Hostname: page-component-7bb8b95d7b-5mhkq Total loading time: 0 Render date: 2024-09-27T02:23:45.596Z Has data issue: false hasContentIssue false

NEW CONGRUENCES FOR THE TRUNCATED APPELL SERIES $F_1$

Published online by Cambridge University Press:  18 April 2024

XIAOXIA WANG*
Affiliation:
Department of Mathematics, Shanghai University, Newtouch Center for Mathematics of Shanghai University, Shanghai 200444, PR China
WENJIE YU
Affiliation:
Department of Mathematics, Shanghai University, Newtouch Center for Mathematics of Shanghai University, Shanghai 200444, PR China e-mail: wenjieyu@shu.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

Liu [‘Supercongruences for truncated Appell series’, Colloq. Math. 158(2) (2019), 255–263] and Lin and Liu [‘Congruences for the truncated Appell series $F_3$ and $F_4$’, Integral Transforms Spec. Funct. 31(1) (2020), 10–17] confirmed four supercongruences for truncated Appell series. Motivated by their work, we give a new supercongruence for the truncated Appell series $F_{1}$, together with two generalisations of this supercongruence, by establishing its q-analogues.

Type
Research Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1. Introduction

In 1880, Appell defined four kinds of double series $F_{1}$ , $F_{2}$ , $F_{3}$ , $F_{4}$ in two variables (see [Reference Slater12, pages 210–211]) by generalising the Gauss hypergeometric series. These four series, called Appell series, are well known in the field of double hypergeometric series.

Based on the definition of the truncated hypergeometric series, Liu [Reference Liu9] introduced the truncated Appell series, defined by

$$ \begin{align*} F_{1} [a;b,b{'} ;c;x,y]_{n}&=\sum_{i=0}^{n} \sum_{j=0}^{n}\frac{(a)_{i+j}(b)_{i} (b{'})_{j}}{(c)_{i+j}} \cdot \frac{x^{i}y^{j}}{i!j!}; \\ F_{2} [a;b,b{'} ;c,c{'};x,y]_{n}&=\sum_{i=0}^{n} \sum_{j=0}^{n}\frac{(a)_{i+j}(b)_{i} (b{'})_{j}}{(c)_{i}(c{'})_{j}} \cdot \frac{x^{i}y^{j}}{i!j!}; \\ F_{3} [a,a{'};b,b{'} ;c;x,y]_{n}&=\sum_{i=0}^{n} \sum_{j=0}^{n}\frac{(a)_{i}(a{'})_{j}(b)_{i} (b{'})_{j}}{(c)_{i+j}} \cdot \frac{x^{i}y^{j}}{i!j!}; \\ F_{4} [a;b;c,c{'};x,y]_{n}&=\sum_{i=0}^{n} \sum_{j=0}^{n}\frac{(a)_{i+j}(b)_{i+j}}{(c)_{i}(c{'})_{j}} \cdot \frac{x^{i}y^{j}}{i!j!}, \end{align*} $$

where $(x)_n$ is the shifted factorial $(x)_{n} =x(x+1)\cdots (x+n-1)$ with $n \in \mathbb {Z}^{+}$ and $(x)_{0}=1$ .

In [Reference Liu9], Liu confirmed two congruences for the truncated Appell series $F_{1}$ and $F_{2}$ by using combinatorial identities: for any prime $p \ge $ 5, modulo $p^{2}$ ,

$$ \begin{align*} F_1[\tfrac12;\tfrac12,\tfrac12;1;1,1]_{(p-1)/2} &\equiv 1 \\ F_2[\tfrac12;\tfrac12,\tfrac12;1,1;1,1]_{(p-1)/2} &\equiv\begin{cases} -\Gamma_p(\frac{1}{4})^{4} & \mbox{if } p \equiv 1 \pmod{4}, \\ 0 & \mbox{if } p \equiv 3 \pmod{4}. \end{cases} \end{align*} $$

Later, Lin and Liu [Reference Lin and Liu8] studied congruence properties of the truncated Appell series $F_{3}$ and $F_{4}$ : for any prime $p \ge $ 5, modulo $p^{2}$ ,

