Some q-supercongruences for truncated basic hypergeometric series

Victor J. W. Guo School of Mathematical Sciences Huaiyin Normal University Huaian, Jiangsu 223300 People’s Republic of China E-mail: [email protected] Jiang Universit´ede Lyon Universit´eLyon 1 Institut Camille Jordan, UMR 5208 CNRS 43, boulevard du 11 novembre 1918 F-69622 Villeurbanne Cedex, France E-mail: [email protected]

Abstract For any odd prime p we obtain q-analogues of van Hamme’s and Rodriguez-Villegas’ supercongruences involving products of three bi- nomial coefficients such as

p−1 2  3 2k X 2k q 2 2 2 2 2 ≡ 0 (mod [p] ) for p ≡ 3 (mod 4), k 2 (−q ; q ) (−q; q) k=0 q k 2k p−1 2   3 2 3 3k X 2k (q; q )k(q ; q )kq 2 6 6 2 ≡ 0 (mod [p] ) for p ≡ 2 (mod 3), k 3 (q ; q ) k=0 q k

p−1 n−1 where [p] = 1+q+···+q and (a; q)n = (1−a)(1−aq) ··· (1−aq ). We also prove q-analogues of the brothers’ generalizations of the

2010 Mathematics Subject Classification: Primary 11B65; Secondary 05A10, 05A30. Key words and phrases: p-adic integer, congruences, supercongruences, basic hyper- geometric series, q-binomial theorem, q-Chu-Vandermonde formula.

1 2 V. J. W. Guo and J. Zeng

above supercongruences. Our proofs are elementary in nature and use the theory of basic hypergeometric series and combinatorial q- binomial identities including a new q-Clausen-type summation for- mula.

1 Introduction

We shall follow the standard q-notations from [4]. The q-shifted factorial n−1 is defined by (a; q)n = (1 − a)(1 − aq) ··· (1 − aq ) for n = 1, 2,..., and 1−qn (a; q)0 = 1, while the q-integer is denoted by [n] := 1−q . In a previous pa- per [6], we proposed several q-analogues of Rodriguez-Villegas and Morten- son type congruences for truncated hypergeometric series conjectured by Rodriguez-Villegas [11, 9] and proved the following q-analogue of one of their supercongruences:

p−1 2 2   2 X (q; q )k −1 1−p 2 (1.1) ≡ q 4 (mod [p] ). (q2; q2)2 p k=0 k · Here and in what follows p denotes an odd prime, and ( p ) is the Legendre symbol modulo p. Recently, by using the properties of generalized Legendre polynomials, Z.-H. Sun [13, Theorem 2.5] proved the following remarkable congruence:

p−1 X 2ka−1 − a 1 (1.2) ≡ 0 (mod p2), k k k 4k k=0 where a is a p-adic integer such that the least nonnegative residue of a mod- ulo p is odd. It is interesting to note that (1.2) is a common generalization of several congruences due to van Hamme and Rodriguez-Villegas. On the other hand, van Hamme [17] proved the following congruence (1.3) p−1 2 3 ( X 2k 1 4x2 − 2p (mod p2), if p = x2 + y2 with x odd, ≡ k 64k 0 (mod p2), if p ≡ 3 (mod 4), k=0 of which a generalization was recently conjectured by H. Swisher [16, (H.3)]. The aim of this paper is to prove some q-supercongruences for certain trun- cated basic hypergeometric series generalizing the above results. n Recall that the q-binomial coefficients k are then defined by  n−k+1 (q ; q)k n n  , if 0 6 k 6 n, = = (q; q)k k k q 0, otherwise. truncated basic hypergeometric series 3

The first aim of this paper is to give a unified q-analogue of (1.3) and Z.-W. Sun’s generalization [14, Theorem 1.1(i)] of (1.3).

p−1 Theorem 1.1. Let 0 6 s 6 2 . Then the following congruence holds modulo [p]2: (1.4)

p−1 2 2 X 2k  2k  q2k k k + s (−q2; q2)2(−q; q)2 k=0 q2 q2 k 2k  2 2 2 2 2  p−1  p−2s−1 p+2s−1 p−1 2 (q ; q ) (q ; q )  s −s 2 2 2 p−1 (−1) q 2 , if s ≡ mod 2, p−2s−1 (q4; q4)2 2 ≡ 4 q4 p−1  2 0, otherwise.

When s = 0, the congruence (1.4) reduces to the following result.

Corollary 1.2. There holds

p−1 2 3 X 2k q2k (1.5) k (−q2; q2)2(−q; q)2 k=0 q2 k 2k   p−1 2  p−1 2 1 q 2 , if p ≡ 1 (mod 4), p−1 (−q2; q2)2 2 ≡ 4 q4 p−1 (mod [p] ).  2 0, otherwise,

To see that (1.5) is a q-analogue of (1.3) we need to recall a known result. Let p be a prime such that p ≡ 1 (mod 4) and p = x2 + y2 with x ≡ 1 (mod 4), then we have the so-called Beukers-Chowla-Dwork-Evans congruence [3, 10]:

 p−1  2p−1 + 1  p  2 ≡ 2x − (mod p2), p−1 2 2x 4 which easily implies Sun’s congruence [12, Lemma 3.4])

2  p−1  1 (1.6) 2 ≡ 4x2 − 2p (mod p2). p−1 2p−1 4 Using (1.6), it is clear that (1.5) reduces to (1.3) when q → 1. Generalizing a result of Mortenson [7, 8], Z.-W. Sun [15, (1.4)] obtained the following congruence

p−1 2 2k2k+2s X −1 p − 1 (1.7) k k+s ≡ (mod p2), for 0 s . 42k+s p 6 6 2 k=0 4 V. J. W. Guo and J. Zeng

For any p-adic integer x, let hxip denote the least nonnegative residue of x modulo p. The second aim of this paper is to give a unified q-analogue of (1.2) and (1.7).

