Skip to content
BY 4.0 license Open Access Published by De Gruyter January 13, 2021

Quantum modularity of partial theta series with periodic coefficients

  • Ankush Goswami and Robert Osburn EMAIL logo
From the journal Forum Mathematicum

Abstract

We explicitly prove the quantum modularity of partial theta series with even or odd periodic coefficients. As an application, we show that the Kontsevich–Zagier series t(q) which matches (at a root of unity) the colored Jones polynomial for the family of torus knots T(3,2t), t2, is a weight 32 quantum modular form. This generalizes Zagier’s result on the quantum modularity for the “strange” series F(q).

MSC 2010: 11F37; 33D15; 57K16

1 Introduction

In [30], Zagier introduced the notion of a quantum modular form of weight k12 as a function g: for which the function rγ:{γ-1(i)} given by

g(α)-(cα+d)-kg(aα+bcα+d)=:rγ(α)

extends to a real-analytic function on 1()Sγ, where Sγ is a finite set, for each γ=(abcd)SL2(). Suitable modifications can be made to restrict the domain of rγ to appropriate subsets of and allow both multiplier systems and transformations on subgroups of SL2(). Since their inception, there has been substantial interest in studying these modular objects which emerge in diverse contexts: Maass forms [9], supersymmetric quantum field theory [12], topological invariants for plumbed 3-manifolds [5, 10, 11], combinatorics [13, 20], unified Witten–Reshetikhin–Turaev invariants [19] and L-functions [24, 26]. For more examples, see [4, Chapter 21].

One of the most influential of the original five examples from [30] is the Kontsevich–Zagier “strange” series [29]

(1.1)F(q):=n0(q)n,

where

(a1,a2,,aj)n=(a1,a2,,aj;q)n:=k=1n(1-a1qk-1)(1-a2qk-1)(1-ajqk-1)

is the standard q-hypergeometric notation, valid for n0{}. F(q) is “strange” in the sense that it does not converge on any open subset of , but is well-defined when q is a root of unity (where it is finite). Zagier proves that, for α, ϕ(α):=eπiα12F(e2πiα) is a quantum modular form of weight 32 on with respect to SL2(). The key to proving this result is the “strange identity”

(1.2)F(q)=-12n1n(12n)qn2-124

where = means that the two sides agree to all orders at every root of unity (for further details, see [29, Sections 2 and 5]) and (12*) is the quadratic character of conductor 12. The idea is to prove quantum modular properties for the right-hand side of (1.2) which are then inherited by F(q). The purpose of this paper is to place the right-hand side of (1.2) and other examples in the literature into the general context of quantum modularity of partial theta series with even or odd periodic coefficients. Before stating our main result, we introduce some notation.

Let f: be an even or odd function with period M2. For any fixed 1k0<2M, consider the set

𝒮(k0):={1kM2:k2k0(mod2M)}.

Let f(k0)𝒮(k0) be non-empty and such that f(j)=0 whenever jf(k0){M-k:kf(k0)}. Clearly, Sf(k0):=f(k0){M-k:kf(k0)} is the support of f. Consider the following partial theta series:

(1.3)θf(z):=n0f(n)qn22M,Θf(z):=n0nf(n)qn22M,

where q=e2πiz, z. For N, let

Γ1(N):={(abcd)SL2():c0(modN),ad1(modN)},

and let ΓM be defined as Γ1(2M) if M is even and

(1.4){(abcd)Γ1(2M):b0(mod2)}

if M is odd. Consider the set

(1.5)BM:={α:αisΓM-equivalent toi}

and let AM be defined by (1.5) if kf(k0)f(k)0 and by

{α:αisΓM-equivalent to 0ori}

otherwise. Here and throughout, means that, whenever M2f(k0), we replace f(M2) in the sum by 12f(M2). We also employ the convention that f(n)=0 if n. For k12 and γ=(abcd)ΓM, we define the Petersson slash operator |k,χ by

(g|k,χγ)(τ):=χ(γ)¯(cτ+d)-kg(aτ+bcτ+d),

where τ and χ is a multiplier. Finally, we write () for the extended Jacobi symbol and let εd=1 or i according as d1 or 3(mod4). Our main result is now as follows.

Theorem 1.1.

Let f be a function with period M2 and support Sf(k0). Let αQ. If f is even, then Θf(α) is a quantum modular form of weight 32 on AM with respect to ΓM. If f is odd, then θf(α) is a “strong” quantum modular form of weight 12 on Q with respect to ΓM and is a quantum modular form of weight 12 on BM with respect to ΓM.

Remark 1.2.

