1 Introduction

In 1997, Van Hamme [24] presented 13 remarkable supercongruences corresponding to Ramanujan’s or to Ramanujan-like formulas for \(1/\pi \). For instance, the two infinite series expansions

$$\begin{aligned}&\sum _{k=0}^\infty (-1)^k(4k+1)\frac{(\frac{1}{2})_k^3}{k!^3} =\frac{2}{\pi }, \\&\sum _{k=0}^\infty (-1)^k(6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} =\frac{3\sqrt{3}}{2\pi }, \end{aligned}$$

correspond to the following two supercongruences for truncated hypergeometric series:

$$\begin{aligned} \sum _{k=0}^{(p-1)/2} (-1)^k(4k+1)\frac{(\frac{1}{2})_k^3}{k!^3}&\equiv (-1)^{(p-1)/2}p\pmod {p^3}, \end{aligned}$$
(1.1)
$$\begin{aligned} \sum _{k=0}^{(p-1)/3} (-1)^k(6k+1)\frac{(\frac{1}{3})_k^3}{k!^3}&\equiv p\pmod {p^3},\quad \text {for}\quad p\equiv 1\pmod {3}, \end{aligned}$$
(1.2)

where p is an odd prime, and \((a)_n=a(a+1)\cdots (a+n-1)\) denotes the Pochhammer symbol. The supercongruence (1.1) was first proved by Mortenson [18] using a technical evaluation of gamma functions, and later reproved by Zudilin [28] and Long [16]. Swisher [23] employed Long’s method to prove four supercongruences of Van Hamme, including (1.2) (i.e., the (E.2) supercongruence in [24]). He [11] also gave a generalization of (1.2). In 2016, Osburn and Zudilin [21] confirmed the last supercongruence conjecture of Van Hamme.

During the past few years, q-analogues of supercongruences have been investigated by many authors (see, for example, [3,4,5,6,7,8,9,10, 12,13,14,15, 19, 20, 22, 25,26,27, 29]). In particular, the first author [3, 4] gave q-analogues of (1.1) and (1.2) as follows: for any odd integer n,

$$\begin{aligned} \sum _{k=0}^{(n-1)/2}(-1)^k [4k+1]\frac{(q;q^2)_k^3}{(q^2;q^2)_k^3}q^{k^2} \equiv (-1)^{(n-1)/2}q^{(n-1)^2/4} [n]\pmod {[n]\Phi _n(q)^2}, \end{aligned}$$

and for any positive integer n with \(n\equiv 1\pmod {3}\),

$$\begin{aligned} \sum _{k=0}^{(n-1)/3}(-1)^k[6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3}q^{(3k^2+k)/2} \equiv&(-1)^{(n-1)/3}q^{(n-1)(n-2)/6}[n] \\&\pmod {[n]\Phi _n(q)^2}. \end{aligned}$$

Here and in what follows, \((a;q)_n=(1-a)(1-aq)\cdots (1-aq^{n-1})\) is the q-shifted factorial, \([n]=[n]_q=(1-q^n)/(1-q)\) is the q-integer, and \(\Phi _n(q)\) denotes the n-th cyclotomic polynomial in q, i.e.,

$$\begin{aligned} \Phi _n(q)=\prod _{\begin{array}{c} 1\leqslant k\leqslant n\\ \gcd (k,n)=1 \end{array}}(q-\zeta ^k), \end{aligned}$$

where \(\zeta \) is an n-th primitive root of unity. The first author and Zudilin [10, Theorem 3.5 with \(r=1\)] also gave another q-analogue of (1.2): for any positive integer n with \(n\equiv 1\pmod {6}\),

$$\begin{aligned} \sum _{k=0}^{(n-1)/3}(-1)^k[6k+1]_{q^2}\frac{(q^2;q^6)_k^3 (-q^3;q^6)_k}{(q^6;q^6)_k^3 (-q^5;q^6)_k}q^k \equiv q^{1-n} [n]_{q^2} \pmod {[n]\Phi _n(q)^2}. \end{aligned}$$
(1.3)

One of the aims of this paper is to establish the following new q-analogue of (1.2).