Theorem 1.3. Let m and r two positive integers with p - m. Let s 6 r m−r r min{h− m ip, h− m ip} be a nonnegative integer. If h− m ip ≡ s + 1 (mod 2), then

p−1 2 m m r m m−r m mk X (q ; q )2k(q ; q )k(q ; q )kq 2 (1.8) m m m m 2m 2m 2 ≡ 0 (mod [p] ), (q ; q )k−s(q ; q )k+s(q ; q ) k=s k p−1 m m r m m−r m mk X (q ; q )2k(q ; q )k(q ; q )kq 2 (1.9) m m m m 2m 2m 2 ≡ 0 (mod [p] ). (q ; q )k−s(q ; q )k+s(q ; q ) k=s k

r If h− m ip ≡ s (mod 2), then the following congruence holds:

p−1 2 m m r m m−r m mk X (q ; q )2k(q ; q )k(q ; q )kq (1.10) m m m m 2m 2m 2 (q ; q )k−s(q ; q )k+s(q ; q ) k=s k (s+ p−1 )m m 2m m 2m 2 r ≡ q (q ; q ) h− ip−s (q ; q ) m−r m h− m ip−s 2 2 −mh− r i m −mh− m−r i m (q m p ; q ) (q m p ; q ) × s s (mod [p]). 2m 2m r 2m 2m (q ; q ) h− ip+s (q ; q ) m−r m h− m ip+s 2 2

r Letting s = 0, − m = a and q → 1 in (1.9), we obtain (1.2). On the other hand, it is not difficult to see that (see [6]), for any prime p > 5,       h− 1 i −3 h− 1 i −2 h− 1 i −1 (−1) 3 p = , (−1) 4 p = , (−1) 6 p = . p p p

Taking r = 1 and m = 3, 4, 6 in (1.8), we obtain

Corollary 1.4. Let > 5 be a prime and let s be a nonnegative integer. Then the following congruences hold modulo [p]2:

p−1 2   3 2 3 3k −3 X 2k (q; q )k(q ; q )kq p − 1 1 + ( p ) ≡ 0, if s and s ≡ mod 2, k + s (q6; q6)2 6 3 2 k=s q3 k p−1 2   4 3 4 4k −2 X 2k (q; q )k(q ; q )kq p − 1 1 + ( p ) ≡ 0, if s and s ≡ mod 2, k + s (q8; q8)2 6 4 2 k=s q4 k p−1 2   6 5 6 6k −1 X 2k (q; q )k(q ; q )kq p − 1 1 + ( p ) ≡ 0, if s and s ≡ mod 2. k + s (q12; q12)2 6 6 2 k=s q6 k truncated basic hypergeometric series 5

The proof of Theorem 1.3 is based on the following q-Clausen type sum- mation formula, which seems new and interesting in its own right.

Theorem 1.5. Let n and s be nonnegative integers with s 6 n. Then

n −2n 2 k ! n −2n 2 k ! X (q ; q )k(x; q)kq X (q ; q )k(q/x; q)kq (1.11) (q; q)k−s(q; q)k+s (q; q)k−s(q; q)k+s k=s k=s n 2 2 2 −n2 n k 2 2 k2−2nk (−1) (q ; q )nq X (−1) (q ; q )n+k(x; q)k(q/x; q)kq = 2 2 2 2 2 2 . (q ; q )n−s(q ; q )n+s (q ; q )n−k(q; q)k−s(q; q)k+s(q; q)2k k=s

We also have the following q-analogue of (1.7), which reduces to (1.1) when s = 0.

p−1 Theorem 1.6. Let p be an odd prime and let 0 6 s 6 2 . Then

p−1 2 2 2   2 X (q; q )k(q; q )k+s −1 1−p 4 2 2 2 2 2 ≡ q (mod [p] ). (q ; q )k(q ; q )k+s p k=0

Finally, we shall prove the following result.

Theorem 1.7. Let p be an odd prime and let m, r be positive integers with m−r p - m and r < m. Then for any integer s with 0 6 s 6 h− m ip, there holds

(1.12)

p−s−1 r m m−r m r r X (q ; q )k(q ; q )k+s r −mh− m ip(h− m ip+1) h− m ip 2 m m m m ≡ (−1) q (mod [p]). (q ; q )k(q ; q )k+s k=0

In particular, if p ≡ ±1 (mod m), then

p−s−1 r m m−r m 2 X (q ; q )k(q ; q )k+s r r(m−r)(1−p ) h− m ip 2m (1.13) m m m m ≡ (−1) q (mod [p]). (q ; q )k(q ; q )k+s k=0

Throughout the paper we will often use the fact that for any prime p, the q-integer [p] is always an irreducible polynomial in Q[q]. Namely, Q[q]/[p] is a field. Therefore, rational functions a(q)/b(q) are well defined modulo [p] r or [p] (r > 1) on condition that b(q) is relatively prime to [p]. The rest of this paper is organized as follows. In Sections 2–5 we prove Theorem 1.1, Theorem 1.5, Theorem 1.3, and Theorems 1.6 and 1.7, respec- tively. We conclude the paper with some open problems. 6 V. J. W. Guo and J. Zeng

2 Proof of Theorem 1.1

We first establish two lemmas.

p−1 Lemma 2.1. Let 0 6 k 6 2 . Then

 p−1    kp−k2 2 + k k 2k q 2 (2.1) ≡ (−1) 2 (mod [p] ). 2k q2 k q2 (−q; q)2k Proof. Since

(1 − qp−2j+1)(1 − qp+2j−1) + (1 − q2j−1)2qp−2j+1 = (1 − qp)2, we have

(1 − qp−2j+1)(1 − qp+2j−1) ≡ −(1 − q2j−1)2qp−2j+1 (mod [p]2).