(1.1) $$ \begin{align} F_3[\tfrac12,\tfrac12;\tfrac12,\tfrac12;1;1,1]_{(p-1)/2} &\equiv (-1)^{{p-1}/{2}},\\F_4[\tfrac12;\tfrac12;1,1;1,1]_{(p-1)/2} &\equiv \begin{cases} (-1)^{{p+1}/{2} }\Gamma_p(\frac{1}{6})^{2}\Gamma_p(\frac{1}{3}) ^{2} & \mbox{if } p \equiv 1 \pmod{3}, \\ 0 & \mbox{if } p \equiv 2 \pmod{3}.\notag \end{cases} \end{align} $$

Here $\Gamma _p$ is the p-adic Gamma function for p an odd prime, given by

$$ \begin{align*} \Gamma_p(\alpha)=\lim_{n\rightarrow \alpha}(-1)^n\prod_{\substack{0<j<n\\ p\nmid j}}j \end{align*} $$

for $\alpha \in \mathbb Z_p$ , and $\mathbb Z_p$ denotes the ring of all p-adic integers.

Recently, Wang and Yu [Reference Wang and Yu14] gave a generalisation of (1.1) with one free parameter d by establishing a q-supercongruence: for n a positive odd integer and d an integer with $n\ge \max \{2d+1, 1-2d \}$ , modulo $\Phi _n(q)^{4}$ ,

$$ \begin{align*} &\sum_{i=0}^{{(n-1)}/{2}-d}\sum_{j=0}^{{(n-1)}/{2}+d}\frac{(q^{2d+1};q^2)_{i}^2(q^{1-2d};q^2)_{j}^2}{(q^2;q^2)_i(q^2;q^2)_j(q^2;q^2)_{i+j}}q^{2ij-4di+4dj} \\ & \equiv \left \{\begin{aligned} & (-1)^{{(n-1)}/{2}}q^{{(1-n^2)}/{4}}, & d = 0, \\ & (1-q^n)^2q^{|d|(2+3|d|-n)-n+{(1-n^2)}/{4}}\sum_{k=1}^{2|d|}(-1)^{k-|d|+{(n-1)}/{2}}q^{k^2-k}H_k(-2|d|-1)\\ & \quad\quad\times\frac{(q^{n+2|d|-2k+1};q^2)_k(q^{4|d|-2k+2};q^2)_{(n-2|d|-1)/2}}{(q^2;q^2)_k(q^2;q^2)_{(n-2|d|-1)/2}}, & d\neq 0,\end{aligned}\right.\end{align*} $$

where $H_k(x)=\sum _{t=1}^{k}{q^{2t+x}}/{(1-q^{2t+x})^2}$ , $k \in \mathbb {Z}^{+}$ . The q-shifted factorial is defined by $(a;q)_{0}=1$ and $(a;q)_{n}=(1-a)(1-aq)\cdots (1-aq^{n-1})$ with $n \in \mathbb {Z}^{+}$ ; the q-integer is $[n]=[n]_{q}=(q^n-1)/(q-1)$ and $\Phi _n(q)$ denotes the nth cyclotomic polynomial in q, which can be factorised as

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

with $\zeta $ a primitive nth root of unity. In addition, the q-binomial coefficient is defined by

$$ \begin{align*} {x\brack k}={x\brack k}_{q} = \begin{cases} & \frac{(q^{1+x-k};q)_k}{(q;q)_k}, & k \geq 0, & \\ & 0, & k < 0. & \end{cases}\end{align*} $$

Inspired by the work mentioned above, and recent progress on congruences and q-congruences (see [Reference Guo2Reference He and Wang7, Reference Liu and Petrov10, Reference Liu and Wang11, Reference Wang and Xu13Reference Wang and Yu15]), we continue the study of congruences for the truncated Appell series $F_{1}$ and obtain new results.

Theorem 1.1. Let p be a prime with $p \equiv 1\pmod {4}$ . Then

$$ \begin{align*} F_1[\tfrac12;\tfrac14,\tfrac14;1;1,1]_{(p-1)/2} \equiv 1 \pmod {p^2}. \end{align*} $$

We establish two generalised q-analogues of Theorem 1.1.