(i) The main novelty of Theorem 1.1 is that one does not require Θf(z) or θf(z) to be a cusp form. For example, consider θψ(z), where ψ is given in Section 4.2 (cf. [20]). Otherwise, one can invoke (2.4), (2.9) and [8, Theorem 1.1].

(ii) In Theorem 1.1, Θf(z) satisfies

Θf(α)-(Θf|32,χγ)(α)=rγ,f(α)

for all γ=(abcd)ΓM and αAM, where

rγ,f(z)=-Meπi42πγ-1(i)iθf(τ)(τ-z¯)-32dτ.

Here, rγ,f: is a C function which is real-analytic in {γ-1(i)} and χ is a multiplier given by

(1.6)χ(γ)=eπiabk0M(2cMd)εd-1.

(iii) For τ-:={τ:Im(τ)<0}, let Θ^f(τ) denote the non-holomorphic Eichler integral

(1.7)Θ^f(τ):=1iMτ¯iΘf(w)(w-τ)-12dw.

In Theorem 1.1, θf(α) is a “strong” quantum modular form in the following sense (see [24] or [30]):

  1. θf and Θ^f “agree to infinite order” at all rational numbers (see Lemma 2.6);

  2. for τ- and γΓM, we have

    Θ^f(τ)-(Θ^f|12,χγ)(τ)=rγ,f(τ),

    where

    rγ,f(τ)=1iMγ-1(i)iΘf(w)(w-τ)-12dw.

Here, rγ,f(τ) is a holomorphic function in -, extends as a function to and is real-analytic in

{γ-1(i)}.

Also, χ is the multiplier as in (1.6). A close inspection of the techniques in [24] reveals that one needs convergence of Θ^f(τ) for τ- (and not necessarily at rational points) to deduce the strong quantum modularity property for θf(z). To ensure this condition, Θf(z) does not have to be a cusp form. For a similar approach, see [3, 17].

(iv) If f is a function with period M2 and support Sf(k0), then θf(z) is a sum of a modular form and a (strong) quantum modular form both of weight 12 and Θf(z) is a sum of a modular form and a quantum modular form both of weight 32. To see this, write f as f(n)=fe(n)+fo(n), where

fe(n):=f(n)+f(-n)2,fo(n):=f(n)-f(-n)2.

Clearly, fe(n) (respectively, fo(n)) is an even (respectively, odd) function of period M with support contained in Sf(k0). Indeed, if Sf,e(k0) denotes (respectively, Sf,o(k0)) the support of fe(n) (respectively, fo(n)), then Sf,e(k0)=f,e(k0){M-k:kf,e(k0)} (respectively, Sf,o(k0)=f,o(k0){M-k:kf,o(k0)}) for some f,e(k0),f,o(k0)f(k0). Thus, we have

θf(z)=n0fe(n)qn22M+n0fo(n)qn22M=:θf(e)(z)+θf(o)(z),
Θf(z)=n0nfe(n)qn22M+n0nfo(n)qn22M=:Θf(e)(z)+Θf(o)(z).

Now, apply Lemma 2.1 to θf(e)(z) and Theorem 1.1 to θf(o)(z). Similarly, apply Theorem 1.1 to Θf(e)(z) and Lemma 2.2 to Θf(o)(z).

(v) As pointed out by the referee, there is a “duality” in the proof of Theorem 1.1. If f is even, then the quantum modularity of Θf(z) is driven by the modularity of θf(z) and if f is odd, then the (strong) quantum modularity of θf(z) is driven by the modularity of Θf(z). See Lemmas 2.1 and 2.2.

The paper is organized as follows. In Section 2, we carefully study some important transformation and limiting properties of θf(z) and Θf(z). In Section 3, we prove Theorem 1.1. In Section 4, we give some examples, including the quantum modularity of the Kontsevich–Zagier series t(q) associated to the family of torus knots T(3,2t), t2. This latter result generalizes the quantum modularity of F(q).

2 Preliminaries

We begin with transformation properties of the partial theta series θf(z) and Θf(z) in (1.3).

Lemma 2.1.

Let f be an even function with period M2 and support Sf(k0). For all γ=(abcd)ΓM, we have

(2.1)θf(γz)=eπiabk0M(2cMd)εd-1(cz+d)12θf(z).

Proof.

From (1.3), we have

θf(z)=1k<Mn=0f(Mn+k)q(Mn+k)22M=1k<Mf(k)n0q(Mn+k)22M
(2.2)=1k<M2f(k)(n0q(Mn+k)22M+n0q(Mn+M-k)22M)+f(M2)n0q(Mn+M2)22M
(2.3)=1k<M2f(k)n=-q(Mn+k)22M+δf(M2)n=-q(Mn+M2)22M,

where

δf(M2):={12f(M2)ifMis even andM2f(k0),0otherwise.