Theorem 1.1

Let \(n\equiv 1\pmod {6}\) be a positive integer. Then

$$\begin{aligned} \sum _{k=0}^{M} (-1)^k[6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3} \equiv q^{2(1-n)/3}[n]\frac{(-q^3;q^3)_{(n-1)/3}}{(-q^2;q^3)_{(n-1)/3}} \pmod {[n]\Phi _n(q)^2}, \end{aligned}$$
(1.4)

where \(M=(n-1)/3\) or \(M=n-1\).

We shall also give the following similar result.

Theorem 1.2

Let \(n\equiv 1\pmod {3}\) be a positive integer. Then

$$\begin{aligned} \sum _{k=0}^{M} [6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3} \equiv q^{2(1-n)/3}[n]\frac{(q^3;q^3)_{(n-1)/3}}{(q^2;q^3)_{(n-1)/3}} \pmod {\Phi _n(q)^3}, \end{aligned}$$
(1.5)

where \(M=(n-1)/3\) or \(M=n-1\).

Note that the supercongruence (1.5) does not hold modulo \([n]\Phi _n(q)^2\) in general, even for \(n\equiv 1\pmod {6}\). We take this opportunity to point out that Theorems 1 and 2 in [8] only hold modulo \(\Phi _n(q)^3\) and \(\Phi _n(q)^2\), respectively, but do not hold modulo [n], since Lemma 3 in [8] is not true (it only holds for even integers d).

Swisher [23] also proved that, for any prime \(p\equiv 2\pmod {3}\),

$$\begin{aligned} \sum _{k=0}^{(2p-1)/3} (-1)^k(6k+1)\frac{(\frac{1}{3})_k^3}{k!^3}&\equiv -2p\pmod {p^3}. \end{aligned}$$
(1.6)

A q-analogue of (1.6) was given by the first author [4, Theorem 1.5 with \((d,r)=(3,1)\)]: for any positive integer \(n\equiv 2\pmod {3}\),

$$\begin{aligned}&\sum _{k=0}^{(2n-1)/3}(-1)^kq^{(3k^2+k)/2}[6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3} \equiv -[2n]q^{(n-1)(2n-1)/3} \\&\pmod {[n]\Phi _n(q)^2}. \end{aligned}$$

In this paper, we shall give a new q-analogue of (1.6).

Theorem 1.3

Let \(n\equiv 5\pmod {6}\) be a positive integer. Then

$$\begin{aligned} \sum _{k=0}^{M} (-1)^k[6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3} \equiv -q^{2(1-2n)/3}[2n]\frac{(-q^3;q^3)_{(2n-1)/3}}{(-q^2;q^3)_{(2n-1)/3}} \pmod {[n]\Phi _n(q)^2}, \end{aligned}$$
(1.7)

where \(M=(2n-1)/3\) or \(M=n-1\).

Similarly, we have the following result.

Theorem 1.4

Let \(n\equiv 2\pmod {3}\) be a positive integer. Then

$$\begin{aligned} \sum _{k=0}^{M} [6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3} \equiv q^{2(1-2n)/3}[2n]\frac{(q^3;q^3)_{(2n-1)/3}}{(q^2;q^3)_{(2n-1)/3}} \pmod {\Phi _n(q)^2}, \end{aligned}$$
(1.8)

where \(M=(2n-1)/3\) or \(M=n-1\).

Note that the q-supercongruence (1.8) does not hold modulo \(\Phi _n(q)^3\) for \(n>2\). We shall prove Theorems 1.1, 1.2, and 1.3 modulo \(\Phi _n(q)^3\) and Theorem 1.4 by using a summation for a very-well-poised \({}_6\phi _5\) series and the ‘creative microscoping’ method introduced by the first author in collaboration with Zudilin [9]. The proof of Theorems 1.1 and 1.3 also requires the use of a lemma previously given by the present authors.