Theorem 1.2. Let d and n be positive integers with $n\equiv 1 \pmod {2d}$ . Then, modulo $\Phi _n(q)^{2}$ ,

(1.2) $$ \begin{align} \sum_{i=0}^{{(n-1)}/{d}}\sum_{j=0}^{{(n-1)}/{d}}\frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj} \equiv \begin{bmatrix} -\dfrac{2}{d} \\[2mm] \dfrac{n-1}{d} \end{bmatrix} _{q^{2d}}q^{{(2n^{2}-2)}/{d}}. \end{align} $$

Theorem 1.3. Let d be an even positive integer and n a positive integer with $n \equiv d-1 \pmod {2d}$ . Then, modulo $\Phi _n(q)^{2}$ ,

$$ \begin{align*} &\sum_{i=0}^{{(n-(d-1))}/{d}}\sum_{j=0}^{{(n-(d-1))}/{d}}\frac{(q^{2d-2};q^{2d})_{i+j}(q^{d-1};q^{2d})_{i}(q^{d-1};q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}} q^{3di-i+2dj+2d-4} \notag \\ &\qquad \equiv \begin{bmatrix} \dfrac{2}{d} -2 \\[2mm] \dfrac{n-(d-1)}{d} \end{bmatrix} _{q^{2d}}q^{{(2n^{2}-2)}/{d}}. \end{align*} $$

Letting n be a prime p and then taking $q \rightarrow 1$ in Theorems 1.2 and 1.3 gives the following conclusions.

Corollary 1.4. Let p be a prime and d a positive integer with $p \equiv 1 \pmod {2d}$ . Then

$$ \begin{align*} F_1\bigg[\frac1d;\frac{1}{2d},\frac{1}{2d};1;1,1\bigg]_{(p-1)/d} \equiv \begin{pmatrix} -\dfrac{2}{d} \\[2mm] \dfrac{p-1}{d} \end{pmatrix} \pmod {p^2}. \end{align*} $$

Corollary 1.5. Let p be a prime and d an even positive integer with $p \equiv d-1 \pmod {2d}$ . Then

$$ \begin{align*} F_1\bigg[\frac{d-1}{d};\frac{d-1}{2d},\frac{d-1}{2d};1;1,1\bigg]_{(p-(d-1))/d} \equiv \begin{pmatrix} \dfrac{2}{d}-2 \\[2mm] \dfrac{p-(d-1)}{d} \end{pmatrix} \pmod {p^2}. \end{align*} $$

Theorem 1.1 is the special case $d=2$ of Corollaries 1.4 and 1.5. In the following two sections, we give the proofs of Theorems 1.2 and 1.3.

The famous q-Chu–Vandermonde identity [Reference Gasper and Rahman1, (1.5.2)] can be converted to

(1.3) $$ \begin{align}\sum_{k=0}^{n} {x\brack k} {y\brack n-k}q^{(x-k)(n-k)} ={x+y\brack n}, \end{align} $$

and this will be frequently used in our proofs.

2. Proof of Theorem 1.2

Since $n \equiv 1\pmod {2d}$ , we have $\gcd (2d,n) = 1$ . Hence, $(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}$ is relatively prime to $\Phi _n(q)$ for $0 \le i+j \le n-1$ . Also, $(q^{2};q^{2d})_{i+j} \equiv $ 0 (mod $\Phi _n(q)$ ) for ${(n-1}/{d} +1\le i\le {2(n-1)}/{d}$ and $(q;q^{2d})_{i} \equiv $ 0 (mod $\Phi _n(q)$ ) for ${(n-1)}/{2d} +1\le i\le {2(n-1)}/{d}$ . So,

$$ \begin{align*} \frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}} \equiv 0 \pmod {\Phi_n(q)^{2}} \quad \text{when } \frac{n-1}{d} +1\le i,i+j\le \frac{2(n-1)}{d}. \end{align*} $$

By symmetry, also

$$ \begin{align*} \frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{j}} \equiv 0 \pmod {\Phi_n(q)^{2}} \quad \text{when } \frac{n-1}{d} +1\le j,i+j\le \frac{2(n-1)}{d}. \end{align*} $$