Note that (2.3) follows by changing n-n-1 in the second sum in (2.2). Since support of f is Sf(k0), (2.3) yields

(2.4)θf(z)=f(k)θ(z;k,M),

where θ(z;k,M) is the theta series

θ(z;k,M):=n=-q(Mn+k)22M.

By [28, Proposition 2.1], we see that θ(z;k,M) satisfies

(2.5)θ(γz;k,M)=eπiabk2M(2cMd)εd-1(cz+d)12θ(z;ak,M)

for all γ=(abcd)ΓM. Also, since γΓM, we have for some integer j that

(2.6)θ(z;ak,M)=n=-q(Mn+ak)22M=n=-q(Mn+(1+2jM)k)22M=θ(z;k,M),

where n has been replaced by n-2jk in the second sum in (2.6). Noting that

(2.7)eπiabk2M=eπiabk0M,

(2.1) now follows from (2.4) and (2.5)–(2.7). ∎

Lemma 2.2.

Let f be an odd function with period M2 and support Sf(k0). For all γ=(abcd)ΓM, we have

(2.8)Θf(γz)=eπiabk0M(2cMd)εd-1(cz+d)32Θf(z).

Proof.

If M is even, then f(M2)=0 for odd f. So we have

Θf(z)=0k<Mn0(Mn+k)f(Mn+k)q(Mn+k)22M
=0kM2f(k)(n0(Mn+k)q(Mn+k)22M-n0(Mn+(M-k))q(Mn+M-k)22M)
=0kM2f(k)n=-(Mn+k)q(Mn+k)22M
(2.9)=kf(k0)f(k)Θ~(z;k,M),

where

Θ~(z;k,M)=n=-(Mn+k)q(Mn+k)22M.

Using [28, Proposition 2.1] (with A=[M], ν=1 and P(m)=m), we have

(2.10)Θ~(γz;k,M)=eπiabk2M(2cMd)εd-1(cz+d)32Θ~(z;ak,M)

for all γ=(abcd)ΓM. Also, since γΓM, we have for some integer j that

(2.11)Θ~(z;ak,M)=n=-(Mn+ak)q(Mn+ak)22M=n=-(Mn+(1+2jM)k)q(Mn+(1+2jM)k)22M=Θ~(z;k,M),

where n has been replaced by n-2jk in the sum in (2.11). Thus, combining (2.9)–(2.11) yields (2.8). ∎

Lemma 2.3.

Let f be an even function with period M2 and support Sf(k0). Then

(2.12)θf(z)=qk02M(qM;qM)qkf(k)(-qM2-k;qM)(-qM2+k;qM),

where, for kMf(k0), k=k2-k02MZ0.

Proof.

We have

(2.13)θ(z;k,M)=qk22Mn=-qM2n2+2kMn2M=qk22Mn=-(qk)n(qM2)n2=qk22M(qM;qM)(-qM2-k;qM)(-qM2+k;qM),

where (2.13) follows from Jacobi’s triple product identity

(2.14)n=-(-1)nznqn2=(q2;q2)(zq;q2)(z-1q;q2)

with z-qk and qqM2 in (2.14). As in the proof of Lemma 2.1, we have

(2.15)θf(z)=f(k)θ(z;k,M).

Thus, (2.13) and (2.15) imply (2.12), where, for kf(k0), k=k2-k02M. Since k implies k2k0, we conclude that k0. ∎

Lemma 2.4.

Let f be an even function with period M2 and support Sf(k0). Assume kMf(k0)f(k)=0. Let αQ be such that αγ(i) for any γΓM. Then we have

(2.16)θf(α+iy)=1M(y-iα)f(k)n=-n0e-πn2M(y-iα)+2πinkM.

Proof.

For any 1k<M, we obtain the following upon using [27, Chapter 5, page 76] with u=(y-iα)M and x=kM:

(2.17)θ(α+iy;k,M)=eπi(α+iy)k2+π(y-iα)k2MM(y-iα)n=-e-πn2M(y-iα)+2πinkM=1M(y-iα)(1+n=-n0e-πn2M(y-iα)+2πinkM).

As in the proof of Lemma 2.1, we have

(2.18)θf(z)=f(k)θ(z;k,M),

and so (2.16) follows from (2.17), (2.18) and kf(k0)f(k)=0. ∎

Corollary 2.5.

Let f be an even function with period M2 and support Sf(k0). Assume kM(k0)f(k)=0. Then we have

θf(iy)=e-πMyMy(cf(M,k0)+o(1)),

where o(1)0 as y0+ and

cf(M,k0):=2f(k)cos(2πkM).

