This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Hikami’s observations on unified WRT invariants
and false theta functions

Toshiki Matsusaka Faculty of Mathematics, Kyushu University, Motooka 744, Nishi-ku, Fukuoka 819-0395, Japan [email protected]
Abstract.

The object of this article is a family of qq-series originating from Habiro’s work on the Witten–Reshetikhin–Turaev invariants. The qq-series usually make sense only when qq is a root of unity, but for some instances, it also determines a holomorphic function on the open unit disc. Such an example is Habiro’s unified WRT invariant H(q)H(q) for the Poincaré homology sphere. In 2007, Hikami observed its discontinuity at roots of unity. More precisely, the value of H(ζ)H(\zeta) at a root of unity is 1/21/2 times the limit value of H(q)H(q) as qq tends towards ζ\zeta radially within the unit disc. In this article, we explain the appearance of the 1/21/2-factor and generalize Hikami’s observations by using Bailey’s lemma and the theory of false theta functions.

2020 Mathematics Subject Classification:
Primary 11F27; Secondary 57K16

Dedicated to the memory of Toshie Takata.

1. Introduction

The WRT invariants are derived from the work of Witten [Witten1989] and Reshetikhin–Turaev [ReshetikhinTuraev1991]. Witten answered Atiyah’s question on a 33-dimensional definition of the Jones polynomials of knot theory and introduced certain invariants of 33-manifolds using quantum field theory. Its rigorous mathematical definition was subsequently given by Reshetikhin and Turaev using the quantum group Uq(sl2)U_{q}(sl_{2}) at roots of unity and has been extensively investigated.

Here is one example. The WRT-invariant τN(Σ(2,3,5))\tau_{N}(\Sigma(2,3,5)) associated to the Poincaré homology sphere M=Σ(2,3,5)M=\Sigma(2,3,5) is computed as

e2πiN121120(e2πiN1)τN(Σ(2,3,5))=eπi/4260N0n60N1Nneπin260Nj=13(eπinNpjeπinNpj)eπinNeπinN\displaystyle e^{\frac{2\pi i}{N}\frac{121}{120}}(e^{\frac{2\pi i}{N}}-1)\tau_{N}(\Sigma(2,3,5))=\frac{e^{\pi i/4}}{2\sqrt{60N}}\sum_{\begin{subarray}{c}0\leq n\leq 60N-1\\ N\nmid n\end{subarray}}e^{-\frac{\pi in^{2}}{60N}}\frac{\prod_{j=1}^{3}(e^{\frac{\pi in}{Np_{j}}}-e^{-\frac{\pi in}{Np_{j}}})}{e^{\frac{\pi in}{N}}-e^{-\frac{\pi in}{N}}}

for N>1N\in\mathbb{Z}_{>1}, where (p1,p2,p3)=(2,3,5)(p_{1},p_{2},p_{3})=(2,3,5) (see Lawrence–Rozansky [LawrenceRozansky1999] and Hikami [Hikami2005IJM]). One of the topics of research on the WRT invariants is to find a “unified” function that can capture the values for all NN. More precisely, we find a function τ(M):𝒵\tau(M):\mathcal{Z}\to\mathbb{C} defined on the set of all roots of unity 𝒵\mathcal{Z} such that the value τ(M)(e2πi/N)\tau(M)(e^{2\pi i/N}) coincides with the WRT invariant τN(M)\tau_{N}(M). A number-theoretic (or analytic) approach was given by Lawrence–Zagier [LawrenceZagier1999] using false theta functions. They considered the qq-series defined by

Φ~(2,3,5)(1,1,1)(τ)=12nsgn(n)χ(2,3,5)(1,1,1)(n)qn2120(q=e2πiτ),\widetilde{\Phi}_{(2,3,5)}^{(1,1,1)}(\tau)=\frac{1}{2}\sum_{n\in\mathbb{Z}}\operatorname{sgn}(n)\chi_{(2,3,5)}^{(1,1,1)}(n)q^{\frac{n^{2}}{120}}\qquad(q=e^{2\pi i\tau}),

where qr=e2πirτq^{r}=e^{2\pi ir\tau} for rr\in\mathbb{Q} and τ={τIm(τ)>0}\tau\in\mathbb{H}=\{\tau\in\mathbb{C}\mid\operatorname{Im}(\tau)>0\}, and

χ(2,3,5)(1,1,1)(n)={1if n31,41,49,59(mod60),1if n1,11,19,29(mod60),0if otherwise.\chi_{(2,3,5)}^{(1,1,1)}(n)=\begin{cases}1&\text{if }n\equiv 31,41,49,59\pmod{60},\\ -1&\text{if }n\equiv 1,11,19,29\pmod{60},\\ 0&\text{if otherwise}.\end{cases}

Then they showed that

limt0Φ~(2,3,5)(1,1,1)(1N+it)=160Nn=160Nnχ(2,3,5)(1,1,1)(n)eπin260N\lim_{t\to 0}\widetilde{\Phi}_{(2,3,5)}^{(1,1,1)}\left(\frac{1}{N}+it\right)=-\frac{1}{60N}\sum_{n=1}^{60N}n\chi_{(2,3,5)}^{(1,1,1)}(n)e^{\pi i\frac{n^{2}}{60N}}

and

eπi60N2limt0Φ~(2,3,5)(1,1,1)(1N+it)=1+e2πiN(1e2πiN)τN(Σ(2,3,5)).-\frac{e^{-\frac{\pi i}{60N}}}{2}\lim_{t\to 0}\widetilde{\Phi}_{(2,3,5)}^{(1,1,1)}\left(\frac{1}{N}+it\right)=1+e^{\frac{2\pi i}{N}}(1-e^{\frac{2\pi i}{N}})\tau_{N}(\Sigma(2,3,5)).

In this sense, the qq-series Φ~(2,3,5)(1,1,1)(τ)\widetilde{\Phi}_{(2,3,5)}^{(1,1,1)}(\tau) unifies the WRT-invariants via the limits to the roots of unity.

Another approach is developed by Habiro [Habiro2008]. Habiro constructed the unified WRT invariant IM(q)I_{M}(q) for the integral homology spheres MM with values in the set so-called “Habiro ring” today. For instance, the unified WRT invariant IΣ(2,3,5)(q)I_{\Sigma(2,3,5)}(q) he constructed is given by

(1.1) H(q):=1+q(1q)IΣ(2,3,5)(q)=n=0qn(qn)n,\displaystyle H(q):=1+q(1-q)I_{\Sigma(2,3,5)}(q)=\sum_{n=0}^{\infty}q^{n}(q^{n})_{n},

where (x)n=(x;q)n=k=0n1(1xqk)(x)_{n}=(x;q)_{n}=\prod_{k=0}^{n-1}(1-xq^{k}) is the usual qq-Pochhammer symbol. A characteristic of this type of series expression is that although an infinite sum defines it, substituting roots of unity for qq truncates the sum to a finite sum. Series with such properties were observed before Habiro. A few famous examples are Kontsevich’s function F(q)=n=0(q)nF(q)=\sum_{n=0}^{\infty}(q)_{n} studied in [Zagier2001] and Ramanujan’s function σ(q)=1+n=1(1)n1qn(q)n1\sigma(q)=1+\sum_{n=1}^{\infty}(-1)^{n-1}q^{n}(q)_{n-1} discovered by Andrews [Andrews1986] and studied in [AndrewsDysonHickerson1988, Cohen1988]. In this case, IΣ(2,3,5)(e2πi/N)=τN(Σ(2,3,5))I_{\Sigma(2,3,5)}(e^{2\pi i/N})=\tau_{N}(\Sigma(2,3,5)), that is,

H(ζ)=n=0ζn(ζn)n=1+ζ(1ζ)τN(Σ(2,3,5))H(\zeta)=\sum_{n=0}^{\infty}\zeta^{n}(\zeta^{n})_{n}=1+\zeta(1-\zeta)\tau_{N}(\Sigma(2,3,5))

holds for ζ=e2πi/N\zeta=e^{2\pi i/N}.

Now we have two ways to unify the WRT invariants. Is there any direct relationship between them? First, it is worth noting that, by the term qnq^{n} in the sum, Habiro’s series in (1.1) can be viewed as an element in [[q]]\mathbb{Z}[[q]], which is a feature not found in Kontsevich’s function F(q)F(q). Then, Hikami [Hikami2007] addressed this question and succeeded in showing the direct equation

H(q)=q1120Φ~(2,3,5)(1,1,1)(τ)H(q)=-q^{-\frac{1}{120}}\widetilde{\Phi}_{(2,3,5)}^{(1,1,1)}(\tau)

as a holomorphic function on |q|<1|q|<1. However, we notice a strange phenomenon. By the above results, we see that

(1.2) H(e2πi/N)=12limτ=1/N+itt0q1120Φ~(2,3,5)(1,1,1)(τ)=12limq=e2πi/Nett0H(q).\displaystyle H(e^{2\pi i/N})=-\frac{1}{2}\lim_{\begin{subarray}{c}\tau=1/N+it\\ t\to 0\end{subarray}}q^{-\frac{1}{120}}\widetilde{\Phi}_{(2,3,5)}^{(1,1,1)}(\tau)=\frac{1}{2}\lim_{\begin{subarray}{c}q=e^{2\pi i/N}e^{-t}\\ t\to 0\end{subarray}}H(q).

The mystery of the 1/21/2-factor was pointed out by Habiro [Habiro2008, Section 16].

