Math. Sci. 2 (2015), Article no. 18.10.1186/s40687-015-0037-6CrossRefGoogle Scholar
Van Hamme, L., ‘Some conjectures concerning partial sums of generalized hypergeometric series’, in $p$ -Adic Functional Analysis, Nijmegen, 1996, Lecture Notes in Pure and Applied Mathematics, 192 (eds. W. H. Schikhof, C. Perez-Garcia and J. Kakol) (Dekker, New York, 1997), 223236.Google Scholar