Proof.

First, we note that γ(i)0 for all γΓM. Thus, we choose α=0 in Lemma 2.4 to get

(2.19)θf(iy)=1Myf(k)n=-n0e-πn2My+2πinkM=1Myf(k)n=1(rM(k,n)+rM(k,-n))q1n2,

where q1=e-πMy and rM(k,n)=e2πinkM. Interchanging the sums in (2.19), we find

θf(iy)=1Myn=1gM(k0,n)q1n2,

where

gM(k0,n):=f(k)(rM(k,n)+rM(k,-n)).

At this point, note that gM(k0,n)=O(1) and q10 as y0+. This yields the result. ∎

Let C: be a periodic function with mean value zero and consider the L-series

L(s,C):=n=1C(n)ns,(s)>0,

which has an analytic continuation to (see [24, Proposition, page 98]).

Lemma 2.6.

Let f be an odd function with period M2 and support Sf(k0). Then, as t0+, we have for (p,q)=1 that

(2.20)θf(pq+it2π)r=0L(-2r,Cf,k0)(-t2M)rr!,
(2.21)Θ^f(pq-it2π)r=0L(-2r,Cf,k0)(t2M)rr!,

where Cf,k0(n):=kMf(k0)f(k)Cf(n,k) and

Cf(n,k):={eπipn2Mq𝑖𝑓nk(modM),-eπipn2Mq𝑖𝑓n-k(modM)),0𝑜𝑡ℎ𝑒𝑟𝑤𝑖𝑠𝑒.

Thus, in the sense of Lawrence and Zagier [24, page 103], (2.20) and (2.21) imply that θf and Θ^f “agree to infinite order” at all rational numbers.

Proof.

For t>0, we have

(2.22)θf(pq+it2π)=n0f(n)eπipn2Mq-tn22M=kf(k0)f(k)(n>0nk(modM)eπipn2Mq-tn22M-n>0n-k(modM)eπipn2Mq-tn22M)=n1Cf,k0(n)e-tn22M.

For each kf(k0), Cf(n,k) is an odd function with period Mq or 2Mq according as 2 divides p or does not divide p. Also, Cf(n,k) has mean value zero. This implies that Cf,k0(n) is an odd function with period Mq or 2Mq according as 2 divides p or does not divide p and with mean value zero. Thus, by [24, Proposition, page 98] and (2.22), we obtain

θf(pq+it2π)r=0L(-2r,Cf,k0)(-t2M)rr!.

Next, we turn to Θ^f(τ), where τ=x+iy with y<0. First, we have, for w,

Θf(w)=kf(k0)f(k)(n>0nk(modM)neπin2wM-n>0n-k(modM)neπin2wM).

Thus, by the change of variable ww+τ and contour integration, it follows that

(2.23)Θ^f(τ)=1iMkf(k0)f(k)(n>0nk(modM)-n>0n-k(modM))neπin2τM-2iyieπin2wMw-12dw.

To evaluate the integral on the right-hand side of (2.23), we let wiMwπn2. This yields

(2.24)-2iyieπin2wMw-12dw=iMπn2-2πyn2Me-ww-12dw=iMπn2Γ(12,-2πyn2M),

where Γ(a,x) is the upper incomplete gamma function defined by

Γ(a,x):=xwa-1e-wdw.

Thus, (2.23) and (2.24) yield

(2.25)Θ^f(τ)=1πkf(k0)f(k)(n>0nk(modM)-n>0n-k(modM))eπin2τMΓ(12,-2πyn2M),

and so (2.25) implies

(2.26)Θ^f(pq-it2π)=1πn1Cf,k0(n)etn22MΓ(12,tn2M).

Since Cf,k0(n) is an odd periodic function, it now follows from [6, Lemma 4.3] and (2.26) that

Θ^f(pq-it2π)r=0L(-2r,Cf,k0)(t2M)rr!.

3 Proof of Theorem 1.1

We are now in a position to prove Theorem 1.1.

Proof of Theorem 1.1.

Let f be an even function with period M2 and support Sf(k0). Define the Eichler integral of θf(z) as follows:

θ~f(z):=ziθf(τ)(τ-z¯)-32dτ.

With Lemma 2.3 and Corollary 2.5, we see that θ~f(z) is well-defined on AM. Thus, for z=x+iyAM, it follows using contour integration that

(3.1)θ~f(z)=πMe-iπ4n0nf(n)Γ(-12,2πn2yM)eπin2Mz¯.

For αAM, we see from (3.1) and the fact that Γ(-12)=-2π,

(3.2)θ~f(α)=-2πMe-iπ4Θf(α).