It follows that  p−1  Qk (1 − qp−2j+1)(1 − qp+2j−1) 2 + k j=1 = 2 2 2k q2 (q ; q )2k Qk (1 − q2j−1)2qp−2j+1 k j=1 ≡ (−1) 2 2 (q ; q )2k   kp−k2 k 2k q 2 = (−1) 2 (mod [p] ), k q2 (−q; q)2k as desired. 

Lemma 2.2. Let n and s be nonnegative integers with s 6 n. Then

n     k (n−k) X n + k 2k 2k (−1) q 2 (2.2) 2k k k + s (−q; q)2 k=0 k   2 n2−s2 n (q; q)n−s(q; q)n+s  s 2 (−1) q n−s 2 2 2 , if n ≡ s (mod 2), = 2 q2 (q ; q )n 0, otherwise.

Proof. We may rewrite the left-hand side of (2.2) as

n −n n+1 1 1 k+(n) X (q ; q)k(q ; q)k(q 2 ; q)k(−q 2 ; q)kq 2 (q; q)k(q; q)k−s(q; q)k+s(−q; q)k k=s n −n n+1 2 s+( )  s−n n+s+1 s+ 1 s+ 1  (q ; q)s(q ; q)s(q; q )2sq 2 q , q , q 2 , −q 2 = 2 2 4φ3 s+1 s+1 2s+1 ; q, q . (q ; q )s(q; q)2s q , −q , q truncated basic hypergeometric series 7

The result then follows from Andrews’ terminating q-analogue of Watson’s formula [4, (II.17)]:  0, if n is odd, q−n, a2qn+1, b, −b   n 2 2 2 2 (2.3) 4φ3 2 ; q, q = b (q, a q /b ; q ) aq, −aq, b n/2 , if n is even  2 2 2 2 (a q , b q; q )n/2

s s+ 1 with the substitution of n, a and b by n − s, q and q 2 , respectively. 

Proof of Theorem 1.1. By the congruence (2.1), we have

p−1 2 2 X 2k  2k  q2k k k + s (−q2; q2)2(−q; q)2 k=0 q2 q2 k 2k p−1 2  p−1      k X + k 2k 2k (−1) 2 ≡ 2 qk +2k−pk (mod [p]2). 2k k k + s (−q2; q2)2 k=0 q2 q2 q2 k p−1 2 The proof then follows from (2.2) with n = 2 and q → q . 

3 Proof of Theorem 1.5

We first establish four lemmas to make the proof easier. Lemma 3.1. Let n be a nonnegative integer. Then n n−k+1   n ( 2 ) X n−k n (aq ; q)kq (ax; q)n(q; q)n (3.1) (−1) −k = −n , k (a; q)k(1 − xq ) (a; q)n(xq ; q)n+1 k=0 n   (n+1) X n k (q; q)n−k q 2 n−k (2) (3.2) (−1) q = n . k (x; q)n−k+1 1 − xq k=0 Proof. For (3.1), by the partial fraction decomposition we have n (ax; q)n(q; q)n X ak (3.3) −n = −k (a; q)n(xq ; q)n+1 1 − xq k=0 with −k   n (1 − xq )(ax; q)n(q; q)n n−k+1 n (aq ; q)k n−k ( 2 ) ak = lim −n = (−1) q . x→qk (a; q)n(xq ; q)n+1 k (a; q)k By the Gauss or q-binomial inversion (see, for example, [1, p. 77, Exercise 2.47]), the identity (3.2) is equivalent to n   (q; q)n X n k+1 1 k ( 2 ) = (−1) q k , (x; q)n+1 k 1 − xq k=0 q n which corresponds to the a = 0 case of (3.3) with x → xq .  8 V. J. W. Guo and J. Zeng

Lemma 3.2. Let n be a positive integer. Then (3.4)

(x; q)n + (a/x; q)n = (x; q)n(a/x; q)n n−1 n k   X (x; q)k(a/x; q)k(1 − q ) X k j j (2) k+j + n−k (−1) q (aq ; q)n−k, (q; q)k(1 − q ) j k=0 j=0 (3.5) (x; q)n + (a/x; q)n = (x; q)n(a/x; q)n + (a; q)n n−1 n−k    (j)+kj j X X n − k − 1 k + j − 1 q 2 a + (x; q) (a/x; q) (1 − qn) (−1)j . k k j − 1 j − 1 1 − qj k=1 j=1 −m Proof. We first prove (3.4). Taking x = q (0 6 m 6 n − 1), we have (3.6) n−1 n k   X (x; q)k(a/x; q)k(1 − q ) X k j j (2) k+j n−k (−1) q (aq ; q)n−k (q; q)k(1 − q ) j k=0 j=0 n−1 −m m n k   X (q ; q)k(aq ; q)k(1 − q ) X k j j (2) k+j = n−k (−1) q (aq ; q)n−k (q; q)k(1 − q ) j k=0 j=0 m   m n k   j X m k (aq ; q)k(1 − q ) X k j (aq ; q)n k (2)−mk j (2) = (−1) q n−k (−1) q j k (1 − q ) j (aq ; q)k k=0 j=0 m   m   m n X m j X m − j k (aq ; q)k(1 − q ) j (2) j k (2)−mk = (−1) q (aq ; q)n (−1) q j n−k . j k − j (aq ; q)k(1 − q ) j=0 k=j It follows from (3.1) that