1.1. Main results

The article aims to generalize the relation between Habiro-type series and false theta functions studied by Hikami [Hikami2007] and provide a plausible explanation for the appearance of the 1/21/2-factor. First, we review Hikami’s results and observations.

For more general Brieskorn homology spheres Σ(2,3,6p1)\Sigma(2,3,6p-1), Hikami explicitly expressed Habiro’s unified WRT invariants as follows. For any integer p>1p>1, we have

(1.3) Hp(1)(q):=(1q)IΣ(2,3,6p1)(q)=sps10qsp(qsp+1)sp+1i=1p1qsi(si+1)[si+1si]q,\displaystyle H_{p}^{(1)}(q):=(1-q)I_{\Sigma(2,3,6p-1)}(q)=\sum_{s_{p}\geq\cdots\geq s_{1}\geq 0}q^{s_{p}}(q^{s_{p}+1})_{s_{p}+1}\prod_{i=1}^{p-1}q^{s_{i}(s_{i}+1)}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q},

where []q\genfrac{[}{]}{0.0pt}{}{\cdot}{\cdot}_{q} is the qq-binomial coefficient defined by

[nk]q=(q)n(q)k(q)nk.\genfrac{[}{]}{0.0pt}{}{n}{k}_{q}=\frac{(q)_{n}}{(q)_{k}(q)_{n-k}}.

Then substituting q=e2πi/Nq=e^{2\pi i/N} truncates the infinite sum defining the unified WRT invariant to a finite sum and IΣ(2,3,6p1)(e2πi/N)=τN(Σ(2,3,6p1))I_{\Sigma(2,3,6p-1)}(e^{2\pi i/N})=\tau_{N}(\Sigma(2,3,6p-1)) holds. On the other hand, Hikami [Hikami2005IJM] generalized Lawrence–Zagier’s qq-series as

Φ~(p1,p2,p3)(1,2,3)(τ)\displaystyle\widetilde{\Phi}_{(p_{1},p_{2},p_{3})}^{(\ell_{1},\ell_{2},\ell_{3})}(\tau) =12nsgn(n)χ(p1,p2,p3)(1,2,3)(n)qn24p1p2p3\displaystyle=\frac{1}{2}\sum_{n\in\mathbb{Z}}\operatorname{sgn}(n)\chi_{(p_{1},p_{2},p_{3})}^{(\ell_{1},\ell_{2},\ell_{3})}(n)q^{\frac{n^{2}}{4p_{1}p_{2}p_{3}}}

with a periodic function χ(p1,p2,p3)(1,2,3):/2p1p2p3{1,0,1}\chi_{(p_{1},p_{2},p_{3})}^{(\ell_{1},\ell_{2},\ell_{3})}:\mathbb{Z}/2p_{1}p_{2}p_{3}\mathbb{Z}\to\{-1,0,1\}, which we define later in (3.1). Then he showed that

(1.4) 12limτ1/Nq(6p+5)224(6p1)Φ~(2,3,6p1)(1,1,1)(τ)=(1e2πi/N)τN(Σ(2,3,6p1))=Hp(1)(e2πi/N)\displaystyle\begin{split}-\frac{1}{2}\lim_{\tau\to 1/N}q^{-\frac{(6p+5)^{2}}{24(6p-1)}}\widetilde{\Phi}_{(2,3,6p-1)}^{(1,1,1)}(\tau)&=(1-e^{2\pi i/N})\tau_{N}(\Sigma(2,3,6p-1))\\ &=H_{p}^{(1)}(e^{2\pi i/N})\end{split}

for any p>1p>1. Here the limit is along the vertical line τ=1/N+it\tau=1/N+it as before. Similarly below, we will consider the vertical limit τ=1/N+it1/N\tau=1/N+it\to 1/N or the radial limit q=e2πi/Nete2πi/Nq=e^{2\pi i/N}e^{-t}\to e^{2\pi i/N} as limits. To observe a similarity to (1.2), we are interested in comparing the limit

limqe2πi/NHp(1)(q)\lim_{q\to e^{2\pi i/N}}H_{p}^{(1)}(q)

from within the unit disc |q|<1|q|<1 and the value Hp(1)(e2πi/N)H_{p}^{(1)}(e^{2\pi i/N}) given in (1.4). In this case, however, numerical calculations show that the difference is no longer a constant multiple. More specifically, when qq tends to a root of unity from within the unit disc, we observe a divergence of Hp(1)(q)H_{p}^{(1)}(q). Our main theorem claims that the “convergent part” of Hp(1)(q)H_{p}^{(1)}(q) converges to the value in (1.4).

Theorem 1.1 (The precise statement is given in Theorem 3.15 and Theorem 3.21).

For any integer p>1p>1, as a holomorphic function on the open unit disc, the series has the expression

Hp(1)(q)=q(6p+5)224(6p1)2η(τ)𝜺=(ε1ε2){0,1}2(1)ε1+ε2(θ~𝝁𝟏+𝜺,𝒄𝟏(2)(τ)+θ~𝝁𝟏+𝜺,𝒄𝟐(2)(τ))H_{p}^{(1)}(q)=\frac{q^{-\frac{(6p+5)^{2}}{24(6p-1)}}}{2\eta(\tau)}\sum_{\bm{\varepsilon}=\bigl{(}\begin{smallmatrix}\varepsilon_{1}\\ \varepsilon_{2}\end{smallmatrix}\bigr{)}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\bigg{(}\widetilde{\theta}_{\bm{\mu_{1}}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\theta}_{\bm{\mu_{1}}+\bm{\varepsilon},\bm{c_{2}}}^{(2)}(\tau)\bigg{)}

in terms of false theta functions θ~𝛍+𝛆,𝐜(2)(τ)\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c}}^{(2)}(\tau) defined in (3.2) and the Dedekind eta function η(τ)=q1/24(q)\eta(\tau)=q^{1/24}(q)_{\infty}. Then the first half converges in the vertical limit τ1/N\tau\to 1/N to

limτ1/Nq(6p+5)224(6p1)2η(τ)𝜺{0,1}2(1)ε1+ε2θ~𝝁𝟏+𝜺,𝒄𝟏(2)(τ)=12limτ1/Nq(6p+5)224(6p1)Φ~(2,3,6p1)(1,1,1)(τ),\lim_{\tau\to 1/N}\frac{q^{-\frac{(6p+5)^{2}}{24(6p-1)}}}{2\eta(\tau)}\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\widetilde{\theta}_{\bm{\mu_{1}}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)=-\frac{1}{2}\lim_{\tau\to 1/N}q^{-\frac{(6p+5)^{2}}{24(6p-1)}}\widetilde{\Phi}_{(2,3,6p-1)}^{(1,1,1)}(\tau),

which coincides with the value of Hp(1)(q)H_{p}^{(1)}(q) at q=e2πi/Nq=e^{2\pi i/N}. The second half of the expression diverges in the same limit generally.

As for p=1p=1, the function H(q)H(q) given in (1.1) is denoted by H1(2)(q)H_{1}^{(2)}(q) in the following general notations. The series also has a similar expression

H1(2)(q)=n=0qn(qn)n=q11202η(τ)𝜺{0,1}2(1)ε1+ε2(θ~𝝁𝟐+𝜺,𝒄𝟏(2)(τ)+θ~𝝁𝟐+𝜺,𝒄𝟐(2)(τ)),H_{1}^{(2)}(q)=\sum_{n=0}^{\infty}q^{n}(q^{n})_{n}=\frac{q^{-\frac{1}{120}}}{2\eta(\tau)}\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\bigg{(}\widetilde{\theta}_{\bm{\mu_{2}}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\theta}_{\bm{\mu_{2}}+\bm{\varepsilon},\bm{c_{2}}}^{(2)}(\tau)\bigg{)},

and the first half converges to the value H1(2)(e2πi/N)H_{1}^{(2)}(e^{2\pi i/N}) in the limit τ1/N\tau\to 1/N. Furthermore, in this case, the first and second terms accidentally coincide. This fact follows from the symmetry of aa and bb in the expression given in Theorem 2.15. Thus the limit of the whole H1(2)(q)H_{1}^{(2)}(q) also converges, and its limit equals 2H1(2)(e2πi/N)2H_{1}^{(2)}(e^{2\pi i/N}). That is a reason for the occurrence of the 1/21/2-factor in (1.2).

Hikami [Hikami2007] also gave many observations on the relations between other Habiro-type series and the limits of Φ~(2,3,6p1)(1,2,3)(τ)\widetilde{\Phi}_{(2,3,6p-1)}^{(\ell_{1},\ell_{2},\ell_{3})}(\tau). More precisely, he introduced another infinite family of Habiro-type series Hp(5)(q)H_{p}^{(5)}(q) and three more examples H2(2)(q),H2(3)(q)H_{2}^{(2)}(q),H_{2}^{(3)}(q), and H2(4)(q)H_{2}^{(4)}(q) in the following notations. Here we generalize Hikami’s examples to five infinite families.

Definition 1.2.