From Theorems 1.2 and 1.4, we can deduce the following supercongruences.

Corollary 1.5

Let \(p\equiv 1\pmod {3}\) be a prime. Then

$$\begin{aligned} \sum _{k=0}^{(p-1)/3} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv p\Gamma _p(\tfrac{2}{3})^3\pmod {p^3}, \end{aligned}$$
(1.9)

where \(\Gamma _p(x)\) denotes the p-adic Gamma function.

Corollary 1.6

Let \(p\equiv 2\pmod {3}\) be an odd prime. Then

$$\begin{aligned} \sum _{k=0}^{(2p-1)/3} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv -6\Gamma _p(\tfrac{2}{3})^3\pmod {p^2}. \end{aligned}$$

2 Proof of Theorem 1.1

We first give the following result, which is due to the present authors [6, Lemma 2.1].

Lemma 2.1

Let d, m and n be positive integers with \(m\leqslant n-1\). Let r be an integer satisfying \(dm\equiv -r\pmod {n}\). Then, for \(0\leqslant k\leqslant m\) and any indeterminate a, we have

$$\begin{aligned} \frac{(aq^r;q^d)_{m-k}}{(q^d/a;q^d)_{m-k}} \equiv (-a)^{m-2k}\frac{(aq^r;q^d)_k}{(q^d/a;q^d)_k} q^{m(dm-d+2r)/2+(d-r)k} \pmod {\Phi _n(q)}. \end{aligned}$$

If \(\gcd (d,n)=1\), then the above q-congruence also holds for \(a=1\).

We also need the following result to prove the truth of (1.4) modulo [n].

Lemma 2.2

Let n be a positive integer coprime with 6, and let a be an indeterminate. Then

$$\begin{aligned}&\sum _{k=0}^{m}(-1)^k[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k } \equiv 0\pmod {[n]}, \end{aligned}$$
(2.1)
$$\begin{aligned}&\sum _{k=0}^{n-1}(-1)^k[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k } \equiv 0\pmod {[n]}, \end{aligned}$$
(2.2)

where \(m=(n-1)/3\) if \(n\equiv 1\pmod {6}\), and \(m=(2n-1)/3\) if \(n\equiv 5\pmod {6}\).

Proof

Clearly, Lemma 2.2 is true for \(n=1\). We now assume that \(n>1\). By Lemma 2.1, we can easily deduce that the k-th and \((m-k)\)-th terms on the left-hand side of (2.1) cancel each other modulo \(\Phi _n(q)\), i.e.,

$$\begin{aligned}&(-1)^{m-k}\frac{[6(m-k)+1](aq;q^3)_{m-k}(q/a;q^3)_{m-k}(q;q^3)_{m-k} }{(aq^3;q^3)_{m-k} (q^3/a;q^3)_{m-k} (q^3;q^3)_{m-k} } \\&\qquad \equiv -(-1)^k[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k }\pmod {\Phi _n(q)}. \end{aligned}$$

Thus, we have proved that the q-congruence (2.1) holds modulo \(\Phi _n(q)\). Since the numerator contains the factor \((q;q^3)_k\), it is easy to see that the k-th summand in (2.2) is congruent to 0 modulo \(\Phi _n(q)\) for \(m< k\leqslant n-1\). This proves the q-congruence (2.2) modulo \(\Phi _n(q)\).