m   m n X m − j k (aq ; q)k(1 − q ) k (2)−mk (−1) q j n−k k − j (aq ; q)k(1 − q ) k=j m m   m+j n (aq ; q)j X m − j m−k+1 m+1 (aq ; q)k−j(1 − q ) k ( 2 )−( 2 ) = j (−1) q 2j n−k (aq ; q)j k − j (aq ; q)k−j(1 − q ) k=j m m n+j n −(m+1) (−1) (aq ; q)j(aq ; q)m−j(q; q)m−j(1 − q )q 2 = j 2j n−m (aq ; q)j(aq ; q)m−j(q ; q)m−j+1 m n+j n −(m+1) (−1) (a; q)j(aq ; q)m−j(q; q)m−j(1 − q )q 2 = n−m . (a; q)m(q ; q)m−j+1 Therefore, the right-hand side of (3.6) can be simplified as (3.7) m+1 m n −( 2 )   (a; q)m+n(1 − q )q X m−j m j (q; q)m−j m (−1) q(2) = (aq ; q) , (a; q) j (qn−m; q) n m j=0 m−j+1 truncated basic hypergeometric series 9

−m where the equality follows from (3.2). Noticing that (q ; q)n = 0 for 0 6 −m m 6 n − 1, we have proved that both sides of (3.4) are equal for x = q m (0 6 m 6 n − 1), and by symmetry, for x = aq (0 6 m 6 n − 1) too. Furthermore, both sides of (3.4) are of the form x−nP (x) with P (x) being a n polynomial in x of degree 2n with the leading coefficient (−1)nq(2). Hence, they must be identical. This proves (3.4). By the q-binomial theorem (see, for example, [2, Theorem 3.3]), for k > 1, we have k   X j k j k+j (3.8) (−1) q(2)(aq ; q) j n−k j=0 k   n−k   X j k j X i n − k i +(k+j)i i = (−1) q(2) (−1) q(2) a j i j=0 i=0 n−k   k   X i n − k i +ik i X j k j +ij = (−1) q(2) a (−1) q(2) i j i=0 j=0 n−k   X i n − k i i +ik i = (−1) (q ; q) q(2) a . i k i=1

Moreover, for k = 0, the left-hand side of (3.8) is clearly equal to (a; q)n. Noticing that   i    n − k (q ; q)k n − k − 1 k + i − 1 1 n−k = i , i (q; q)k(1 − q ) i − 1 i − 1 1 − q we complete the proof of (3.5). 

Let n and h be positive integers and let m and s be nonnegative integers such that s 6 m and h 6 n − m (so n > m). Let

−n −n f(x; j, k) := (x; q)j(x; q)k(q ; q)j(q ; q)k j−m−h+1 k−m−h+1 2j+k 2k+j (q ; q)h−1(q ; q)h−1(q − q ) × m−s−1 2 , (−1) (q; q)n(q; q)h−1(q; q)j−s(q; q)j+s(q; q)k−s(q; q)k+s and, for integers a, b > s, let a b X X La,b(x) := f(x; j, k). j=s k=s

m2+3m−s2+s 2 Lemma 3.3. Let A = 2 − m(n + h) − h + h. Then s+1 n−s−h A (x; q)s(x; q)m+h(q /x; q)n−s−hx q (3.9) Lm,n(x) = . (q; q)m−s(q; q)m+s(q; q)n−s(q; q)n+s(q; q)n−m−h 10 V. J. W. Guo and J. Zeng

Proof. Without loss of generality, we assume that q is a complex number with |q| < 1. We first note that f(x; j, k) = −f(x; k, j), and so Lm,m(x) = 0. Since both sides of (3.9) are polynomials in x of degree m+n with the same leading coefficient, it suffices to show that these two polynomials have m+n common roots, counted with multiplicity. We proceed by dividing the roots of the right-hand side of (3.9) into four cases.

2 • If s > 1, then it is easily seen that (x; q)s divides Lm,n(x), which −r means that the numbers q (0 6 r 6 s − 1) are roots of Lm,n(x) with multiplicity 2.

−r • For r with s 6 r 6 m, we have (q ; q)k = 0 if k > m, and so −r −r Lm,n(q ) = Lm,m(q ) = 0.

• For r with m + 1 6 r 6 m + h − 1, we have r − m − h + 1 6 0, and so k−m−h+1 (q ; q)h−1 = 0 for m + 1 6 k 6 r, while for r < k 6 n we have −r −r −r (q ; q)k = 0. Hence, we again get Lm,n(q ) = Lm,m(q ) = 0.

r • For r with s + 1 6 r 6 n − h, it is clear that Lm,n(q ) = 0 follows from the identity

n −n r k−m−h+1 k−j 2j+k X (q ; q)k(q ; q)k(q ; q)h−1(1 − q )q (3.10) = 0, (q; q)k−s(q; q)k+s k=s which will be proved as follows: The left-hand side of (3.10) can be written as (3.11)

−n r (q ; q)s(q ; q)s n −n+s r+s k−m−h+1 k−j 2j+k X (q ; q)k−s(q ; q)k−s(q ; q)h−1(1 − q )q × (q; q)k−s(q; q)k+s k=s n−s   −n r X k n − s −(n−s)k+ k = (q ; q) (q ; q) (−1) q (2)R , s s k k k=0 where r+s k+s−m−h+1 k+s−j 2j+k+s (q ; q)k(q ; q)h−1(1 − q )q Rk = . (q; q)k+2s Since (qr+s; q) (qk+2s+1; q) k = r−s−1 , (q; q)k+2s (q; q)r+s−1 truncated basic hypergeometric series 11

k we see that Rk is a polynomial in q of degree r−s−1+h−1+2 6 n−s with constant term 0. By the q-binomial theorem (see [2, Theorem 3.3]),

n   X k n k+1 k (−1) q( 2 )x = (xq; q) , k n k=0 q

we have n   ( X k n k+1 −ik 0, for 1 6 i 6 n, (3.12) (−1) q( 2 ) = k (q; q) , for i = 0. k=0 q n

It follows that the right-hand side of (3.11) is equal to 0. Namely, the identity (3.10) holds.