For any positive integer p1p\geq 1, we define five Habiro-type series by

Hp(1)(q)\displaystyle H_{p}^{(1)}(q) =sps10qsp(qsp+1)sp+1i=1p1qsi(si+1)[si+1si]q,\displaystyle=\sum_{s_{p}\geq\cdots\geq s_{1}\geq 0}q^{s_{p}}(q^{s_{p}+1})_{s_{p}+1}\prod_{i=1}^{p-1}q^{s_{i}(s_{i}+1)}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q},
Hp(2)(q)\displaystyle H_{p}^{(2)}(q) =sps10qsp(qsp)spi=1p1qsi2[si+1si]q,\displaystyle=\sum_{s_{p}\geq\cdots\geq s_{1}\geq 0}q^{s_{p}}(q^{s_{p}})_{s_{p}}\prod_{i=1}^{p-1}q^{s_{i}^{2}}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q},
Hp(3)(q)\displaystyle H_{p}^{(3)}(q) =sps10q2sp(qsp+1)spi=1p1qsi(si+1)[si+1si]q,\displaystyle=\sum_{s_{p}\geq\cdots\geq s_{1}\geq 0}q^{2s_{p}}(q^{s_{p}+1})_{s_{p}}\prod_{i=1}^{p-1}q^{s_{i}(s_{i}+1)}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q},
Hp(4)(q)\displaystyle H_{p}^{(4)}(q) =sps10qsp(qsp+1)spi=1p1qsi(si+1)[si+1si]q,\displaystyle=\sum_{s_{p}\geq\cdots\geq s_{1}\geq 0}q^{s_{p}}(q^{s_{p}+1})_{s_{p}}\prod_{i=1}^{p-1}q^{s_{i}(s_{i}+1)}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q},
Hp(5)(q)\displaystyle H_{p}^{(5)}(q) =sps10qsp(qsp+1)spi=1p1qsi2[si+1si]q.\displaystyle=\sum_{s_{p}\geq\cdots\geq s_{1}\geq 0}q^{s_{p}}(q^{s_{p}+1})_{s_{p}}\prod_{i=1}^{p-1}q^{s_{i}^{2}}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q}.

If the notations are to match those adapted by the spirit of Hikami [Hikami2007], then the above series should be named Hp(1)(q)=Mp(1)(q)H_{p}^{(1)}(q)=M_{p}^{(1)}(q), Hp(2)(q)=Mp(p)(q)H_{p}^{(2)}(q)=M_{p}^{(p)}(q), Hp(3)(q)=Mp(2p1)(q)H_{p}^{(3)}(q)=M_{p}^{(2p-1)}(q), Hp(4)(q)=Mp(2p)(q)H_{p}^{(4)}(q)=M_{p}^{(2p)}(q), and Hp(5)(q)=Mp(3p1)(q)H_{p}^{(5)}(q)=M_{p}^{(3p-1)}(q). However, since the superscripts overlap when p=1p=1, the notations here are purposely changed. These five series are infinite families that extend each of Hikami’s M2(k)(q)M_{2}^{(k)}(q) for k=1,2,3,4,5k=1,2,3,4,5. Moreover, H1(2)(q)=M1(1)(q)H_{1}^{(2)}(q)=M_{1}^{(1)}(q) and H1(4)(q)=H1(5)(q)=M1(2)(q)H_{1}^{(4)}(q)=H_{1}^{(5)}(q)=M_{1}^{(2)}(q) hold in Hikami’s notations.

Our main theorems stated in Theorem 3.15 and Theorem 3.21 give similar expressions in terms of false theta functions and limit formulas of these five families as in Theorem 1.1. For instance, we have

limτ1/Nconvergent part of H2(2)(q)\displaystyle\lim_{\tau\to 1/N}\text{convergent part of }H_{2}^{(2)}(q) =12limτ1/Nq1264Φ~(2,3,11)(1,1,2)(τ)\displaystyle=-\frac{1}{2}\lim_{\tau\to 1/N}q^{-\frac{1}{264}}\widetilde{\Phi}_{(2,3,11)}^{(1,1,2)}(\tau)
=e2πi264N264Nn=1132Nnχ(2,3,11)(1,1,2)(n)eπin2132N,\displaystyle=\frac{e^{-\frac{2\pi i}{264N}}}{264N}\sum_{n=1}^{132N}n\chi_{(2,3,11)}^{(1,1,2)}(n)e^{\pi i\frac{n^{2}}{132N}},

where

χ(2,3,11)(1,1,2)(n)={1if n67,89,109,131(mod132),1if n1,23,43,65(mod132),0if otherwise.\chi_{(2,3,11)}^{(1,1,2)}(n)=\begin{cases}1&\text{if }n\equiv 67,89,109,131\pmod{132},\\ -1&\text{if }n\equiv 1,23,43,65\pmod{132},\\ 0&\text{if otherwise}.\end{cases}

Moreover, numerical calculations suggest that the above limit value coincides with the value H2(2)(e2πi/N)H_{2}^{(2)}(e^{2\pi i/N}), that is,

(1.5) H2(2)(e2πi/N)=e2πi264N264Nn=1132Nnχ(2,3,11)(1,1,2)(n)eπin2132N\displaystyle H_{2}^{(2)}(e^{2\pi i/N})=\frac{e^{-\frac{2\pi i}{264N}}}{264N}\sum_{n=1}^{132N}n\chi_{(2,3,11)}^{(1,1,2)}(n)e^{\pi i\frac{n^{2}}{132N}}

holds. The similarity with Theorem 1.1 leads us to expect the coincidence to hold, but it is a conjecture. For other cases, too, Hikami [Hikami2007, Conjectures 1–3] conjectured the coincidence between the limits of Φ~(2,3,6p1)(1,2,3)(τ)\widetilde{\Phi}_{(2,3,6p-1)}^{(\ell_{1},\ell_{2},\ell_{3})}(\tau) and the values of Habiro-type series through numerical calculations, but they are still open problems.

To conclude this introduction section, we introduce some related studies. First, Hikami also studied the unified WRT invariants for the Brieskorn homology spheres Σ(2,3,6p+1)\Sigma(2,3,6p+1) with p1p\geq 1, but we do not deal with the cases in this article. Second, the above conjecture for Hp(1)(q)H_{p}^{(1)}(q),

(limτ1/Nconvergent part of Hp(1)(q))=(the value of Hp(1)(q) at q=e2πi/N)\bigg{(}\lim_{\tau\to 1/N}\text{convergent part of }H_{p}^{(1)}(q)\bigg{)}=\bigg{(}\text{the value of }H_{p}^{(1)}(q)\text{ at }q=e^{2\pi i/N}\bigg{)}

proved in Theorem 1.1, is derived from the fact that both sides have the topological interpretations (1.3) and (1.4), namely begin the WRT invariants. On the other hand, the Habiro-type series defined in Definition 1.2 are found by numerical experiments so that the analogy of Theorem 1.1 holds, and so far, its roles in the theory of WRT invariants are unclear. Third, many other known methods exist to unify the WRT invariants for more general 33-manifolds. For instance, Hikami [Hikami2006JMP] further generalized Lawrence–Zagier’s series Φ~𝒑(τ)\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau) for the Seifert fibered homology 33-spheres with nn-singular fibers. More recently, Gukov–Pei–Putrov–Vafa [GPPV2020] introduced qq-series called homological blocks for any plumbed 33-manifolds associated with negative definite plumbing tree graphs based on Gukov–Putrov–Vafa [GukovPutrovVafa2017]. Andersen–Mistegård [AM2022] and Fuji–Iwaki–H. Murakami–Terashima [FIMT2021] independently studied the limit of the homological blocks at roots of unity in different contexts and showed that the homological blocks also unify the WRT invariants for Seifert fibered integral homology 33-spheres. As for other manifolds, Mori–Y. Murakami [MoriMurakami2022] dealt with the case for the H\mathrm{H}-graph, and Y. Murakami [Murakami2022+] extended it to more general cases. Furthermore, the modular transformation theory for the homological blocks is developing by Bringmann–Mahlburg–Milas [BMM2020], Bringmann–Kaszian–Milas–Nazaroglu [BKMN2023], and Matsusaka–Terashima [MatsusakaTerashima2021] et al.

This article is organized as follows. In Section 2, we give Hecke-type series expressions of the five families of the Habiro-type series. This expression yields the relation between the Habiro-type series and the false theta functions. Since the key to the proof is Bailey’s work on the Rogers–Ramanujan identities, we begin by reviewing it in the first half of Section 2. In Section 3, we introduce the notion of the false theta functions based on the recent work of Bringmann–Nazaroglu [BringmannNazaroglu2019]. Then, under this setting, we review Hikami’s work [Hikami2005IJM] on the function Φ~𝒑(τ)\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau). In Section 3.2, we give our first main theorem (Theorem 3.15) on the expressions of the Habiro-type series in terms of the false theta functions θ~𝝁,𝒄(2)(τ)\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau). The transformation called “false theta decomposition” decomposes the false theta functions into a sum of products of the (lower-dimensional) false theta functions θ~M,μ(1)(τ)\widetilde{\theta}_{M,\mu}^{(1)}(\tau) and the ordinary theta functions θM,μ(1)(τ)\theta_{M,\mu}^{(1)}(\tau) (Theorem 3.12). This decomposition allows us to calculate the limit of Habiro-type series at roots of unity and obtain our second main theorem (Theorem 3.21). Finally, we revisit Hikami’s question on the modular transformation theory of the Hecke-type series related to the Habiro-type series.

2. Hecke-type formulas

This section aims to transform the five Habiro-type series defined in Definition 1.2 into a Hecke-type series. Since the basic idea is based on Bailey’s lemma, developed by Andrews, we review it first. Then, as an application of Bailey’s lemma, we show five critical identities related to the Habiro-type series in Proposition 2.8 and a series of lemmas. In Section 2.3, we derive the desired Hecke-type expressions. Finally, in Section 2.4, although off-topic, we remark on a well-known equation of multiple zeta values derived from Bailey’s transform.