Now, the left-hand side of (1.2) can be evaluated as

$$ \begin{align*} &\sum_{i=0}^{{(n-1)}/{d}} \sum_{j=0}^{({n-1)}/{d}}\frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj}\notag\ \\[5pt] &\qquad\equiv \sum_{i=0}^{{(n-1)}/{d}}\sum_{j=0}^{{(n-1)}/{d}}\frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj}\notag\ \\[5pt] &\qquad\quad +\sum_{{(n-1)}/{d}+1\le i,i+j\le {2(n-1)}/{d}}^{ }\frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj} \notag\ \\[5pt] &\qquad\quad +\sum_{{(n-1)}/{d}+1\le j,i+j\le {2(n-1)}/{d}}^{ }\frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj} \notag\ \\[5pt]&\qquad\equiv \sum_{0 \le i+j \le {2(n-1)}/{d}}^{}\frac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj}\notag\ \\ &\qquad\equiv \sum_{m=0}^{{2(n-1)}/{d} } \begin{bmatrix} -\dfrac{1}{d} \\ m \end{bmatrix}_{q^ {2d}}\sum_{i=0}^{m} \begin{bmatrix} -\dfrac{1}{2d} \\ i \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{1}{2d} \\ m-i \end{bmatrix}_{q^{2d}}q^{i+3m+2di^{2}-2dmi+2dm^{2}} \notag \\[4pt] &\qquad \equiv \sum_{m=0}^{{2(n-1)}/{d} } \begin{bmatrix} -\dfrac{1}{d} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{1}{d} \\ m \end{bmatrix}_{q^{2d}}q^{4m+2dm^{2}} \pmod {\Phi_n(q)^{2}}, \end{align*} $$

where we have performed the replacement $m=i+j$ and applied the q-Chu–Vandermonde identity (1.3).

When ${(n-1)}/{d} < m< n$ ,

$$ \begin{align*} \begin{bmatrix}-\dfrac{1}{d}\\[1mm] m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{1}{d} \\[1mm] m \end{bmatrix}_{q^{2d}} = \dfrac{(q^{2};q^{2d})^{2}_{m}}{(q^{2d};q^{2d})^{2}_{m}}q^{-4d\binom{m}{2}-4m} \equiv 0 \pmod {\Phi_n(q)^{2}}, \end{align*} $$

for $(q^{2};q^{2d})_{m} \equiv $ 0 (mod $\Phi _n(q)$ ), and $(q^{2d};q^{2d})_{m}$ is relatively prime to $\Phi _n(q)$ . Therefore, modulo $\Phi _n(q)^{2}$ , the left-hand side of (1.2) can be simplified as

$$ \begin{align*} \sum_{i=0}^{{(n-1)}/{d}}\sum_{j=0}^{{(n-1)}/{d}}\dfrac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj} \equiv \!\sum_{m=0}^{{(n-1)}/{d} } \begin{bmatrix} -\dfrac{1}{d} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{1}{d} \\ m \end{bmatrix}_{q^{2d}}q^{4m+2dm^{2}}\!. \end{align*} $$

lt is easy to check that

$$ \begin{align*} (1-q^{n-t})(1-q^{n+t})+(1-q^{t})^{2}q^{n-t} = (1-q^{n})^{2}, \end{align*} $$

from which we deduce

$$ \begin{align*} &\begin{bmatrix} \dfrac{n-1}{d} _{} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} \dfrac{n-(d-1)}{d} +m_{} \\ m \end{bmatrix}_{q^{2d}} =\dfrac{1}{(q^{2d};q^{2d})^{2}_{m}} \prod_{t=1}^{m}(1-q^{2n+(2+2td-2d)})(1-q^{2n-(2+2td-2d)}) \notag \\[4pt] &\quad=\dfrac{1}{(q^{2d};q^{2d})^{2}_{m}}\prod_{t=1}^{m}\big\{(1-q^{2n})^{2}-(1-q^{2n-(2+2td-2d)})^{2}q^{2n-(2+2td-2d)}\big\} \notag \\[4pt] &\quad \equiv (-1)^{m}\dfrac{(q^{2};q^{2d})^{2}_{m}}{(q^{2d};q^{2d})^{2}_{m}}q^{(2n-dm+d-2)m} \notag \\[4pt] &\quad\equiv (-1)^{m}\begin{bmatrix} -\dfrac{1}{d} _{} \\ m \end{bmatrix}_{q^{2d}}\begin{bmatrix} -\dfrac{1}{d} _{} \\ m \end{bmatrix}_{q^{2d}}q^{2nm+dm^{2}-dm+2m} \pmod{\Phi_n(q)^{2}}. \end{align*} $$