Now we can prove (2.1) and (2.2) modulo [n]. Let \(\zeta \ne 1\) be an n-th root of unity, not necessarily primitive. In other words, \(\zeta \) is a primitive root of unity of degree s satisfying \(s\mid n\) and \(s>1\). Let \(c_q(k)\) stand for the k-th term on the left-hand side of (2.2), i.e.,

$$\begin{aligned} c_q(k) =(-1)^k[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k }. \end{aligned}$$

Taking \(n=s\) in the q-congruences (2.1) and (2.2) modulo \(\Phi _n(q)\), we get

$$\begin{aligned} \sum _{k=0}^{s_1}c_\zeta (k)=\sum _{k=0}^{s-1}c_\zeta (k)=0, \end{aligned}$$

where \(s_1=(s-1)/3\) if \(s\equiv 1\pmod {6}\), and \(s_1=(2s-1)/3\) if \(s\equiv 5\pmod {6}\). It is not difficult to see that

$$\begin{aligned} \lim _{q\rightarrow \zeta }\frac{c_q(\ell s+k)}{c_q(\ell s)} =\frac{c_\zeta (\ell s+k)}{c_\zeta (\ell s)} =c_\zeta (k). \end{aligned}$$

Therefore,

$$\begin{aligned} \sum _{k=0}^{n-1}c_\zeta (k)=\sum _{\ell =0}^{n/s-1} \sum _{k=0}^{s-1}c_\zeta (\ell s+k) =\sum _{\ell =0}^{n/s-1}c_\zeta (\ell s) \sum _{k=0}^{s-1}c_\zeta (k)=0, \end{aligned}$$
(2.3)

and

$$\begin{aligned} \sum _{k=0}^{m}c_\zeta (k) =\sum _{\ell =0}^{(m-s_1)/s-1} c_\zeta (\ell s)\sum _{k=0}^{s-1}c_\zeta (k) +c_\zeta (m-s_1)\sum _{k=0}^{s_1}c_\zeta (k)=0. \end{aligned}$$

This proves that both of the sums \(\sum _{k=0}^{n-1}c_q(k)\) and \(\sum _{k=0}^{m}c_q(k)\) are divisible by \(\Phi _s(q)\) for any divisor \(s>1\) of n. Since

$$\begin{aligned} \prod _{s\mid n,\, s>1}\Phi _s(q)=[n], \end{aligned}$$

we complete the proof of (2.1) and (2.2). \(\square \)

Like most of the q-supercongruences in [9], we have the following parametric generalization of Theorem 1.1.

Theorem 2.3

Let \(n\equiv 1\pmod {6}\) be a positive integer. Then, modulo \([n](1-aq^n)(a-q^n)\),

$$\begin{aligned} \sum _{k=0}^{M} (-1)^k[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k } \equiv q^{2(n-1)/3}[n]\frac{(-q^3;q^3)_{(n-1)/3}}{(-q^2;q^3)_{(n-1)/3}}, \end{aligned}$$
(2.4)

where \(M=(n-1)/3\) or \(M=n-1\).

Proof

We start with the following summation for a very-well-poised \({}_6\phi _5\) series (see [2, Appendix (II.20)]):

$$\begin{aligned}&\sum _{k=0}^\infty \frac{(1-aq^{2k})(a;q)_k(b;q)_k(c;q)_k(d;q)_k}{(1-a)(q;q)_k(aq/b;q)_k(aq/c;q)_k(aq/d;q)_k}\biggl (\frac{aq}{bcd}\biggr )^k \nonumber \\&\quad =\frac{(aq;q)_\infty (aq/bc;q)_\infty (aq/bd;q)_\infty (aq/cd;q)_\infty }{(aq/b;q)_\infty (aq/c;q)_\infty (aq/d;q)_\infty (aq/bcd;q)_\infty }. \end{aligned}$$
(2.5)

(The infinite series in (2.5) converges for \(|q|<1\) and \(|aq/bcd|<1\).) Specializing (2.5) by letting \(q\mapsto q^3\), \(a=q\), \(b=q^{1-n}\), \(c=q^{1+n}\), and \(d=-q^2\), we have

$$\begin{aligned}&\sum _{k=0}^{(n-1)/3} (-1)^k[6k+1]\frac{(q^{1-n};q^3)_k(q^{1+n};q^3)_k(q;q^3)_k }{(q^{3-n};q^3)_k (q^{3+n};q^3)_k (q^3;q^3)_k } \\&\quad \quad =\frac{(q^4;q^3)_{(n-1)/3} (-q^{1-n};q^3)_{(n-1)/3}}{(q^{3-n};q^3)_{(n-1)/3}(-q^2;q^3)_{(n-1)/3}} \\&\quad \quad =q^{2(1-n)/3}[n]\frac{(-q^3;q^3)_{(n-1)/3}}{(-q^2;q^3)_{(n-1)/3}}. \end{aligned}$$