2.1. Bailey’s lemma

Bailey’s lemma has a long history, dating back to Bailey’s work [Bailey1947, Bailey1948] in the 1940s, which clarifies the structure of Rogers’ second proof of the Rogers–Ramanujan identities. The original idea of Bailey is simple but has several powerful applications. For example, Andrews [Andrews1986TAMS] found Hecke-type formulas of Ramanujan’s mock theta functions by constructing particular Bailey pairs. This discovery by Andrews led Zwegers [Zwegers2002] to establish the modular transformation theory of mock theta functions. In this subsection, we recall the claims and ideas of Bailey’s transform and Bailey’s lemma. Its more detailed and extensive history can be found in Andrews [Andrews1986AMS], Warnaar [Warnaar1999], and Sills [Sills2018].

Lemma 2.1 (Bailey’s transform).

If sequences (αn)n,(βn)n,(γn)n,(δn)n,(un)n(\alpha_{n})_{n},(\beta_{n})_{n},(\gamma_{n})_{n},(\delta_{n})_{n},(u_{n})_{n}, and (vn)n(v_{n})_{n} satisfy suitable convergence conditions and the equations

βn=k=0nαkunkvn+k,γn=k=nδkuknvk+n,\displaystyle\beta_{n}=\sum_{k=0}^{n}\alpha_{k}u_{n-k}v_{n+k},\qquad\gamma_{n}=\sum_{k=n}^{\infty}\delta_{k}u_{k-n}v_{k+n},

then we have

n=0αnγn=n=0βnδn.\sum_{n=0}^{\infty}\alpha_{n}\gamma_{n}=\sum_{n=0}^{\infty}\beta_{n}\delta_{n}.

The proof is simply an exchange of the order of the sums, where the “suitable convergence conditions” are required. In particular, let us choose un=1/(q)nu_{n}=1/(q)_{n} and vn=1/(aq)nv_{n}=1/(aq)_{n} with a complex number aa\in\mathbb{C}. Here (x)n=(x;q)n=k=0n1(1xqk)(x)_{n}=(x;q)_{n}=\prod_{k=0}^{n-1}(1-xq^{k}) is the usual qq-Pochhammer symbol with |q|<1|q|<1. Then the four sequences are required to satisfy the following equations.

(2.1) βn=k=0nαk(q)nk(aq)n+k,γn=k=nδk(q)kn(aq)k+n.\displaystyle\beta_{n}=\sum_{k=0}^{n}\frac{\alpha_{k}}{(q)_{n-k}(aq)_{n+k}},\qquad\gamma_{n}=\sum_{k=n}^{\infty}\frac{\delta_{k}}{(q)_{k-n}(aq)_{k+n}}.

A pair of sequences (α,β)(\alpha,\beta) satisfying the above first equation is called a Bailey pair relative to aa. Similarly, a pair (γ,δ)(\gamma,\delta) satisfying the second equation is called a conjugate Bailey pair relative to aa.

In applications, Bailey [Bailey1948, §.4] found the following conjugate Bailey pair (γ,δ)(\gamma,\delta).

Lemma 2.2.

For any ρ1,ρ2\rho_{1},\rho_{2}\in\mathbb{C} (such that no zeros appear in the denominators) and a non-negative integer N0N\geq 0, a pair of

γn\displaystyle\gamma_{n} =(aq/ρ1)N(aq/ρ2)N(aq)N(aq/ρ1ρ2)N(1)n(ρ1)n(ρ2)n(qN)n(aq/ρ1)n(aq/ρ2)n(aqN+1)n(aqρ1ρ2)nqnNn(n1)2,\displaystyle=\frac{(aq/\rho_{1})_{N}(aq/\rho_{2})_{N}}{(aq)_{N}(aq/\rho_{1}\rho_{2})_{N}}\frac{(-1)^{n}(\rho_{1})_{n}(\rho_{2})_{n}(q^{-N})_{n}}{(aq/\rho_{1})_{n}(aq/\rho_{2})_{n}(aq^{N+1})_{n}}\left(\frac{aq}{\rho_{1}\rho_{2}}\right)^{n}q^{nN-\frac{n(n-1)}{2}},
δn\displaystyle\delta_{n} =(ρ1)n(ρ2)n(qN)nqn(ρ1ρ2qN/a)n\displaystyle=\frac{(\rho_{1})_{n}(\rho_{2})_{n}(q^{-N})_{n}q^{n}}{(\rho_{1}\rho_{2}q^{-N}/a)_{n}}

is a conjugate Bailey pair relative to aa.

Proof.

The key to the proof is qq-analogue of the Saalschütz summation formula for the qq-hypergeometric series ϕ23{}_{3}\phi_{2}. The proof can be found in Andrews [Andrews1986AMS, p.25–27]. ∎

Since γn=δn=0\gamma_{n}=\delta_{n}=0 for n>Nn>N, the “suitable convergence conditions” required in Lemma 2.1 is satisfied. The following Bailey’s lemma tells us that a Bailey pair (α,β)(\alpha,\beta) yields a new Bailey pair (α,β)(\alpha^{\prime},\beta^{\prime}).

Theorem 2.3 (Bailey’s lemma).

If (α,β)(\alpha,\beta) is a Bailey pair relative to aa, then a pair of

αn\displaystyle\alpha^{\prime}_{n} =(ρ1)n(ρ2)n(aqρ1ρ2)n(aq/ρ1)n(aq/ρ2)nαn,\displaystyle=\frac{(\rho_{1})_{n}(\rho_{2})_{n}\left(\frac{aq}{\rho_{1}\rho_{2}}\right)^{n}}{(aq/\rho_{1})_{n}(aq/\rho_{2})_{n}}\alpha_{n},
βn\displaystyle\beta^{\prime}_{n} =j=0n(ρ1)j(ρ2)j(aq/ρ1ρ2)nj(aqρ1ρ2)j(q)nj(aq/ρ1)n(aq/ρ2)nβj\displaystyle=\sum_{j=0}^{n}\frac{(\rho_{1})_{j}(\rho_{2})_{j}(aq/\rho_{1}\rho_{2})_{n-j}\left(\frac{aq}{\rho_{1}\rho_{2}}\right)^{j}}{(q)_{n-j}(aq/\rho_{1})_{n}(aq/\rho_{2})_{n}}\beta_{j}

is also a Bailey pair relative to aa, that is,

βn=k=0nαk(q)nk(aq)n+k\displaystyle\beta^{\prime}_{n}=\sum_{k=0}^{n}\frac{\alpha^{\prime}_{k}}{(q)_{n-k}(aq)_{n+k}}

holds.

Proof.

We give a sketch of the proof given in [Andrews1986AMS, p.27]. A direct calculation yields

k=0Nαk(q)Nk(aq)N+k\displaystyle\sum_{k=0}^{N}\frac{\alpha^{\prime}_{k}}{(q)_{N-k}(aq)_{N+k}} =(aq/ρ1ρ2)N(aq/ρ1)N(aq/ρ2)N(q)Nk=0γkαk,\displaystyle=\frac{(aq/\rho_{1}\rho_{2})_{N}}{(aq/\rho_{1})_{N}(aq/\rho_{2})_{N}(q)_{N}}\sum_{k=0}^{\infty}\gamma_{k}\alpha_{k},

where γk\gamma_{k} is defined in Lemma 2.2 and γk=0\gamma_{k}=0 for k>Nk>N. By Lemma 2.1, the last sum equals k=0Nβkδk\sum_{k=0}^{N}\beta_{k}\delta_{k}. A more direct calculation yields the definition of βN\beta^{\prime}_{N}. ∎

For later applications, we will compute the particular case of Bailey’s lemma.

Corollary 2.4.

If (α,β)(\alpha,\beta) is a Bailey pair relative to aa, then a pair of

αn=anqn2αn,βn=j=0najqj2(q)njβj\displaystyle\alpha^{\prime}_{n}=a^{n}q^{n^{2}}\alpha_{n},\qquad\beta^{\prime}_{n}=\sum_{j=0}^{n}\frac{a^{j}q^{j^{2}}}{(q)_{n-j}}\beta_{j}

is also a Bailey pair relative to aa.

Proof.

In the definition of (αn,βn)(\alpha^{\prime}_{n},\beta^{\prime}_{n}), we take a limit as ρ1,ρ2\rho_{1},\rho_{2}\to\infty. ∎

Example 2.5.

We explain how Rogers–Ramanujan’s identities follow from Bailey’s lemma. The most basic example of Bailey pairs (relative to aa) is the unit Bailey pair defined by

(2.2) αn=(1aq2n)(a)n(1)nqn(n1)2(1a)(q)n,βn=δn,0={1if n=0,0if n>0\displaystyle\begin{split}\alpha_{n}&=\frac{(1-aq^{2n})(a)_{n}(-1)^{n}q^{\frac{n(n-1)}{2}}}{(1-a)(q)_{n}},\\ \beta_{n}&=\delta_{n,0}=\begin{cases}1&\text{if }n=0,\\ 0&\text{if }n>0\end{cases}\end{split}

(see Andrews [Andrews1984PJM, (2.12) and (2.13)]). First, we let a=1a=1. By applying Corollary 2.4 twice, we see that a pair of

(2.3) αn′′=q2n2αn={1if n=0,(1)nqn(5n1)2(1+qn)if n>0,βn′′=j=0nqj2(q)njk=0jqk2(q)jkβk=j=0nqj2(q)nj(q)j\displaystyle\begin{split}\alpha^{\prime\prime}_{n}&=q^{2n^{2}}\alpha_{n}=\begin{cases}1&\text{if }n=0,\\ (-1)^{n}q^{\frac{n(5n-1)}{2}}(1+q^{n})&\text{if }n>0,\end{cases}\\ \beta^{\prime\prime}_{n}&=\sum_{j=0}^{n}\frac{q^{j^{2}}}{(q)_{n-j}}\sum_{k=0}^{j}\frac{q^{k^{2}}}{(q)_{j-k}}\beta_{k}=\sum_{j=0}^{n}\frac{q^{j^{2}}}{(q)_{n-j}(q)_{j}}\end{split}

is also a Bailey pair relative to 11. By the first relation in (2.1) and taking a limit as nn\to\infty, we have

j=0qj2(q)j=1(q)k(1)kqk(5k1)2.\displaystyle\sum_{j=0}^{\infty}\frac{q^{j^{2}}}{(q)_{j}}=\frac{1}{(q)_{\infty}}\sum_{k\in\mathbb{Z}}(-1)^{k}q^{\frac{k(5k-1)}{2}}.