For γ=(abcd)ΓM, it follows from Lemma 2.1, (3.2) and

d(γτ)=dτ(cτ+d)2,γτ-γz¯=τ-z¯(cτ+d)(cz¯+d)

that

Θf(α)-(Θf|32,χγ)(α)=:rγ,f(α),

where

rγ,f(z)=-Meiπ42πγ-1(i)iθf(τ)(τ-z¯)-32dτ.

It only remains to observe that rγ,f(z) is C and real-analytic in {γ-1(i)}.

Let f be an odd function with period M2 and support Sf(k0). For τ- and γΓM, it follows from (1.7) and Lemma 2.2 that

(3.3)Θ^f(τ)-(Θ^f|12,χγ)(τ)=rγ,f(τ),

where

rγ,f(τ)=1iMγ-1(i)iΘf(w)(w-τ)-12dw.

Since, by Lemma 2.6, θf and Θ^f “agree to infinite order” at all rational numbers and Θ^f(τ) satisfies the transformation property in (3.3) for all τ-, it follows in the sense of Lawrence and Zagier [24, page 103] that θf(z) is a strong quantum modular form of weight 12 on with respect to ΓM. It is also clear that rγ,f(τ) is a holomorphic function in -, extends as a function to and is real-analytic in {γ-1(i)}. Here, χ is the multiplier given by (1.6). Finally, we note that, as Θf(z) vanishes at z=i (and thus at all αBM), Θ^f is well-defined on BM. Thus, for τ-BM, Θ^f(τ) satisfies (3.3). Now, one can check that

(3.4)θf(z)=kf(k0)f(k)(n0nk(modM)-n0n-k(modM))qn22M.

For τ=αBM, it follows from (2.25), (3.4) and Γ(12)=π that θf(α)=Θ^f(α). Thus, θf(α) is a quantum modular form of weight 12 on BM with respect to ΓM. ∎

4 Examples

In this section, we illustrate Theorem 1.1 with four examples.

4.1 Kontsevich–Zagier series t(q) for torus knots T(3,2t)

Let K be a knot and JN(K;q) the usual colored Jones polynomial, normalized to be 1 for the unknot. For the importance of this quantum knot invariant, see, for example, [1, 14, 25, 30]. If T(3,2) is the right-handed torus knot, then [15, 23]

(4.1)JN(T(3,2);q)=q1-Nn0q-nN(q1-N)n.

Upon comparing (1.1) and (4.1), we immediately observe that F(q) matches the colored Jones polynomial for T(3,2) at a root of unity q=ζN:=e2πiN, that is, ζNF(ζN)=JN(T(3,2);ζN).

Consider the family of torus knots T(3,2t) for an integer t2. In this case, a q-hypergeometric expression for the colored Jones polynomial has been computed, namely (see [21, page 41, Théorème 3.2], cf. [18])

(4.2)JN(T(3,2t);q)=(-1)h′′(t)q2t-1-h(t)-Nn0(q1-N)nq-Nnm(t)×3=1m(t)-1j1(modm(t))(-q-N)=1m(t)-1jq-a(t)+=1m(t)-1jm(t)+=1m(t)-1(j2)×k=0m(t)-1q-kN=1m(t)-1[n+I(k)j],

where

h′′(t)={2t-13iftis even,2t-23iftis odd,  h(t)={2t-43iftis even,2t-53iftis odd,  a(t)={2t-1+13iftis even,2t+13iftis odd,

m(t)=2t-1, I(*) is the characteristic function and

[nk]=[nk]q:=(q)n(q)n-k(q)k

is the q-binomial coefficient. We now define the Kontsevich–Zagier series for torus knots T(3,2t) as[1]

(4.3)t(q)=(-1)h′′(t)q-h(t)n0(q)n3=1m(t)-1j1(modm(t))(-1)=1m(t)-1jq-a(t)+=1m(t)-1jm(t)+=1m(t)-1(j2)×k=0m(t)-1=1m(t)-1[n+I(k)j].

The expression t(q) converges in a similar manner as F(q) and, by (4.2) and (4.3), satisfies

ζN2t-1t(ζN)=JN(T(3,2t);ζN).

An application of Theorem 1.1 is the following. For an integer t2, set st:=(2t+1-3)232t+2.

Corollary 4.1.

For an integer t2 and αQ, ϕt(α):=e2πistαFt(e2πiα) is a quantum modular form of weight 32 on A32t+1={αQ:α𝑖𝑠Γ1(32t+2)-equivalent to 0𝑜𝑟i} with respect to Γ1(32t+2).

Proof.