This shows that both sides of (2.4) are equal for \(a=q^{-n}\) and \(a=q^n\). This means that the congruence (2.4) holds modulo \(1-aq^n\) and \(a-q^n\).

Moreover, by Lemma 2.2, the left-hand side of (2.4) is congruent to 0 modulo [n]. Since \(1-q^n\) (n is odd) is relatively prime to \(1+q^k\), we see that the right-hand side of (2.4) is also congruent to 0 modulo [n]. Noticing that \(1-aq^n\), \(a-q^n\), and [n] are pairwise coprime polynomials in q, we finish the proof of the theorem. \(\square \)

Proof of Theorem 1.1

Since \((1-q^n)^2\) contains the factor \(\Phi _n(q)^2\) and \((q^3;q^3)_M\) is coprime with \(\Phi _n(q)\), letting \(a=1\) in (2.4), we conclude that (1.4) is true modulo \(\Phi _n(q)^3\). Note that Lemma 2.2 also holds for \(a=1\). Namely, the q-congruence (1.4) is true modulo [n] and is therefore also true modulo \([n]\Phi _n(q)^2\). This completes the proof. \(\square \)

3 Proof of Theorem 1.2

We first give the following parametric generalization of Theorem 1.2: for \(n\equiv 1\pmod {3}\), modulo \(\Phi _n(q)(1-aq^n)(a-q^n)\),

$$\begin{aligned} \sum _{k=0}^{M} [6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k } \equiv q^{2(n-1)/3}[n]\frac{(q^3;q^3)_{(n-1)/3}}{(q^2;q^3)_{(n-1)/3}},\nonumber \\ \end{aligned}$$
(3.1)

where \(M=(n-1)/3\) or \(M=n-1\). The proof of (3.1) is analogous to that of (2.4). This time, we make the substitutions \(q\mapsto q^3\), \(a=q\), \(b=q^{1-n}\), \(c=q^{1+n}\), and \(d=q^2\) in (2.5) to obtain

$$\begin{aligned}&\sum _{k=0}^{(n-1)/3} [6k+1]\frac{(q^{1-n};q^3)_k(q^{1+n};q^3)_k(q;q^3)_k }{(q^{3-n};q^3)_k (q^{3+n};q^3)_k (q^3;q^3)_k } \\&\quad \quad =\frac{(q^4;q^3)_{(n-1)/3} (q^{1-n};q^3)_{(n-1)/3}}{(q^{3-n};q^3)_{(n-1)/3}(q^2;q^3)_{(n-1)/3}} \\&\quad \quad =q^{2(1-n)/3}[n]\frac{(q^3;q^3)_{(n-1)/3}}{(q^2;q^3)_{(n-1)/3}}. \end{aligned}$$

Thus, the two sides of (3.1) are equal for \(a=q^{-n}\) and \(a=q^n\). This means that the congruence (3.1) is true modulo \(1-aq^n\) and \(a-q^n\).

Moreover, by Lemma 2.1, for \(m=(n-1)/3\) we can deduce that the k-th and \((m-k)\)-th terms on the left-hand side of (3.1) cancel each other modulo \(\Phi _n(q)\), i.e.,

$$\begin{aligned}&\frac{[6(m-k)+1](aq;q^3)_{m-k}(q/a;q^3)_{m-k}(q;q^3)_{m-k} }{(aq^3;q^3)_{m-k} (q^3/a;q^3)_{m-k} (q^3;q^3)_{m-k} } \\&\qquad \equiv -[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k }\pmod {\Phi _n(q)}. \end{aligned}$$

(Note that we have utilized the fact that \(q^{n/2}\equiv -1\pmod {\Phi _n(q)}\) for even n.) This proves (3.1) modulo \(\Phi _n(q)\).