By Jacobi’s triple product

(2.4) n(1)nqn(n1)2ζn=(q)(ζ)(ζ1q),\displaystyle\sum_{n\in\mathbb{Z}}(-1)^{n}q^{\frac{n(n-1)}{2}}\zeta^{n}=(q)_{\infty}(\zeta)_{\infty}(\zeta^{-1}q)_{\infty},

we have

j=0qj2(q)j=(q5;q5)(q2;q5)(q3;q5)(q;q)=1(q;q5)(q4;q5),\sum_{j=0}^{\infty}\frac{q^{j^{2}}}{(q)_{j}}=\frac{(q^{5};q^{5})_{\infty}(q^{2};q^{5})_{\infty}(q^{3};q^{5})_{\infty}}{(q;q)_{\infty}}=\frac{1}{(q;q^{5})_{\infty}(q^{4};q^{5})_{\infty}},

which is so-called Rogers–Ramanujan’s first identity.

Similarly, we let a=qa=q in (2.2). Again, from the twice application of Corollary 2.4 and Jacobi’s triple product, we obtain the second Rogers–Ramanujan identity,

j=0qj(j+1)(q)j=1(q)k(1)kqk(5k+3)2=1(q2;q5)(q3;q5).\sum_{j=0}^{\infty}\frac{q^{j(j+1)}}{(q)_{j}}=\frac{1}{(q)_{\infty}}\sum_{k\in\mathbb{Z}}(-1)^{k}q^{\frac{k(5k+3)}{2}}=\frac{1}{(q^{2};q^{5})_{\infty}(q^{3};q^{5})_{\infty}}.

2.2. Bailey chains related to the Rogers–Ramanujan identities

As seen in Example 2.5, repeated application of Bailey’s lemma yields a sequence of Bailey pairs. We call the sequence a Bailey chain. To obtain Hecke-type expansions of the Habiro-type series defined in Definition 1.2, we recall two Bailey chains considered by Hikami [Hikami2007] and show three auxiliary lemmas below.

Let []q\genfrac{[}{]}{0.0pt}{}{\cdot}{\cdot}_{q} be the qq-binomial coefficient defined by

[nk]q=(q)n(q)k(q)nk.\genfrac{[}{]}{0.0pt}{}{n}{k}_{q}=\frac{(q)_{n}}{(q)_{k}(q)_{n-k}}.

The first Bailey chain follows from a unit Bailey pair with a=1a=1 considered in Example 2.5 related to the first Rogers–Ramanujan identity.

Proposition 2.6.

For any integer p1p\geq 1, a pair (α(1,p),β(1,p))(\alpha^{(1,p)},\beta^{(1,p)}) defined by

αn(1,p)\displaystyle\alpha_{n}^{(1,p)} =(1)nq(p+12)n212n(1+qn)δn,0,\displaystyle=(-1)^{n}q^{\left(p+\frac{1}{2}\right)n^{2}-\frac{1}{2}n}(1+q^{n})-\delta_{n,0},
βn(1,p)\displaystyle\beta_{n}^{(1,p)} =1(q)nn=sps10i=1p1qsi2[si+1si]q\displaystyle=\frac{1}{(q)_{n}}\sum_{n=s_{p}\geq\cdots\geq s_{1}\geq 0}\prod_{i=1}^{p-1}q^{s_{i}^{2}}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q}

is a Bailey pair relative to 11, where the empty product is understood as 11, that is, βn(1,1)=1/(q)n\beta_{n}^{(1,1)}=1/(q)_{n}.

Proof.

It follows from induction on pp. The initial case of p=1p=1 is given by applying Corollary 2.4 to the unit Bailey pair with a=1a=1. By the definition, we see that

αn(1,p)=qn2αn(1,p1),βn(1,p)=sp1=0nqsp12(q)nsp1βsp1(1,p1).\displaystyle\alpha_{n}^{(1,p)}=q^{n^{2}}\alpha_{n}^{(1,p-1)},\qquad\beta_{n}^{(1,p)}=\sum_{s_{p-1}=0}^{n}\frac{q^{s_{p-1}^{2}}}{(q)_{n-s_{p-1}}}\beta_{s_{p-1}}^{(1,p-1)}.

The induction assumption and Bailey’s lemma in Corollary 2.4 imply that (α(1,p),β(1,p))(\alpha^{(1,p)},\beta^{(1,p)}) is also a Bailey pair relative to 11. ∎

The following second Bailey chain is related to the second Rogers–Ramanujan identity.

Proposition 2.7.

For any integer p1p\geq 1, a pair (α(2,p),β(2,p))(\alpha^{(2,p)},\beta^{(2,p)}) defined by

αn(2,p)\displaystyle\alpha_{n}^{(2,p)} =(1)nq(p+12)n2+(p12)n1q2n+11q,\displaystyle=(-1)^{n}q^{\left(p+\frac{1}{2}\right)n^{2}+\left(p-\frac{1}{2}\right)n}\frac{1-q^{2n+1}}{1-q},
βn(2,p)\displaystyle\beta_{n}^{(2,p)} =1(q)nn=sps10i=1p1qsi(si+1)[si+1si]q\displaystyle=\frac{1}{(q)_{n}}\sum_{n=s_{p}\geq\cdots\geq s_{1}\geq 0}\prod_{i=1}^{p-1}q^{s_{i}(s_{i}+1)}\genfrac{[}{]}{0.0pt}{}{s_{i+1}}{s_{i}}_{q}

is a Bailey pair relative to qq.

Proof.

The proof is the same as that of Proposition 2.6. More precisely, the claim follows from repeatedly applying Corollary 2.4 to the unit Bailey pair with a=qa=q. ∎

Returning to the definition of Bailey pairs, the above two propositions are equivalent to the following equations.

Proposition 2.8.

For any integer p1p\geq 1, we have

βn(1,p)\displaystyle\beta_{n}^{(1,p)} =k=nn(1)kq(p+12)k212k(q)nk(q)n+k,\displaystyle=\sum_{k=-n}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}-\frac{1}{2}k}}{(q)_{n-k}(q)_{n+k}},
βn(2,p)\displaystyle\beta_{n}^{(2,p)} =k=n1n(1)kq(p+12)k2+(p12)k(q)nk(q)n+k+1.\displaystyle=\sum_{k=-n-1}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k+1}}.
Proof.

The first equation follows from a straightforward calculation. As for the second equation, we have

βn(2,p)=k=0nαk(2,p)(q)nk(q2)n+k=k=0n(1)kq(p+12)k2+(p12)k(q)nk(q)n+k+1(1q2k+1).\displaystyle\beta_{n}^{(2,p)}=\sum_{k=0}^{n}\frac{\alpha_{k}^{(2,p)}}{(q)_{n-k}(q^{2})_{n+k}}=\sum_{k=0}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k+1}}(1-q^{2k+1}).

We obtain the desired equation by dividing the sum into two parts and changing the variable. ∎

As can be immediately expected from the definitions of βn(1,p)\beta_{n}^{(1,p)} and βn(2,p)\beta_{n}^{(2,p)}, these are closely related to the Habiro-type series. For later calculations, we prepare three auxiliary lemmas.

Lemma 2.9.

For any integer n1n\geq 1, we have

(1qn)βn(1,p)=(1qn)k=nn(1)kq(p+12)k212k(q)nk(q)n+k=k=n+1n(1)kq(p+12)k212k(q)nk(q)n+k1.(1-q^{n})\beta_{n}^{(1,p)}=(1-q^{n})\sum_{k=-n}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}-\frac{1}{2}k}}{(q)_{n-k}(q)_{n+k}}=\sum_{k=-n+1}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}-\frac{1}{2}k}}{(q)_{n-k}(q)_{n+k-1}}.
Proof.

We put

an,k\displaystyle a_{n,k} =(1qn)(1)kq(p+12)k212k(q)nk(q)n+k(nkn),\displaystyle=(1-q^{n})\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}-\frac{1}{2}k}}{(q)_{n-k}(q)_{n+k}}\quad(-n\leq k\leq n),
bn,k\displaystyle b_{n,k} =(1)kq(p+12)k212k(q)nk(q)n+k1(n+1kn).\displaystyle=\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}-\frac{1}{2}k}}{(q)_{n-k}(q)_{n+k-1}}\quad(-n+1\leq k\leq n).

Then we can quickly check that an,n+an,n=bn,na_{n,-n}+a_{n,n}=b_{n,n} and an,k+an,k=bn,k+bn,ka_{n,-k}+a_{n,k}=b_{n,-k}+b_{n,k} for any 0kn10\leq k\leq n-1, which concludes the proof. ∎

Lemma 2.10.