The Kontsevich–Zagier series t(q) satisfies the “strange” identity (see [2, Proposition 2.4])[2]

(4.4)t(q)=-12Θχt(z),

where

(4.5)χt(n):={1ifn2t+1-3, 3+2t+2(mod32t+1),-1ifn2t+1+3, 2t+2-3(mod32t+1),0otherwise.

Note that χt is an even function with period M=32t+1. For k0=(2t+1-3)2(mod32t+2), consider the set χt(k0)={2t+1-3,2t+1+3}. Thus, Sχt(k0)={±(2t+1-3),±(2t+1+3)}. By Theorem 1.1 and (4.4), the result follows. ∎

4.2 Generating function for odd balanced unimodal sequences

Let v(n) denote the number of odd-balanced unimodal sequences of weight 2n+2 and v(m,n) the number of such sequences having rank m. In [20], the authors study the bivariate generating function

𝒱(x,q):=n0(-xq,-xq)nqn(q,q2)n+1=n0mv(m,n)xmqn

and prove that, for α, q-7𝒱(-1,q-8)|zα is a quantum modular form of weight 32 on

A={α:αisΓ0(16)-equivalent toi}

with respect to Γ0(16). A slight variant of this result is as follows. If we let qq2 in the identity (see [20, page 3693])

𝒱(-1,q-1)=-q2n0(2n+1)qn(n+1)2,

then

q-74𝒱(-1,q-2)=-12n0nψ(n)qn24,

where ψ(n) is the (non-primitive) Dirichlet character modulo 2 which is 0 or 1 according as n is even or odd. Note that ψ(n) is even with period 2 and Sψ(k0)=ψ(k0)={1}, where k0=1. For α, it follows from Theorem 1.1 that Θψ(α):=e-7πiα2𝒱(-1,e-4πiα) is a quantum modular form of weight 32 on

A2={α:αisΓ1(4)-equivalent toi}

with respect to Γ1(4). Precisely, Θψ(α) satisfies

Θψ(α)-(Θψ|32,χγ)(α)=rγ,ψ(α)

for all γ=(abcd)Γ1(4) and αA2, and where

rγ,ψ(z)=eπi422πγ-1(i)iθψ(τ)(τ-z¯)-32dτ.

Here, rγ,ψ: is a C function which is real-analytic in {γ-1(i)}, and χ is a multiplier given by

χ(γ)=eπiab2(4cd)εd-1.

4.3 Kontsevich–Zagier series for torus knots T(2,2m+1)

Let m. For 0m-1, define the Kontsevich–Zagier series for the torus knot T(2,2m+1) as follows:

Xm()(q):=k1,k2,,km=0(q)kmqk12++km-12+k+1++km-1i=1m-1[ki+1+δi,ki],

where δi, is the characteristic function. Hikami [16] established the strange identity

(4.6)Xm()(q)=-12n=0nχ8m+4()(n)qn2-(2m-2-1)28(2m+1),

where

χ8m+4()(n):={1ifn2m-2-1, 6m+2+5(mod8m+4),-1ifn2m+2+3, 6m-2+1(mod8m+4),0otherwise.

Note that f(n):=χ8m+4()(n) is an even function with period 8m+4. For k0=(2m-2-1)2(mod16m+8), consider the set f(k0)={2m-2-1,2m+2+3}. Thus, we have Sf(k0)={±(2m-2-1),±(2m+2+3)}. Observe that kf(k0)f(k)=0. Thus, for α, Theorem 1.1 and (4.6) imply that

eπiα(2m-2-1)28m+4Xm()(e2πiα)=-12Θf(α)=:Θ~m,(α)

is a quantum modular form of weight 32 on A8m+4={α:αisΓ1(16m+8)-equivalent to 0ori} with respect to Γ1(16m+8). Precisely, we have

Θ~m,(α)-(Θ~m,|32,χγ)(α)=rγ,f(α)

for all γ=(abcd)Γ1(16m+8) and αA8m+4, where

rγ,f(z)=8m+4eπi44πγ-1(i)iθf(τ)(τ-z¯)-32dτ.

Here, rγ,f: is a C function which is real-analytic in {γ-1(i)}, and χ is a multiplier given by

χ(γ)=eπiab(2m-2-1)2(8m+4)(2c(8m+4)d)εd-1.

We remark that Hikami proved Θ~m,(z) is a vector-valued quantum modular form of weight 32 on SL2() (see [16, page 195] for details).