Thus, the left-hand of (1.2) becomes

$$ \begin{align*} &\sum_{i=0}^{{(n-1)}/{d}} \sum_{j=0}^{{(n-1)}/{d}}\dfrac{(q^{2};q^{2d})_{i+j}(q;q^{2d})_{i}(q;q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{i+2di+2dj}\\[5pt] &\qquad \equiv \sum_{m=0}^{{(n-1)}/{d} } (-1)^{m} \begin{bmatrix} \dfrac{n-1}{d} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} \dfrac{n-(d-1)}{d} +m \\ m \end{bmatrix}_{q^{2d}}q^{(2+d-2n)m+dm^{2}} \\[5pt] &\qquad\equiv \sum_{m=0}^{{(n-1)}/{d} } \begin{bmatrix} \dfrac{n-1}{d} \\[2mm] \dfrac{n-1}{d}-m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{n-(d-1)}{d}-1 \\ m \end{bmatrix}_{q^{2d}}q^{2dm^{2}+4m}\\[5pt] &\qquad\equiv \begin{bmatrix} -\dfrac{2}{d} \\[2mm] \dfrac{n-1}{d} \end{bmatrix} _{q^{2d}}q^{{(2n^{2}-2)}/{d}} \pmod{\Phi_n(q)^{2}}, \end{align*} $$

where we have used the q-Chu–Vandermonde identity (1.3) in the last line. This completes the proof of Theorem 1.2.

3. Proof of Theorem 1.3

The proof of Thereom 1.3 is very similar to the proof of Thereom 1.2. We give a sketch of its proof. The left-hand side is

(3.1) $$ \begin{align} &\sum_{i=0}^{{(n-(d-1))}/{d}} \sum_{j=0}^{{(n-(d-1))}/{d}}\dfrac{(q^{2d-2};q^{2d})_{i+j}(q^{d-1};q^{2d})_{i}(q^{d-1};q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{3di-i+2dj+2d-4} \notag \\[5pt] &\qquad\equiv \sum_{0 \le i+j \le {2(n-(d-1))}/{d}}^{} \dfrac{(q^{2d-2};q^{2d})_{i+j}(q^{d-1};q^{2d})_{i}(q^{d-1};q^{2d})_{j}}{(q^{2d};q^{2d})_{i+j}(q^{2d};q^{2d})_{i}(q^{2d};q^{2d})_{j}}q^{3di-i+2dj+2d-4} \notag\\[5pt] &\qquad\equiv \sum_{m=0}^{{2(n-(d-1))}/{d} } \begin{bmatrix} -\dfrac{d-1}{d} \\ m \end{bmatrix}_{q^ {2d}}\sum_{k=0}^{m} \begin{bmatrix} -\dfrac{d-1}{2d} \\ k \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{d-1}{2d} \\ m-k \end{bmatrix}_{q^{2d}}\notag\\[5pt]&\quad\qquad \times q^{(3d-3)m+(d-1)i+2dk^{2}-2dmk+2dm^{2}+2d-4}\notag\\[5pt] &\qquad \equiv \sum_{m=0}^{{2(n-(d-1))}/{d} } \begin{bmatrix} -\dfrac{d-1}{d} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{d-1}{d} \\ m \end{bmatrix}_{q^{2d}}q^{4dm-4m+2dm^{2}+2d-4} \notag\\[5pt] &\qquad \equiv \sum_{m=0}^{{(n-(d-1))}/{d} } \begin{bmatrix} -\dfrac{d-1}{d} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{d-1}{d} \\ m \end{bmatrix}_{q^{2d}}q^{4dm-4m+2dm^{2}+2d-4} \pmod{\Phi_n(q)^{2}}. \end{align} $$

To simplify this expression, note that

$$ \begin{align*} &\begin{bmatrix} -\dfrac{d-1}{d} _{} \\ m \end{bmatrix}_{q^{2d}}\begin{bmatrix} -\dfrac{d-1}{d} _{} \\ m \end{bmatrix}_{q^{2d}}\\ &\quad\equiv (-1)^{m} \begin{bmatrix} \dfrac{n-(d-1)}{d} _{} \\ m \end{bmatrix}_{q^{2d}}\begin{bmatrix} \dfrac{n-1}{d}+m _{} \\ m \end{bmatrix}_{q^{2d}} q^{-2nm-dm^{2}-dm+2m} \pmod{\Phi_n(q)^{2}}. \end{align*} $$