Finally, letting \(a=1\) in (3.1), we arrive at the q-supercongruence (1.5).

4 Proof of Theorems 1.3 and 1.4

Proof of Theorem 1.3

We first give a parametric generalization of Theorem 1.4: for \(n\equiv 5\pmod {6}\), modulo \([n](1-aq^{2n})(a-q^{2n})\),

$$\begin{aligned}&\sum _{k=0}^{M} (-1)^k[6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k } \equiv -q^{2(1-2n)/3}[2n]\nonumber \\&\quad \frac{(-q^3;q^3)_{(2n-1)/3}}{(-q^2;q^3)_{(2n-1)/3}}, \end{aligned}$$
(4.1)

where \(M=(2n-1)/3\) or \(n-1\). The proof of (4.1) is very similar to that of (2.4). Specializing (2.5) by \(q\mapsto q^3\), \(a=q\), \(b=q^{1-2n}\), \(c=q^{1+2n}\), and \(d=-q^2\), we have

$$\begin{aligned}&\sum _{k=0}^{(2n-1)/3} (-1)^k[6k+1]\frac{(q^{1-2n};q^3)_k(q^{1+2n};q^3)_k(q;q^3)_k }{(q^{3-2n};q^3)_k (q^{3+2n};q^3)_k (q^3;q^3)_k } \\&\quad \quad =\frac{(q^4;q^3)_{(2n-1)/3} (-q^{1-2n};q^3)_{(2n-1)/3}}{(q^{3-2n};q^3)_{(2n-1)/3}(-q^2;q^3)_{(2n-1)/3}} \\&\quad \quad =-q^{2(1-2n)/3}[2n]\frac{(-q^3;q^3)_{(2n-1)/3}}{(-q^2;q^3)_{(2n-1)/3}}. \end{aligned}$$

This proves the congruence (4.1) modulo \(1-aq^{2n}\) and \(a-q^{2n}\). Moreover, the proof of (4.1) modulo [n] follows from Lemma 2.2. \(\square \)

Finally, taking \(a=1\) in (4.1), we arrive at the desired q-supercongruence (1.7).

Proof of Theorem 1.4

We have the following congruence with a parameter a: for \(n\equiv 5\pmod {6}\), modulo \((1-aq^{2n})(a-q^{2n})\),

$$\begin{aligned} \sum _{k=0}^{M} [6k+1]\frac{(aq;q^3)_k(q/a;q^3)_k(q;q^3)_k }{(aq^3;q^3)_k (q^3/a;q^3)_k (q^3;q^3)_k } \equiv q^{2(1-2n)/3}[2n]\frac{(q^3;q^3)_{(2n-1)/3}}{(q^2;q^3)_{(2n-1)/3}}, \end{aligned}$$
(4.2)

where \(M=(2n-1)/3\) or \(M=n-1\). The congruence (4.2) is equivalent to say that both sides are equal for \(a=q^{2n}\) and \(a=q^{-2n}\). But this again follows from (2.5) by performing the parameter substitutions \(q\mapsto q^3\), \(a=q\), \(b=q^{1-2n}\), \(c=q^{1+2n}\), and \(d=q^2\). At last, letting \(a=1\) in (4.2), we get (1.8). \(\square \)

5 Proof of Corollaries 1.5 and 1.6

Proof of Corollary 1.5

Letting \(n=p\), where p is a prime congruent to \(1\pmod {3}\), and \(q\rightarrow 1\) in (1.5), we obtain

$$\begin{aligned} \sum _{k=0}^{(p-1)/3} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv p\frac{(\frac{p-1}{3})!}{(\frac{2}{3})_{(p-1)/3}}\pmod {p^3}. \end{aligned}$$