For any integer n0n\geq 0, we have

qnβn(2,p)=qnk=n1n(1)kq(p+12)k2+(p12)k(q)nk(q)n+k+1=k=nn(1)kq(p+12)k2+(p+12)k(q)nk(q)n+k.q^{n}\beta_{n}^{(2,p)}=q^{n}\sum_{k=-n-1}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k+1}}=\sum_{k=-n}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p+\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k}}.
Proof.

We put

an,k\displaystyle a_{n,k} =qn(1)kq(p+12)k2+(p12)k(q)nk(q)n+k+1(n1kn),\displaystyle=q^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k+1}}\quad(-n-1\leq k\leq n),
bn,k\displaystyle b_{n,k} =(1)kq(p+12)k2+(p+12)k(q)nk(q)n+k(nkn).\displaystyle=\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p+\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k}}\quad(-n\leq k\leq n).

Again, we can also show by a direct calculation that an,n1+an,n=bn,na_{n,-n-1}+a_{n,n}=b_{n,n} and an,k1+an,k=bn,k1+bn,ka_{n,-k-1}+a_{n,k}=b_{n,-k-1}+b_{n,k} for any 0kn10\leq k\leq n-1. ∎

Lemma 2.11.

For any integer n1n\geq 1, we have

(1q2n)βn(2,p)=(1q2n)k=n1n(1)kq(p+12)k2+(p12)k(q)nk(q)n+k+1=k=n+1n(1)kq(p+12)k2+(p12)k(q)nk(q)n+k1.(1-q^{2n})\beta_{n}^{(2,p)}=(1-q^{2n})\sum_{k=-n-1}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k+1}}=\sum_{k=-n+1}^{n}\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k-1}}.
Proof.

As in the above two proofs, we put

an,k\displaystyle a_{n,k} =(1q2n)(1)kq(p+12)k2+(p12)k(q)nk(q)n+k+1(n1kn),\displaystyle=(1-q^{2n})\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k+1}}\quad(-n-1\leq k\leq n),
bn,k\displaystyle b_{n,k} =(1)kq(p+12)k2+(p12)k(q)nk(q)n+k1(n+1kn).\displaystyle=\frac{(-1)^{k}q^{\left(p+\frac{1}{2}\right)k^{2}+\left(p-\frac{1}{2}\right)k}}{(q)_{n-k}(q)_{n+k-1}}\quad(-n+1\leq k\leq n).

The situation becomes slightly different, but an,n1+an,n=bn,na_{n,-n-1}+a_{n,n}=b_{n,n}, an,n+an,n1=bn,n1a_{n,-n}+a_{n,n-1}=b_{n,n-1}, and an,k1+an,k=bn,k1+bn,ka_{n,-k-1}+a_{n,k}=b_{n,-k-1}+b_{n,k} (0kn2)(0\leq k\leq n-2) are still shown in the same way. ∎

2.3. Hecke-type expansions of Habiro-type series

Using the Bailey chains prepared in the previous subsection, we transform the five Habiro-type series into the Hecke-type series. Since Hikami [Hikami2007] has done the first and the fifth cases, we will review his proofs and then work on the remaining three. We note that for the remaining three, Hikami led to Hecke-type expansions only for the case p=2p=2. Our result extends Hikami’s results, but our proof method is slightly different from his.

The following equation is also used in the proof (see Hikami [Hikami2007, Lemma 3.6]).

Lemma 2.12.

For any non-negative integer c0c\geq 0, we have

m=0qm(q)2m+c(q)m+c(q)m=1(q)k=0(1)kq32k2+(c+32)k.\sum_{m=0}^{\infty}q^{m}\frac{(q)_{2m+c}}{(q)_{m+c}(q)_{m}}=\frac{1}{(q)_{\infty}}\sum_{k=0}^{\infty}(-1)^{k}q^{\frac{3}{2}k^{2}+\left(c+\frac{3}{2}\right)k}.
Proof.

In the equation given in [Fine1988, (25.96)],

m=0(aq)2m(bq)m(aq)m(q)mtm=(btq)(t)k=0(bq)k(t)k(q)k(btq)2k(at)kqk(3k+1)2,\sum_{m=0}^{\infty}\frac{(aq)_{2m}(bq)_{m}}{(aq)_{m}(q)_{m}}t^{m}=\frac{(btq)_{\infty}}{(t)_{\infty}}\sum_{k=0}^{\infty}\frac{(bq)_{k}(t)_{k}}{(q)_{k}(btq)_{2k}}(-at)^{k}q^{\frac{k(3k+1)}{2}},

we take a=qc,b=0a=q^{c},b=0 and t=qt=q. ∎

Theorem 2.13 (Hikami [Hikami2007, Theorem 3.5]).

For p1p\geq 1, we have the Hecke-type expansion for Hp(1)(q)H_{p}^{(1)}(q) as

Hp(1)(q)=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+2p+12a+52b.H_{p}^{(1)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{2p+1}{2}a+\frac{5}{2}b}.
Proof.

By the definition,

Hp(1)(q)=sp=0qsp(qsp+1)sp+1(q)spβsp(2,p)=b=0qb(q)2b+1βb(2,p).\displaystyle H_{p}^{(1)}(q)=\sum_{s_{p}=0}^{\infty}q^{s_{p}}(q^{s_{p}+1})_{s_{p}+1}(q)_{s_{p}}\beta_{s_{p}}^{(2,p)}=\sum_{b=0}^{\infty}q^{b}(q)_{2b+1}\beta_{b}^{(2,p)}.

Then, Proposition 2.8 implies that

Hp(1)(q)\displaystyle H_{p}^{(1)}(q) =b=0qb(q)2b+1a=b1b(1)aq(p+12)a2+(p12)a(q)ba(q)b+a+1.\displaystyle=\sum_{b=0}^{\infty}q^{b}(q)_{2b+1}\sum_{a=-b-1}^{b}\frac{(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}}{(q)_{b-a}(q)_{b+a+1}}.

By dividing the range of the inner sum into 0ab0\leq a\leq b and b1a1-b-1\leq a\leq-1 and changing the order of the sums,

Hp(1)\displaystyle H_{p}^{(1)} =a0(1)aq(p+12)a2+(p12)ab=aqb(q)2b+1(q)ba(q)b+a+1\displaystyle=\sum_{a\geq 0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}\sum_{b=a}^{\infty}\frac{q^{b}(q)_{2b+1}}{(q)_{b-a}(q)_{b+a+1}}
+a<0(1)aq(p+12)a2+(p12)ab=a1qb(q)2b+1(q)ba(q)b+a+1.\displaystyle\qquad+\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}\sum_{b=-a-1}^{\infty}\frac{q^{b}(q)_{2b+1}}{(q)_{b-a}(q)_{b+a+1}}.

Changing the variables in two sums so that the range of bb is b0b\geq 0 implies

Hp(1)(q)\displaystyle H_{p}^{(1)}(q) =a0(1)aq(p+12)a2+(p12)ab=0qa+b(q)2b+2a+1(q)b(q)b+2a+1\displaystyle=\sum_{a\geq 0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}\sum_{b=0}^{\infty}\frac{q^{a+b}(q)_{2b+2a+1}}{(q)_{b}(q)_{b+2a+1}}
+a<0(1)aq(p+12)a2+(p12)ab=0qba1(q)2b2a1(q)b(q)b2a1.\displaystyle\qquad+\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}\sum_{b=0}^{\infty}\frac{q^{b-a-1}(q)_{2b-2a-1}}{(q)_{b}(q)_{b-2a-1}}.

Finally, Lemma 2.12 yields that

Hp(1)(q)\displaystyle H_{p}^{(1)}(q) =1(q)a0(1)aq(p+12)a2+(p+12)ab=0(1)bq32b2+(2a+52)b\displaystyle=\frac{1}{(q)_{\infty}}\sum_{a\geq 0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p+\frac{1}{2}\right)a}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\left(2a+\frac{5}{2}\right)b}
+1(q)a<0(1)aq(p+12)a2+(p32)a1b=0(1)bq32b2+(2a+12)b\displaystyle\qquad+\frac{1}{(q)_{\infty}}\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{3}{2}\right)a-1}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\left(-2a+\frac{1}{2}\right)b}
=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+(p+12)a+52b,\displaystyle=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\left(p+\frac{1}{2}\right)a+\frac{5}{2}b},

which is the desired Hecke-type expansion. ∎

Theorem 2.14 (Hikami [Hikami2007, Theorem 3.9]).

For p1p\geq 1, we have the Hecke-type expansion for Hp(5)(q)H_{p}^{(5)}(q) as

Hp(5)(q)=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+12a+32b.H_{p}^{(5)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{1}{2}a+\frac{3}{2}b}.
Proof.

The idea of the proof is entirely the same as that of Theorem 2.13. By the definition,

Hp(5)(q)=sp=0qsp(qsp+1)sp(q)spβsp(1,p)=b=0qb(q)2bβb(1,p).H_{p}^{(5)}(q)=\sum_{s_{p}=0}^{\infty}q^{s_{p}}(q^{s_{p}+1})_{s_{p}}(q)_{s_{p}}\beta_{s_{p}}^{(1,p)}=\sum_{b=0}^{\infty}q^{b}(q)_{2b}\beta_{b}^{(1,p)}.

Then, Proposition 2.8 implies that

Hp(5)(q)=b=0qb(q)2ba=bb(1)aq(p+12)a212a(q)ba(q)b+a.H_{p}^{(5)}(q)=\sum_{b=0}^{\infty}q^{b}(q)_{2b}\sum_{a=-b}^{b}\frac{(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}-\frac{1}{2}a}}{(q)_{b-a}(q)_{b+a}}.