4.4 Rogers’ false theta function

For M and 1j<M with jM2, consider the false theta function of Rogers

Fj,M(z):=nj(modM)sgn(n)qn22M=(n>0nj(modM)-n>0n-j(modM))qn22M=n>0f(n)qn22M,

where f(n) is the function defined by 1 or -1 according as nj or -j(modM) and 0 otherwise. Note that FM2,M(z)=0. Here, f is an odd function with period M. In this case, (k0)={j} (respectively, (k0)={M-j}) for 1j<M2 (respectively, M2<j<M) with k0=j2(mod2M) (respectively, k0=(M-j)2(mod2M)). So Sf(k0)={j,M-j}. Thus, for α, Theorem 1.1 implies that Fj,M(α) is a strong quantum modular form of weight 12 on with respect to ΓM (given by (1.4)). This result (with z replaced by zM and M even) was discussed in [6, Theorem 4.1] (see [7] for a vector-valued version). More generally, for 1k0<2M, if

FM(z):=n>0h(n)qn22M=jh(k0)h(j)Fj,M(z),

where h(n) is an odd function with period M and support Sh(k0), then Theorem 1.1 shows that FM(z) is a strong quantum modular form of weight 12 on with respect to ΓM. Finally, Fj,M(α) and, more generally, FM(α) are quantum modular forms of weight 12 on BM with respect to ΓM.


Communicated by Jan Bruinier


Funding source: Austrian Science Fund

Award Identifier / Grant number: SFB F50-06

Funding statement: The first author is supported by grant SFB F50-06 of the Austrian Science Fund (FWF).

Acknowledgements

The second author would like to thank the Max-Planck-Institut für Mathematik for their support during the initial stages of this project, the Ireland Canada University Foundation for the James M. Flaherty Visiting Professorship award and McMaster University for their hospitality during his stay from May 17 to August 9, 2019. The second author also thanks Yingkun Li for a clarifying remark during a visit to TU Darmstadt on April 23, 2019. Finally, the authors thank Jeremy Lovejoy for insightful comments on a preliminary version of this paper, Sergei Gukov for kindly reminding us of [14] and the referee for helpful comments and suggestions.

References

[1] S. Bettin and S. Drappeau, Modularity and value distribution of quantum invariants of hyperbolic knots, preprint (2019), https://arxiv.org/abs/1905.02045. 10.1007/s00208-021-02288-2Search in Google Scholar

[2] C. Bijaoui, H. U. Boden, B. Myers, R. Osburn, W. Rushworth, A. Tronsgard and S. Zhou, Generalized Fishburn numbers and torus knots, J. Combin. Theory Ser. A 178 (2021), Article ID 105355. 10.1016/j.jcta.2020.105355Search in Google Scholar

[3] K. Bringmann, T. Creutzig and L. Rolen, Negative index Jacobi forms and quantum modular forms, Res. Math. Sci. 1 (2014), Article ID 11. 10.1186/s40687-014-0011-8Search in Google Scholar

[4] K. Bringmann, A. Folsom, K. Ono and L. Rolen, Harmonic Maass Forms and Mock Modular Forms: Theory and Applications, Amer. Math. Soc. Colloq. Publ. 64, American Mathematical Society, Providence, 2017. 10.1090/coll/064Search in Google Scholar

[5] K. Bringmann, K. Mahlburg and A. Milas, Quantum modular forms and plumbing graphs of 3-manifolds, J. Combin. Theory Ser. A 170 (2020), Article ID 105145. 10.1016/j.jcta.2019.105145Search in Google Scholar

[6] K. Bringmann and A. Milas, 𝒲-algebras, false theta functions and quantum modular forms, I, Int. Math. Res. Not. IMRN 2015 (2015), no. 21, 11351–11387. 10.1093/imrn/rnv033Search in Google Scholar

[7] K. Bringmann and C. Nazaroglu, A framework for modular properties of false theta functions, Res. Math. Sci. 6 (2019), no. 3, Paper No. 30. 10.1007/s40687-019-0192-2Search in Google Scholar

[8] K. Bringmann and L. Rolen, Half-integral weight Eichler integrals and quantum modular forms, J. Number Theory 161 (2016), 240–254. 10.1016/j.jnt.2015.03.001Search in Google Scholar

[9] R. Bruggeman, Quantum Maass forms, The Conference on L-Functions, World Scientific, Hackensack (2007), 1–15. 10.1142/9789812772398_0001Search in Google Scholar

[10] M. C. Cheng, S. Chun, F. Ferrari, S. Gukov and S. M. Harrison, 3d modularity, J. High Energy Phys. 2019 (2019), no. 10, Article ID 010. 10.1007/JHEP10(2019)010Search in Google Scholar

[11] M. C. Cheng, F. Ferrari and G. Sgroi, Three-manifold quantum invariants and mock theta functions, Philos. Trans. Roy. Soc. A 378 (2020), no. 2163, Article ID 20180439. 10.1098/rsta.2018.0439Search in Google Scholar