Recall that the p-adic Gamma function has the properties: for any p-adic integer x,

$$\begin{aligned}{} & {} \frac{\Gamma _p(x+1)}{\Gamma _p(x)} ={\left\{ \begin{array}{ll}\displaystyle -x,\quad p\not \mid x,\\ \displaystyle -1,\quad p\mid x, \end{array}\right. }\\{} & {} \quad \Gamma _p(x)\Gamma _p(1-x)=(-1)^{a_0(x)}, \end{aligned}$$

where \(a_0(x)\in \{1,2,\ldots ,p\}\) satisfies \(a_0(x)\equiv x\pmod {p}\). Let \(\Gamma (x)\) be the classical Gamma function. Then

$$\begin{aligned} \frac{(\frac{p-1}{3})!}{(\frac{2}{3})_{(p-1)/3}}&=\frac{\Gamma (\frac{p+2}{3})\Gamma (\frac{2}{3})}{\Gamma (1)\Gamma (\frac{p+1}{3})} =\frac{\Gamma _p(\frac{p+2}{3})\Gamma _p(\frac{2}{3})}{\Gamma _p(1)\Gamma _p(\frac{p+1}{3})}\\&=(-1)^{(2p+1)/3}\frac{\Gamma _p(\frac{p+2}{3})\Gamma _p(\frac{2-p}{3})\Gamma _p(\frac{2}{3})}{\Gamma _p(1)}. \end{aligned}$$

By [17, Theorem 14]), for \(p\geqslant 5\), we have

$$\begin{aligned} \Gamma _p(a+mp)\equiv \Gamma _p(a)+\Gamma _p'(a)mp\pmod {p^2}, \end{aligned}$$
(5.1)

and so \(\Gamma _p(\frac{p+2}{3})\Gamma _p(\frac{2-p}{3})\equiv \Gamma _p(\frac{2}{3})^2 \pmod {p^2}\). The proof then follows from the fact \(\Gamma _p(1)=(-1)^{(2p+1)/3}=-1\). \(\square \)

Proof of Corollary 1.6

Letting \(n=p\), where p is an odd prime congruent to \(2\pmod {3}\), and \(q\rightarrow 1\) in (1.8), we obtain

$$\begin{aligned} \sum _{k=0}^{(2p-1)/3} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv 2p\frac{(\frac{2p-1}{3})!}{(\frac{2}{3})_{(2p-1)/3}}\pmod {p^2}. \end{aligned}$$

Further,

$$\begin{aligned} \frac{p(\frac{2p-1}{3})!}{(\frac{2}{3})_{(2p-1)/3}} =p\frac{\Gamma (\frac{2p+2}{3})\Gamma (\frac{2}{3})}{\Gamma (1)\Gamma (\frac{2p+1}{3})} =3\frac{\Gamma _p(\frac{2p+2}{3})\Gamma _p(\frac{2}{3})}{\Gamma _p(1)\Gamma _p(\frac{2p+1}{3})} =3\frac{\Gamma _p(\frac{2p+2}{3})\Gamma _p(\frac{2-2p}{3})\Gamma _p(\frac{2}{3})}{\Gamma _p(1)}, \end{aligned}$$

and by (5.1), \(\Gamma _p(\frac{2p+2}{3})\Gamma _p(\frac{2-2p}{3})\equiv \Gamma _p(\frac{2}{3})^2 \pmod {p^2}\). \(\square \)