By dividing the range of the sum into 0ab0\leq a\leq b and ba1-b\leq a\leq-1, changing the order of the sums, and changing the variables in two sums so that the range of bb is b0b\geq 0, we have

Hp(5)(q)\displaystyle H_{p}^{(5)}(q) =a0(1)aq(p+12)a2+12ab=0qb(q)2b+2a(q)b+2a(q)b\displaystyle=\sum_{a\geq 0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\frac{1}{2}a}\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b+2a}}{(q)_{b+2a}(q)_{b}}
+a<0(1)aq(p+12)a232ab=0qb(q)2b2a(q)b2a(q)b.\displaystyle\qquad+\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}-\frac{3}{2}a}\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b-2a}}{(q)_{b-2a}(q)_{b}}.

Finally, Lemma 2.12 yields the result. ∎

In the two cases above, the straightforward calculations transformed the Habiro-type series into a form in which Lemma 2.12 can be applied. However, in the remaining three cases, additional modifications are required.

Theorem 2.15.

For p1p\geq 1, we have the Hecke-type expansion for Hp(2)(q)H_{p}^{(2)}(q) as

Hp(2)(q)=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+12a+12b.H_{p}^{(2)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{1}{2}a+\frac{1}{2}b}.
Proof.

By the definition, we have

Hp(2)(q)=sp=0qsp(qsp)sp(q)spβsp(1,p)=1+b=1qb(q)2b1(1qb)βb(1,p).H_{p}^{(2)}(q)=\sum_{s_{p}=0}^{\infty}q^{s_{p}}(q^{s_{p}})_{s_{p}}(q)_{s_{p}}\beta_{s_{p}}^{(1,p)}=1+\sum_{b=1}^{\infty}q^{b}(q)_{2b-1}(1-q^{b})\beta_{b}^{(1,p)}.

By the auxiliary Lemma 2.9,

Hp(2)(q)=1+b=1qb(q)2b1a=b+1b(1)aq(p+12)a212a(q)ba(q)b+a1.H_{p}^{(2)}(q)=1+\sum_{b=1}^{\infty}q^{b}(q)_{2b-1}\sum_{a=-b+1}^{b}\frac{(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}-\frac{1}{2}a}}{(q)_{b-a}(q)_{b+a-1}}.

We now divide the range of the sum into a>0a>0, a=0a=0, and a<0a<0. Then the same transformation yields that

Hp(2)(q)\displaystyle H_{p}^{(2)}(q) =1+qb=0qb(q)2b+1(q)b(q)b+1+a>0(1)aq(p+12)a212ab=aqb(q)2b1(q)ba(q)b+a1\displaystyle=1+q\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b+1}}{(q)_{b}(q)_{b+1}}+\sum_{a>0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}-\frac{1}{2}a}\sum_{b=a}^{\infty}\frac{q^{b}(q)_{2b-1}}{(q)_{b-a}(q)_{b+a-1}}
+a<0(1)aq(p+12)a212ab=a+1qb(q)2b1(q)ba(q)b+a1.\displaystyle\qquad+\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}-\frac{1}{2}a}\sum_{b=-a+1}^{\infty}\frac{q^{b}(q)_{2b-1}}{(q)_{b-a}(q)_{b+a-1}}.

By applying Lemma 2.12,

Hp(2)(q)\displaystyle H_{p}^{(2)}(q) =1+1(q)b=0(1)bq32b2+52b+1+1(q)a>0(1)aq(p+12)a2+12ab=0(1)bq32b2+(2a+12)b\displaystyle=1+\frac{1}{(q)_{\infty}}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\frac{5}{2}b+1}+\frac{1}{(q)_{\infty}}\sum_{a>0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\frac{1}{2}a}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\left(2a+\frac{1}{2}\right)b}
+1(q)a<0(1)aq(p+12)a232a+1b=0(1)bq32b2+(2a+52)b.\displaystyle\qquad+\frac{1}{(q)_{\infty}}\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}-\frac{3}{2}a+1}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\left(-2a+\frac{5}{2}\right)b}.

Finally, in Jacobi’s triple product (2.4), by setting qq3q\mapsto q^{3} and ζq\zeta\mapsto q, we obtain Euler’s pentagonal number theorem

(2.5) (q)=b(1)bq32b212b=b0(1)bq32b2+52b+1+b0(1)bq32b2+12b.\displaystyle(q)_{\infty}=\sum_{b\in\mathbb{Z}}(-1)^{b}q^{\frac{3}{2}b^{2}-\frac{1}{2}b}=-\sum_{b\geq 0}(-1)^{b}q^{\frac{3}{2}b^{2}+\frac{5}{2}b+1}+\sum_{b\geq 0}(-1)^{b}q^{\frac{3}{2}b^{2}+\frac{1}{2}b}.

Then we obtain the desired result by rearranging the equation. ∎

Theorem 2.16.

For p1p\geq 1, we have the Hecke-type expansion for Hp(3)(q)H_{p}^{(3)}(q) as

Hp(3)(q)=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+2p+32a+32b.H_{p}^{(3)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{2p+3}{2}a+\frac{3}{2}b}.
Proof.

By the definition and Lemma 2.10, we have

Hp(3)(q)\displaystyle H_{p}^{(3)}(q) =sp=0q2sp(qsp+1)sp(q)spβsp(2,p)=b=0qb(q)2bqbβb(2,p)\displaystyle=\sum_{s_{p}=0}^{\infty}q^{2s_{p}}(q^{s_{p}+1})_{s_{p}}(q)_{s_{p}}\beta_{s_{p}}^{(2,p)}=\sum_{b=0}^{\infty}q^{b}(q)_{2b}q^{b}\beta_{b}^{(2,p)}
=b=0qb(q)2ba=bb(1)aq(p+12)a2+(p+12)a(q)ba(q)b+a.\displaystyle=\sum_{b=0}^{\infty}q^{b}(q)_{2b}\sum_{a=-b}^{b}\frac{(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p+\frac{1}{2}\right)a}}{(q)_{b-a}(q)_{b+a}}.

The same calculation and Lemma 2.12 yield that

Hp(3)(q)\displaystyle H_{p}^{(3)}(q) =a0(1)aq(p+12)a2+(p+32)ab=0qb(q)2b+2a(q)b+2a(q)b\displaystyle=\sum_{a\geq 0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p+\frac{3}{2}\right)a}\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b+2a}}{(q)_{b+2a}(q)_{b}}
+a<0(1)aq(p+12)a2+(p12)ab=0qb(q)2b2a(q)b2a(q)b\displaystyle\qquad+\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b-2a}}{(q)_{b-2a}(q)_{b}}
=1(q)a0(1)aq(p+12)a2+(p+32)ab=0(1)bq32b2+(2a+32)b\displaystyle=\frac{1}{(q)_{\infty}}\sum_{a\geq 0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p+\frac{3}{2}\right)a}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\left(2a+\frac{3}{2}\right)b}
+1(q)a<0(1)aq(p+12)a2+(p12)ab=0(1)bq32b2+(2a+32)b,\displaystyle\qquad+\frac{1}{(q)_{\infty}}\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}\sum_{b=0}^{\infty}(-1)^{b}q^{\frac{3}{2}b^{2}+\left(-2a+\frac{3}{2}\right)b},

which finishes the proof. ∎

Theorem 2.17.

For p1p\geq 1, we have the Hecke-type expansion for Hp(4)(q)H_{p}^{(4)}(q) as

Hp(4)(q)=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+2p+12a+12b.H_{p}^{(4)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{2p+1}{2}a+\frac{1}{2}b}.
Proof.

The structure of the proof is entirely the same as that for Hp(2)(q)H_{p}^{(2)}(q). First, by the definition and Lemma 2.11,

Hp(4)(q)\displaystyle H_{p}^{(4)}(q) =sp=0qsp(qsp+1)sp(q)spβsp(2,p)=1+b=1qb(q)2b1(1q2b)βb(2,p)\displaystyle=\sum_{s_{p}=0}^{\infty}q^{s_{p}}(q^{s_{p}+1})_{s_{p}}(q)_{s_{p}}\beta_{s_{p}}^{(2,p)}=1+\sum_{b=1}^{\infty}q^{b}(q)_{2b-1}(1-q^{2b})\beta_{b}^{(2,p)}
=1+b=1qb(q)2b1a=b+1b(1)aq(p+12)a2+(p12)a(q)ba(q)b+a1.\displaystyle=1+\sum_{b=1}^{\infty}q^{b}(q)_{2b-1}\sum_{a=-b+1}^{b}\frac{(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{1}{2}\right)a}}{(q)_{b-a}(q)_{b+a-1}}.

By dividing the range of the sum into a>0a>0, a=0a=0, and a<0a<0, the equation can be transformed to

Hp(4)(q)\displaystyle H_{p}^{(4)}(q) =1+qb=0qb(q)2b+1(q)b+1(q)b+a>0(1)aq(p+12)a2+(p+12)ab=0qb(q)2b+2a1(q)b(q)b+2a1\displaystyle=1+q\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b+1}}{(q)_{b+1}(q)_{b}}+\sum_{a>0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p+\frac{1}{2}\right)a}\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b+2a-1}}{(q)_{b}(q)_{b+2a-1}}
+a<0(1)aq(p+12)a2+(p32)a+1b=0qb(q)2b2a+1(q)b(q)b2a+1.\displaystyle\qquad+\sum_{a<0}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p-\frac{3}{2}\right)a+1}\sum_{b=0}^{\infty}\frac{q^{b}(q)_{2b-2a+1}}{(q)_{b}(q)_{b-2a+1}}.