Thus, we have found out all the m + n roots of Lm,n(x), which are clearly the same as those of the right-hand side of (3.9). This completes the proof. 

Remark. We can use the identity (3.12) to give a short proof of Jackson’s terminating q-analogue of Dixon’s identity, see [5].

Lemma 3.4. Let n and h be positive integers and let m and s be nonnegative integers with h 6 n − m and s 6 m. Then (3.13) m n −2n 2 −2n 2 k−j j+k+jh    X X (q ; q )j(q ; q )k(1 − q )q k − m − 1 m + h − j − 1 (q; q)j−s(q; q)j+s(q; q)k−s(q; q)k+s h − 1 h − 1 j=sk=m+h n−m−h 2 2 2 m2−n2−2mn+mh (−1) (q ; q )n(−q; q)2n−hq = 2 2 2 2 2 2 . (q; q)m−s(q; q)m+s(q ; q )n−s(q ; q )n+s(q; q)h−1(q , q )n−m−h

k−m−1 Proof. By the definition of q-binomial coefficients, there holds h−1 = 0 for m+1 6 k < m+h. Hence, the left-hand side of (3.13) remains unchanged Pn Pn if we replace k=m+h by k=m+1. Furthermore, we have

k − m − 1m + h − j − 1 h − 1 h − 1 j−m−h+1 k−m−h+1 (m−j)(h−1)−(h) (q ; q)h−1(q ; q)h−1q 2 = h−1 2 . (−1) (q; q)h−1

−n The proof then follows from the identity (3.9) with x = −q .  12 V. J. W. Guo and J. Zeng

Proof of Theorem 1.5. The left-hand side of (1.11) may be expanded as

(3.14) n X (q−2n; q2)2q2k k (x; q) (q/x; q) (q; q)2 (q; q)2 k k k=s k−s k+s −2n 2 −2n 2 j+k  X (q ; q )j(q ; q )kq (x; q)j(q/x; q)k + (x; q)k(q/x; q)j + . (q; q)j−s(q; q)j+s(q; q)k−s(q; q)k+s s6j

For 0 6 j < k, from (3.5) we deduce that

(3.15) (x; q)j(q/x; q)k + (x; q)k(q/x; q)j j j+1  = (x; q)j(q/x; q)j (xq ; q)k−j + (q /x; q)k−j 2j+1 = (x; q)k(q/x; q)k + (x; q)j(q/x; q)j(q ; q)k−j k−j−1 X k−j + (x; q)j+i(q/x; q)j+i(1 − q ) i=1 k−j−i    (h+1)+(i+2j)h X k − j − i − 1 i + h − 1 q 2 × (−1)h , h − 1 h − 1 1 − qh h=1

= (x; q)k(q/x; q)k + (x; q)j(q/x; q)j k−j−1 X k−j + (x; q)j+i(q/x; q)j+i(1 − q ) i=0 k−j−i    (h+1)+(i+2j)h X k − j − i − 1 i + h − 1 q 2 × (−1)h , h − 1 h − 1 1 − qh h=1 where in the last step we have used the q-binomial theorem:

k−j   2j+1 X h k − j h+1 +2jh (q ; q) = 1 + (−1) q( 2 ) . k−j h h=1

Pn By (3.15), we may write (3.14) as m=s am(x; q)m(q/x; q)m, where

n −2n 2 −2n 2 j+m X (q ; q )j(q ; q )mq (3.16) a = m (q; q) (q; q) (q; q) (q; q) j=s j−s j+s m−s m+s m n −2n 2 −2n 2 k−j j+k X X (q ; q )j(q ; q )k(1 − q )q + (q; q)j−s(q; q)j+s(q; q)k−s(q; q)k+s j=s k=m+1 k−m    (h+1)+(m+j)h X k − m − 1 m + h − j − 1 q 2 × (−1)h . h − 1 h − 1 1 − qh h=1 truncated basic hypergeometric series 13

It is easy to see that

n −2n 2 j −2n 2 s n −2n+2s 2 j−s X (q ; q )jq (q ; q )sq X (q ; q )jq = (q; q) (q; q) (q; q) (q; q) (q2s+1; q) j=s j−s j+s 2s j=s j−s j−s −2n 2 s  −n+s −n+s  (q ; q )sq q , −q , = 2φ1 2s+1 ; q, q (q; q)2s q −2n 2 n+s+1 s−(n−s)2 n−s (q ; q )s(−q ; q)n−sq = (−1) 2s+1 (q; q)2s(q ; q)n−s by the q-Chu-Vandermonde summation formula [4, Appendix (II.6)]. Hence,