6 Some Open Problems

Although the q-supercongruence (1.5) is not true modulo [n] in general, using the same arguments as in the proof of Theorem 1.1, we can show that, for \(n\equiv 1\pmod {3}\) and \(n>1\),

$$\begin{aligned} \sum _{k=0}^{M} [6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3} \equiv 0 \pmod {\prod _{\begin{array}{c} j|n,\ j>1\\ j\equiv 1 \bmod 3 \end{array}}\Phi _{j}(q)}, \end{aligned}$$
(6.1)

where \(M=(n-1)/3\) or \(M=n-1\). Letting \(n=p^r\) and \(q\rightarrow 1\) in the above q-congruence, we obtain the following result: for any prime \(p\equiv 1\pmod {3}\) and integer \(r\geqslant 1\),

$$\begin{aligned} \sum _{k=0}^{(p^r-1)/d} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv 0 \pmod {p^r}, \end{aligned}$$
(6.2)

where \(d=1,3\). Inspired by Dwork’s work [1] and Swisher’s conjectures [23, (A.3)–(L.3)], we propose the following conjecture on Dwork-type supercongruences, which is a uniform generalization of (1.9) and (6.2).

Conjecture 6.1

Let \(p\equiv 1\pmod {3}\) be a prime and let \(r\geqslant 1\). Then

$$\begin{aligned} \sum _{k=0}^{(p^r-1)/d} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv p\Gamma _p(\tfrac{2}{3})^3 \sum _{k=0}^{(p^{r-1}-1)/d} (6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \pmod {p^{3r}}, \end{aligned}$$

where \(d=1,3\).

Note that the following Dwork-type supercongruence (see [23, (E.3)] and [4, Conjecture 5.3]) has been proved by the first author and Zudilin [9, Theorem 3.5] by establishing its q-analogue:

For any prime \(p\equiv 1\pmod {3}\) and integer \(r\geqslant 1\),

$$\begin{aligned} \sum _{k=0}^{(p^r-1)/d} (-1)^k(6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \equiv p \sum _{k=0}^{(p^{r-1}-1)/d} (-1)^k(6k+1)\frac{(\frac{1}{3})_k^3}{k!^3} \pmod {p^{3r}}, \end{aligned}$$
(6.3)

where \(d=1,3\).

We believe that the following new q-analogue of (6.3), which is also a generalization of Theorem 1.1, should be true.

Conjecture 6.2

Let \(n>1\) be an integer with \(n\equiv 1\pmod {6}\) and let \(r\geqslant 1\). Then, modulo \([n^r]\prod _{j=1}^r\Phi _{n^j}(q)^2\),

$$\begin{aligned} \sum _{k=0}^{(n^r-1)/d}(-1)^k[6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3}&\equiv q^{2(1-n)/3}[n]\dfrac{(-q^3;q^3)_{(n^r-1)/3}(-q^{2n};q^{3n})_{(n^{r-1}-1)/3}}{(-q^2;q^3)_{(n^r-1)/3} (-q^{3n};q^{3n})_{(n^{r-1}-1)/3}}\nonumber \\&\quad \times \sum _{k=0}^{(n^{r-1}-1)/d}(-1)^k[6k+1]_{q^n}\frac{(q^n;q^{3n})_k^3}{(q^{3n};q^{3n})_k^3}, \end{aligned}$$

where \(d=1,3\).

Likewise, we conjecture a Dwork-type generalization of Theorem 1.2 as follows.

Conjecture 6.3

Let \(n>1\) be an integer with \(n\equiv 1\pmod {3}\) and let \(r\geqslant 1\). Then, modulo \(\prod _{j=1}^r\Phi _{n^j}(q)^3\),

$$\begin{aligned} \sum _{k=0}^{(n^r-1)/d}[6k+1]\frac{(q;q^3)_k^3}{(q^3;q^3)_k^3}&\equiv q^{2(1-n)/3}[n]\dfrac{(q^3;q^3)_{(n^r-1)/3}(q^{2n};q^{3n})_{(n^{r-1}-1)/3}}{(q^2;q^3)_{(n^r-1)/3} (q^{3n};q^{3n})_{(n^{r-1}-1)/3}}\nonumber \\&\quad \times \sum _{k=0}^{(n^{r-1}-1)/d}[6k+1]_{q^n}\frac{(q^n;q^{3n})_k^3}{(q^{3n};q^{3n})_k^3}, \end{aligned}$$

where \(d=1,3\).