So the right-hand side of (3.1) becomes

$$ \begin{align*} &\sum_{m=0}^{{(n-(d-1))}/{d} } (-1)^{m} \begin{bmatrix} \dfrac{n-(d-1)}{d} \\ m \end{bmatrix}_{q^{2d}} \begin{bmatrix} \dfrac{n-1}{d} +m \\ m \end{bmatrix}_{q^{2d}}q^{(3d-2-2n)m+dm^{2}+2d-4}\\ &\qquad\equiv \sum_{m=0}^{{(n-(d-1))}{d} } \begin{bmatrix} \dfrac{n-(d-1)}{d} \\[2mm] \dfrac{n-(d-1)}{d}-m \end{bmatrix}_{q^{2d}} \begin{bmatrix} -\dfrac{n-1}{d}-1 \\[2mm] m \end{bmatrix}_{q^{2d}}q^{2dm^{2}+4(d-1)m}\\[1mm] &\qquad \equiv \begin{bmatrix} \dfrac{2}{d} -2 \\ \dfrac{n-(d-1)}{d} \end{bmatrix}_{q^{2d}}q^{{(2n^{2}-2)}{d}} \pmod{\Phi_n(q)^{2}}. \end{align*} $$

We can then complete the proof of Theorem 1.3 with the help of the q-Chu–Vandermonde identity (1.3).

Footnotes

This work is supported by National Natural Science Foundation of China (No. 12371331) and Natural Science Foundation of Shanghai (No. 22ZR1424100).

References

Gasper, G. and Rahman, M., Basic Hypergeometric Series, 2nd edn, Encyclopedia of Mathematics and Its Applications, 96 (Cambridge University Press, Cambridge, 2004).CrossRefGoogle Scholar
Guo, V. J. W., ‘A new $q$ -analogue of Van Hamme’s (A.2) supercongruence’, Bull. Aust. Math. Soc. 107(1) (2023), 2230.CrossRefGoogle Scholar
Guo, V. J. W. and Schlosser, M. J., ‘Some $q$ -supercongruences from transformation formulas for basic hypergeometric series’, Constr. Approx. 53(1) (2021), 155200.CrossRefGoogle Scholar
Guo, V. J. W. and Schlosser, M. J., ‘A family of $q$ -supercongruences modulo the cube of a cyclotomic polynomial’, Bull. Aust. Math. Soc. 105(2) (2022), 296302.CrossRefGoogle Scholar
Guo, V. J. W. and Zudilin, W., ‘A $q$ -microscope for supercongruences’, Adv. Math. 346 (2019), 329358.CrossRefGoogle Scholar
He, H. and Wang, X., ‘Some congruences that extend Van Hamme’s (D.2) supercongruence’, J. Math. Anal. Appl. 527(1) (2023), Article no. 127344.CrossRefGoogle Scholar
He, H. and Wang, X., ‘Two curious $q$ -supercongruences and their extensions’, Forum Math., to appear; doi:10.1515/forum-2023-0164.CrossRefGoogle Scholar
Lin, K.-Y. and Liu, J.-C., ‘Congruences for the truncated Appell series ${F}_3$ and ${F}_4$ ’, Integral Transforms Spec. Funct. 31(1) (2020), 1017.CrossRefGoogle Scholar
Liu, J.-C., ‘Supercongruences for truncated Appell series’, Colloq. Math. 158(2) (2019), 255263.CrossRefGoogle Scholar
Liu, J.-C. and Petrov, F., ‘Congruences on sums of $q$ -binomial coefficents’, Adv. Appl. Math. 116(2020), Article no. 102003.CrossRefGoogle Scholar
Liu, Y. and Wang, X., ‘Further $q$ -analogues of the (G.2) supercongruence of Van Hamme’, Ramanujan J. 59(3) (2020), 791802.CrossRefGoogle Scholar
Slater, L. J., Generalized Hypergeometric Functions (Cambridge University Press, Cambridge, 1996).Google Scholar
Wang, X. and Xu, C., ‘ $q$ -Supercongruences on triple and quadruple sums’, Results Math. 78(1) (2023), Article no. 27.CrossRefGoogle Scholar
Wang, X. and Yu, M., ‘A generalisation of a supercongruence on the truncated Appell series ${F}_3$ ’, Bull. Aust. Math. Soc. 107(2) (2023), 296303.CrossRefGoogle Scholar
Wang, X. and Yu, M., ‘Some new $q$ -supercongruences involving one free parameter’, Rocky Mountain J. Math. 53(4) (2023), 12851290.CrossRefGoogle Scholar