n −2n 2 −2n 2 j+m X (q ; q )j(q ; q )mq (3.17) (q; q) (q; q) (q; q) (q; q) j=s j−s j+s m−s m+s (q−2n; q2) (q−2n; q2) (−qn+s+1; q) qm+s−(n−s)2 = (−1)n−s s m n−s (q; q)m−s(q; q)m+s(q; q)n+s n−m 2 2 2 m2−n2−2mn (−1) (q ; q )n(−q; q)2nq = 2 2 2 2 2 2 . (q; q)m−s(q; q)m+s(q ; q )n−s(q ; q )n+s(q , q )n−m Substituting (3.17) and (3.13) into (3.16), we obtain (3.18) 2 2 n−m h+1 n−m 2 2 2 m −n −2mn ( 2 )+2mh (−1) (q ; q )nq X (−q; q)2n−hq am = 2 2 2 2 h 2 2 . (q; q)m−s(q; q)m+s(q ; q )n−s(q ; q )n+s (−1) (q; q)h(q , q )n−m−h h=0 Replacing h by n − m − h, we have (3.19) n−m h (h+1)+2mh X (−1) (−q; q)2n−hq 2 2 2 (q; q)h(q , q )n−m−h h=0 n−m m−n n+m+1 (−q; q)n+m n−m n−m+1 +2m(n−m) X (q ; q)h(−q ; q)h −2hm = (−1) q( 2 ) q (q; q)n−m (−q; q)h(q; q)h h=0  −(n−m) m+n+1  (−q; q)n+m n−m (n−m+1)+2m(n−m) q , −q −2m = (−1) q 2 2φ1 ; q, q (q; q)n−m −q −n−m (−q; q)n+m(q ; q)n−m n−m n−m+1 +2m(n−m) = (−1) q( 2 ) (q; q)n−m(−q; q)n−m 2 2 (q ; q )m+n = 2 2 , (q ; q )n−m(q; q)2m where we have used the q-Chu-Vandermonde summation formula [4, Ap- pendix (II.7)]. It follows from (3.18) and (3.19) that am is just the coeffi- cient of (x; q)m(q/x; q)m in the right-hand side of (1.11). This completes the proof.  14 V. J. W. Guo and J. Zeng

4 Proof of Theorem 1.3

We first give a congruence modulo [p].

Lemma 4.1. Let m and r be two positive integers with p - m. Let s 6 r m−r min{h− m ip, h− m ip} be a nonnegative integer. Then the following congru- ence holds modulo [p]:

(4.1)

p−1 2 m 2m r m mk X (q ; q )k(q ; q )kq m m m m (q ; q )k−s(q ; q )k+s k=s r  (h− m ip+s)m r m 2m −mh− ip m q 2 (q ; q ) h− r i −s (q m ; q )  m p s  2 r  , if h− ip ≡ s mod 2, ≡ 2m 2m r m (q ; q ) h− m ip+s  2 0 otherwise.

r m−r p−1 Proof. It is easy to see that h− m ip+h− m ip = p−1, and so s 6 2 . Since p m 2m p+1 is an odd prime, we see that (q ; q )k ≡ 0 (mod [p]) for 2 6 k 6 p−s−1, which means that

p−1 2 m 2m r m mk p−s−1 m 2m r m mk X (q ; q )k(q ; q )kq X (q ; q )k(q ; q )kq m m m m ≡ m m m m (mod [p]). (q ; q )k−s(q ; q )k+s (q ; q )k−s(q ; q )k+s k=s k=s

r mh− m ip+r r Let a = p . Then m|r − ap and r − ap = −mh− m ip 6 0. It is clear r m r−ap m r−ap m r that (q ; q )k ≡ (q ; q )k (mod [p]) and (q ; q )k = 0 for k > h− m ip. m−r r Moreover, we have p − s − 1 > p − h− m ip − 1 = h− m ip > s, and therefore

p−s−1 m 2m r−ap m mk X (q ; q )k(q ; q )kq m m (q ; q )k−s(q; q)k+s k=s h− r i m p m 2m r−ap m mk X (q ; q )k(q ; q )kq ≡ m m m m (q ; q )k−s(q ; q )k+s k=s r m 2m −mh− ip m ms (q ; q )s(q m ; q )sq = m m (q ; q )2s  −m(h− r i −s) (s+ 1 )m (s+ 1 )m  q m p , q 2 , −q 2 × φ ; qm, qm (mod [p]). 3 2 0, q(2s+1)m

The proof then follows from Andrews’ identity (2.3).  truncated basic hypergeometric series 15

p−1 Proof of Theorems 1.3. By Lemma 2.1, for 0 6 k 6 2 , we have m m   (q ; q )2k 2k 1 2m 2m 2 = m m (q ; q )k k q2m (−q ; q )2k  p−1  k mk2−mkp 2 + k m m ≡ (−1) q (−q ; q )2k 2k q2m k mk2−mkp 2m 2m (−1) q (q ; q ) p−1 2 +k 2 = 2m 2m m m (mod [p] ), (q ; q ) p−1 (q ; q )2k 2 −k and so (4.2) p−1 2 m m r m m−r m mk X (q ; q )2k(q ; q )k(q ; q )kq m m m m 2m 2m 2 (q ; q )k−s(q ; q )k+s(q ; q ) k=s k p−1 k m m r m m−r m mk2−mk(p−1) 2 (−1) (q ; q ) p−1 (q ; q )k(q ; q )kq X 2 +k 2 ≡ 2m 2m m m m m m m (mod [p] ). (q ; q ) p−1 (q ; q )k−s(q ; q )k+s(q ; q )2k k=s 2 −k

m r p−1 Letting q → q , x = q and n = 2 in Theorem 1.5, we see that the right-hand side of (4.2) can be written as (4.3) 2 p−1 2m 2m 2m 2m (p−1) (−1) 2 (q ; q ) p−1 (q ; q ) p−1 q 4 2 −s 2 +s 2m 2m 2 (q ; q ) p−1 2  p−1  p−1  2 m(1−p) 2m r m mk 2 m(1−p) 2m m−r m mk X (q ; q )k(q ; q )kq X (q ; q )k(q ; q )kq × m m m m  m m m m  . (q ; q )k−s(q ; q )k+s (q ; q )k−s(q ; q )k+s k=s k=s

r If h− m ip ≡ s + 1 (mod 2), then by the congruence (4.1), we have

p−1 p−1 2 m(1−p) 2m r m mk 2 m 2m r m mk X (q ; q )k(q ; q )kq X (q ; q )k(q ; q )kq m m m m ≡ m m m m ≡ 0 mod [p], (q ; q )k−s(q ; q )k+s (q ; q )k−s(q ; q )k+s k=s k=s

m−r and also h− m ip ≡ s + 1 (mod 2) which means that

p−1 2 m(1−p) 2m m−r m mk X (q ; q )k(q ; q )kq m m m m ≡ 0 (mod [p]). (q ; q )k−s(q ; q )k+s k=s