[12] A. Dabholkar, D. Jain and A. Rudra, APS η-invariant, path integrals, and mock modularity, J. High Energy Phys. 2019 (2019), no. 11, Article ID 080. 10.1007/JHEP11(2019)080Search in Google Scholar

[13] A. Folsom, M.-J. Jang, S. Kimport and H. Swisher, Quantum modular forms and singular combinatorial series with repeated roots of unity, Acta Arith. 194 (2020), no. 4, 393–421. 10.4064/aa190326-23-10Search in Google Scholar

[14] S. Gukov, Three-dimensional quantum gravity, Chern–Simons theory, and the A-polynomial, Comm. Math. Phys. 255 (2005), no. 3, 577–627. 10.1007/s00220-005-1312-ySearch in Google Scholar

[15] K. Habiro, On the colored Jones polynomials of some simple links, RIMS Kōkyūroku 1172 (2000), 34–43. Search in Google Scholar

[16] K. Hikami, q-series and L-functions related to half-derivatives of the Andrews–Gordon identity, Ramanujan J. 11 (2006), no. 2, 175–197. 10.1007/s11139-006-6506-1Search in Google Scholar

[17] K. Hikami, Quantum invariants, modular forms, and lattice points. II, J. Math. Phys. 47 (2006), no. 10, Article ID 102301. 10.1063/1.2349484Search in Google Scholar

[18] K. Hikami and A. N. Kirillov, Hypergeometric generating function of L-function, Slater’s identities, and quantum invariant, Algebra i Analiz 17 (2005), no. 1, 190–208. 10.1090/S1061-0022-06-00897-1Search in Google Scholar

[19] K. Hikami and J. Lovejoy, Hecke-type formulas for families of unified Witten–Reshetikhin–Turaev invariants, Commun. Number Theory Phys. 11 (2017), no. 2, 249–272. 10.4310/CNTP.2017.v11.n2.a1Search in Google Scholar

[20] B. Kim, S. Lim and J. Lovejoy, Odd-balanced unimodal sequences and related functions: Parity, mock modularity and quantum modularity, Proc. Amer. Math. Soc. 144 (2016), no. 9, 3687–3700. 10.1090/proc/13027Search in Google Scholar

[21] I. Konan, Autour des q,-séries, des formes modulaires quantiques et des nœuds toriques, Master’s thesis, Université Denis Diderot, Paris, https://www.irif.fr/_media/users/konan/memoire.pdf. Search in Google Scholar

[22] R. Lawrence and D. Zagier, Modular forms and quantum invariants of 3-manifolds, Asian J. Math. 3 (1999), no. 1, 93–107. 10.4310/AJM.1999.v3.n1.a5Search in Google Scholar

[23] T. T. Q. Lê, Quantum invariants of 3-manifolds: Integrality, splitting, and perturbative expansion, Topology Appl. 127 (2003), 125–152. 10.1016/S0166-8641(02)00056-1Search in Google Scholar

[24] J. Lewis and D. Zagier, Cotangent sums, quantum modular forms, and the generalized Riemann hypothesis, Res. Math. Sci. 6 (2019), no. 1, Paper No. 4. 10.1007/s40687-018-0159-8Search in Google Scholar

[25] H. Murakami and Y. Yokota, Volume Conjecture for Knots, SpringerBriefs Math. Phys. 30, Springer, Singapore, 2018. 10.1007/978-981-13-1150-5Search in Google Scholar

[26] A. Nordentoft, A note on additive twists, reciprocity laws and quantum modular forms, Ramanujan J. (2020), 10.1007/s11139-020-00270-1. 10.1007/s11139-020-00270-1Search in Google Scholar

[27] H. Rademacher, Topics in Analytic Number Theory, Springer, New York, 1973. 10.1007/978-3-642-80615-5Search in Google Scholar

[28] G. Shimura, On modular forms of half integral weight, Ann. of Math. (2) 97 (1973), 440–481. 10.2307/1970831Search in Google Scholar

[29] D. Zagier, Vassiliev invariants and a strange identity related to the Dedekind eta-function, Topology 40 (2001), no. 5, 945–960. 10.1016/S0040-9383(00)00005-7Search in Google Scholar

[30] D. Zagier, Quantum modular forms, Quanta of Maths, Clay Math. Proc. 11, American Mathematical Society, Providence (2010), 659–675. Search in Google Scholar

Received: 2020-07-28
Revised: 2020-12-03
Published Online: 2021-01-13
Published in Print: 2021-03-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/forum-2020-0201/html
Scroll to top button