2m 2m Noticing the fact (q ; q ) p−1 6≡ 0 (mod [p]), we conclude that the right- 2 hand side of (4.2) is congruent to 0 modulo [p]2. This proves (1.8). To prove (1.9), just observe that (see the proof of Lemma 4.1)

r m m−r m (q ; q )k ≡ (q ; q )k ≡ 0 (mod [p]) 16 V. J. W. Guo and J. Zeng

 r m−r for max h− m ip, h− m ip < k 6 p − 1, and

m m r m m−r m (q ; q )2k (q ; q )k(q ; q )k ≡ m m m m ≡ 0 (mod [p]) (q ; q )k−s(q ; q )k+s

p−1  r m−r for 2 < k 6 max h− m ip, h− m ip . Finally, the proof of (1.10) follows from factorizing (4.2) into (4.3), apply- ing the first case of the congruence (4.1), and then using the aforementioned r m−r relation h− m ip + h− m ip = p − 1. 

5 Proof of Theorems 1.6 and 1.7

The following lemma is probably known. For the reader’s convenience, we include a proof.

Lemma 5.1. Let m, n and s be nonnegative integers with s 6 n. Then n    X k n m + k k −nk n − n+1 (5.1) (−1) q(2) = (−1) q ( 2 ), k n k=0 n     k2−k−2nk X k n + k 2k + 2s q n −n(n+1) (5.2) (−1) 2k+1 = (−1) q . 2k 2 k + s 2 (−q ; q)2s k=0 q q

Proof. It is not difficult to see that the two identities (5.1) and (5.2) are equivalent, respectively, to

n    X k n m + n − k k+1 (5.3) (−1) q( 2 ) = 1, k n k=0 n     2n−2k+1 2 X n 2n − k (q ; q )s k+1 k 2( 2 ) (5.4) (−1) 2n−2k+2 2 q = 1. k 2 n 2 (q ; q )s k=0 q q

m+n−k −k Since n can be written as a polynomial in q of degree n with con- stant term 1/(q; q)n. Identity (5.3) then follows from (3.12). On the other hand, since 0 6 s 6 n, we see that

  2n−2k+1 2 2n−2k+2s+2 2 2n−2k+1 2 2n − k (q ; q )s (q ; q )n−s(q ; q )s 2n−2k+2 2 = 2 2 n q2 (q ; q )s (q ; q )n

−2k 1 is a polynomial in q of degree n with constant term 2 2 . Therefore, (q ;q )n the identity (5.4) follows from (3.12) with q → q2. This completes the proof.  truncated basic hypergeometric series 17

Proof of Theorem 1.6. It is easy to see that

2   (q; q )k 2k 1 2 2 = . (q ; q )k k q2 (−q; q)2k Hence, by Lemmas 2.1 and 5.1, we have

p−1 p−1 2 2 2 2     X (q; q )k(q; q )k+s X 2k 2k + 2s 1 2 2 2 2 = (q ; q )k(q ; q )k+s k 2 k + s 2 (−q; q)2k(−q; q)2k+2s k=0 k=0 q q p−1 2  p−1    k2−kp X + k 2k + 2s (−q; q)2kq ≡ (−1)k 2 2k 2 k + s 2 (−q; q)2k+2s k=0 q q 2 p−1 1−p 2 = (−1) 2 q 4 (mod [p] ), as desired.  r mh− m ip+r Proof of Theorem 1.7. Again, let a = p . Then a is a positive integer, m|ap − r, and so

r m k mj+r−m (q ; q )k Y 1 − q (5.5) = (qm; qm) 1 − qmj k j=1 k Y (1 − qap−mj−r+m)qmj+r−m ≡ (−1)k 1 − qmj j=1  ap−r  k mk(k−1) +kr = (−1) m q 2 k qm  r  k h− ip mk(k−1) −mkh− r i ≡ (−1) m q 2 m p (mod [p]), k qm

m−r m k+s ap+mj−r  ap−r  (q ; q )k+s Y 1 − q + k + s (5.6) ≡ = m (qm; qm) 1 − qmj k + s k+s j=1 qm h− r i + k + s = m p (mod [p]). k + s qm By the congruences (5.5) and (5.6), we have

p−s−1 r m m−r m X (q ; q )k(q ; q )k+s m m m m (q ; q )k(q ; q )k+s k=0 p−s−1  r   r  X k h− ip h− ip + k + s mk(k−1) −mkh− r i ≡ (−1) m m q 2 m p k k + s k=0 qm qm −mh− r i (h− r i +1) h− r i m p m p = (−1) m p q 2 (mod [p]), 18 V. J. W. Guo and J. Zeng

m−r r where in the last step we have used p − s − 1 > p − h− m ip − 1 = h− m ip and the identity (5.1). This proves (1.12). r(m−r)(1−p2) To prove (1.13), just notice that if p ≡ ±1 (mod m), then 2m is an integer and

−mh− r i (h− r i + 1) r(m − r)(1 − p2) m p m p ≡ (mod p). 2 2m 

6 Concluding remarks and open problems

It seems that the congruence (1.9) can be further generalized as follows.

Conjecture 6.1. Let m and r be two positive integers with p - m. Let r s 6 p − 1 be a nonnegative integer. If h− m ip ≡ s + 1 (mod 2), then

p−1 m m r m m−r m mk X (q ; q )2k(q ; q )k(q ; q )kq 2 (6.1) m m m m 2m 2m 2 ≡ 0 (mod [p] ). (q ; q )k−s(q ; q )k+s(q ; q ) k=s k

r m−r Note that, if s > max{h− m ip, h− m ip}, then the congruence (6.1) is obviously true, since in this case each summand in the left-hand side is congruent to 0 modulo [p]2. It is easy to see that, when r = 1, m = 3, 4, 6 and q → 1, Conjecture 6.1 reduces to a result of Z.-W. Sun [14, Theorem 1.3(i)]. We conjecture that Theorem 1.7 can be further strengthened.

Conjecture 6.2. Let m and r be positive integers with p ≡ ±1 (mod m) m−r and r < m. Then for any integer s with 0 6 s 6 h− m ip, there holds

p−s−1 r m m−r m 2 X (q ; q )k(q ; q )k+s r r(m−r)(1−p ) h− m ip 2m 2 m m m m ≡ (−1) q (mod [p] ). (q ; q )k(q ; q )k+s k=0 Like [6, Conjecture 7.1], Conjecture 6.2 seems to have a further general- ization as follows:

Conjecture 6.3. Let m and |r| be positive integers with p - m and m - r. Then there exists a unique integer fp,m,r such that, for any s with 0 6 s 6 m−r h− m ip, there holds

p−s−1 r m m−r m X (q ; q )k(q ; q )k+s r h− m ip fp,m,r 2 m m m m ≡ (−1) q (mod [p] ). (q ; q )k(q ; q )k+s k=0 truncated basic hypergeometric series 19

Furthermore, the numbers fp,m,r satisfy the symmetry fp,m,r = fp,m,m−r and the recurrence relation: ( −fp,m,r, if r ≡ 0 (mod p), fp,m,m+r = fp,m,r − r, otherwise.

Here we give some values of fp,m,r:

f3,2,1 = −2, f3,2,3 = −3, f3,2,5 = 3, f3,2,7 = −2, f3,2,9 = −9,

f3,2,11 = 9, f3,2,13 = −2, f5,3,1 = −8, f5,3,2 = −8, f5,3,4 = −9,

f5,3,5 = −10, f5,3,7 = −13, f5,3,8 = 10, f5,3,10 = −20, f5,3,11 = 2,

f5,3,13 = 20, f5,3,14 = −9, f5,3,16 = 7, f5,3,17 = −23, f5,3,19 = −9,

f5,8,1 = −23, f7,9,1 = −54, f7,9,2 = −21, f7,9,4 = −37, f7,9,5 = −37,

f7,9,7 = −21, f7,9,8 = −54, f7,9,10 = −55, f7,9,11 = −23 f7,9,13 = −41,

f7,9,14 = −42, f7,9,16 = −22, f7,9,17 = −33.

Finally, supercongruences (or q-supercongruences) have now been around for quite a time, and it would be desirable to have some more con- ceptual proofs of these phenomena, such as, combinatorial interpretations, connections to elliptic curves or representations of p-adic groups.

Acknowledgement

This work was partially supported by the National Natural Science Foun- dation of China (grant no. 11371144).

References

[1] M. Aigner, A Course in Enumeration, Graduate Texts in Mathematics, Vol. 238, Springer, 2007. [2] G. E. Andrews, The Theory of Partitions, Cambridge University Press, Cam- bridge, 1998. [3] S. Chowla, B. Dwork, and R. Evans, On the mod p2 determination of (p−1)/2 (p−1)/4 , J. Number Theory 24 (1986), 188–196. [4] G. Gasper and M. Rahman, Basic Hypergeometric Series, 2nd ed., Encyclo- pedia Math. Appl. 96, Cambridge Univ. Press, Cambridge, 2004. [5] V. J. W. Guo and J. Zeng, A short proof of the q-Dixon identity, Discrete Math. 296 (2005), 259–261. [6] V. J. W. Guo and J. Zeng, Some q-analogues of supercongruences of Rodriguez-Villegas, J. Number Theory 145 (2014), 301–316. 20 V. J. W. Guo and J. Zeng

[7] E. Mortenson, A supercongruence conjecture of Rodriguez-Villegas for a cer- tain truncated hypergeometric function, J. Number Theory 99 (2003), 139– 147.

[8] E. Mortenson, Supercongruences between truncated 2F1 hypergeometric func- tions and their Gaussian analogs, Trans. Amer. Math. Soc. 355 (2003), 987– 1007.

[9] E. Mortenson, Supercongruences for truncated n+1Fn hypergeometric series with applications to certain weight three newforms, Proc. Amer. Math. Soc. 133 (2005), 321–330. (p−1)/2 2 [10] H. , An elementary approach to (p−1)/4 modulo p , Taiwanese J. Math. 16 (2012), 2197–2202. [11] F. Rodriguez-Villegas, Hypergeometric families of Calabi-Yau manifolds, in: Calabi-Yau Varieties and Mirror Symmetry (Toronto, ON, 2001), Fields Inst. Commun., 38, Amer. Math. Soc., Providence, RI, 2003, 223–231. [12] Z.-H. Sun, Congruences concerning Legendre polynomials II, J. Number The- ory 133 (2013), 1950–1976. [13] Z.-H. Sun, Generalized Legendre polynomials and related supercongruences, J. Number Theory 143 (2014), 293–319. [14] Z.-W. Sun, On sums involving products of three binomial coefficients, Acta Arith. 156 (2012) 123–141. [15] Z.-W. Sun, Supercongruences involving products of two binomial coefficients, Finite Fields Appl. 22 (2013), 24–44. [16] H. Swisher, On the supercongruence conjectures of van Hamme, preprint, 2015, arXiv:1504.01028v1. [17] L. van Hamme, Some conjectures concerning partial sums of generalized hy- pergeometric series, in: p-Adic Functional Analysis (Nijmegen, 1996), Lec- ture Notes in Pure and Appl. Math., Vol. 192, Dekker, 1997, 223–236.