Finally, Lemma 2.12 and Euler’s pentagonal number theorem (2.5) yield the desired Hecke-type expansion. ∎

In conclusion, the goal of this section, which is to transform the five Habiro-type series into a Hecke-type series, has been achieved.

2.4. Appendix on another conjugate Bailey pair and multiple zeta values

Although we are in the middle of a discussion, we would like to make one more observation about a (conjugate) Bailey pair for another pair (u,v)(u,v).

Here we consider the most straightforward pair un=vn=1u_{n}=v_{n}=1 for any n0n\geq 0 instead of the qq-Pochhammer symbol. Then using a similar approach to the Bailey transform, if sequences (α)n(\alpha)_{n}, (βn)n(\beta_{n})_{n}, (γn)n(\gamma_{n})_{n}, and (δn)n(\delta_{n})_{n} satisfy suitable convergence conditions and the equation

(2.6) βn=k=0n1αk,γn=k=n+1δk,\displaystyle\beta_{n}=\sum_{k=0}^{n-1}\alpha_{k},\qquad\gamma_{n}=\sum_{k=n+1}^{\infty}\delta_{k},

then we have

(2.7) n=0αnγn=n=1βnδn.\displaystyle\sum_{n=0}^{\infty}\alpha_{n}\gamma_{n}=\sum_{n=1}^{\infty}\beta_{n}\delta_{n}.

In this setting, we try finding an interesting (conjugate) Bailey pair.

Lemma 2.18.

For any positive integer m>0m>0, the pair of sequences defined by

γn=1mn!m!(n+m)!,δn=1nn!m!(n+m)!\gamma_{n}=\frac{1}{m}\frac{n!m!}{(n+m)!},\qquad\delta_{n}=\frac{1}{n}\frac{n!m!}{(n+m)!}

satisfies the condition in (2.6).

Proof.

By a telescoping sum, we have

k=n+1δk=1mk=n+1((k1)!m!(k1+m)!k!m!(k+m)!)=1mn!m!(n+m)!=γn.\sum_{k=n+1}^{\infty}\delta_{k}=\frac{1}{m}\sum_{k=n+1}^{\infty}\left(\frac{(k-1)!m!}{(k-1+m)!}-\frac{k!m!}{(k+m)!}\right)=\frac{1}{m}\frac{n!m!}{(n+m)!}=\gamma_{n}.

This lemma is a critical identity in Seki–Yamamoto’s work [SekiYamamoto2019] on the proof of duality of multiple zeta values. What is the “unit Bailey pair/chain” in this case? A simple observation yields that if a pair (α,β)(\alpha,\beta) satisfies the condition (2.6), the equation (2.7) implies an identity on the sequence (αn)n(\alpha_{n})_{n},

(2.8) n=0αnmn!m!(n+m)!=n=1βnnn!m!(n+m)!=n=1(k=0n1αkn)n!m!(n+m)!.\displaystyle\sum_{n=0}^{\infty}\frac{\alpha_{n}}{m}\frac{n!m!}{(n+m)!}=\sum_{n=1}^{\infty}\frac{\beta_{n}}{n}\frac{n!m!}{(n+m)!}=\sum_{n=1}^{\infty}\left(\sum_{k=0}^{n-1}\frac{\alpha_{k}}{n}\right)\frac{n!m!}{(n+m)!}.

Putting α0(1)=0\alpha^{(1)}_{0}=0 and αn(1)=k=0n1αk/n\alpha^{(1)}_{n}=\sum_{k=0}^{n-1}\alpha_{k}/n and applying (2.8), we have

n=0αn(1)mn!m!(n+m)!=n=1(k=0n1αk(1)n)n!m!(n+m)!=n=1(n>k>l0αlkn)n!m!(n+m)!.\sum_{n=0}^{\infty}\frac{\alpha_{n}^{(1)}}{m}\frac{n!m!}{(n+m)!}=\sum_{n=1}^{\infty}\left(\sum_{k=0}^{n-1}\frac{\alpha_{k}^{(1)}}{n}\right)\frac{n!m!}{(n+m)!}=\sum_{n=1}^{\infty}\left(\sum_{n>k>l\geq 0}\frac{\alpha_{l}}{kn}\right)\frac{n!m!}{(n+m)!}.

By repeating this process, we obtain the following.

Lemma 2.19.

For any sequence (αn)n(\alpha_{n})_{n} with suitable convergence condition, it holds that

1mpn=0αnn!m!(n+m)!=n=0n=sp>sp1>>s00αs0spsp1s1n!m!(n+m)!.\frac{1}{m^{p}}\sum_{n=0}^{\infty}\alpha_{n}\frac{n!m!}{(n+m)!}=\sum_{n=0}^{\infty}\sum_{n=s_{p}>s_{p-1}>\cdots>s_{0}\geq 0}\frac{\alpha_{s_{0}}}{s_{p}s_{p-1}\cdots s_{1}}\frac{n!m!}{(n+m)!}.

We now let αn=δn,0\alpha_{n}=\delta_{n,0}. Then the above lemma yields that

1mp=sp>sp1>>s1>01spsp1s1sp!m!(sp+m)!\frac{1}{m^{p}}=\sum_{s_{p}>s_{p-1}>\cdots>s_{1}>0}\frac{1}{s_{p}s_{p-1}\cdots s_{1}}\frac{s_{p}!m!}{(s_{p}+m)!}

for any p>0p>0. Moreover, by taking the sum over mm, we have

ζ(p+1)\displaystyle\zeta(p+1) =m=11mp+1=sp>sp1>>s1>01sp1s1m=1(sp1)!(m1)!(sp+m)!\displaystyle=\sum_{m=1}^{\infty}\frac{1}{m^{p+1}}=\sum_{s_{p}>s_{p-1}>\cdots>s_{1}>0}\frac{1}{s_{p-1}\cdots s_{1}}\sum_{m=1}^{\infty}\frac{(s_{p}-1)!(m-1)!}{(s_{p}+m)!}
=sp>sp1>>s1>01sp2sp1s1=ζ(2,1,,1p1),\displaystyle=\sum_{s_{p}>s_{p-1}>\cdots>s_{1}>0}\frac{1}{s_{p}^{2}s_{p-1}\cdots s_{1}}=\zeta(2,\underbrace{1,\dots,1}_{p-1}),

the particular case of the sum formula of multiple zeta values [Granville1997]. We also note that the factor (n1)!(m1)!/(n+m)!(n-1)!(m-1)!/(n+m)! appears in Cloitre and Oloa’s expressions of the Riemann zeta values ζ(2)\zeta(2) and ζ(3)\zeta(3) (see Kawamura–Maesaka–Seki’s recent work [KawamuraMaesakaSeki2022]). It will be interesting to see what kind of identities can be obtained by considering various pairs (u,v)(u,v), (γ,δ)(\gamma,\delta), and (α,β)(\alpha,\beta).

3. False theta functions

False theta functions also come from the work of Rogers [Rogers1917London]. The defining equation looks like the ordinary theta functions but contains an extra sign term that breaks the modular property of the theta functions. The name “false theta functions” also appeared in Ramanujan’s last letter to Hardy, and it became known at about the same time as Ramanujan’s “mock theta functions”. However, as Sills [Sills2018] points out, mock theta functions have long been the subject of active research, starting with Watson [Watson1936], whereas false theta functions have not received much attention until recently. One of the reasons why the study of false theta functions has become active in recent years is due to their relationship with quantum invariants, as revealed by Lawrence–Zagier [LawrenceZagier1999]. Their work was subsequently generalized extensively by Hikami [Hikami2005IJM, Hikami2006JMP, Hikami2007] et al., and the advent of the notion of “quantum modular forms” introduced by Zagier [Zagier2010] has further accelerated the research. More recently, as a counterpart to Zwegers’ discovery [Zwegers2002] of the modular aspect of the mock theta functions, a framework for modular properties of false theta functions has been revealed by Bringmann–Nazaroglu et al. [BringmannNazaroglu2019, BKMN2023] and Goswami–Osburn [GoswamiOsburn2021] (see also Matsusaka–Terashima [MatsusakaTerashima2021]).

Let us define the false theta functions. Let AA be a symmetric r×rr\times r-matrix with integer coefficients, which is positive definite. We consider a bilinear form B:r×rB:\mathbb{R}^{r}\times\mathbb{R}^{r}\to\mathbb{R}, B(𝒙,𝒚)=𝒙TA𝒚B(\bm{x},\bm{y})=\bm{x}^{T}A\bm{y} and the associated quadratic form Q(𝒙)=12B(𝒙,𝒙)Q(\bm{x})=\frac{1}{2}B(\bm{x},\bm{x}). We take a lattice LrL\subset\mathbb{R}^{r} of rank rr such that B(L×L)B(L\times L)\subset\mathbb{Z}. Then the dual lattice LL^{*} is defined by

L={𝒙rB(𝒙,𝒚) for any 𝒚L}.L^{*}=\{\bm{x}\in\mathbb{R}^{r}\mid B(\bm{x},\bm{y})\in\mathbb{Z}\text{ for any $\bm{y}\in L$}\}.
Definition 3.1.

For an arbitrary vector 𝒄r\bm{c}\in\mathbb{R}^{r} satisfying 2Q(𝒄)=12Q(\bm{c})=1 and 𝝁L\bm{\mu}\in L^{*}, the false theta function θ~𝝁,𝒄(τ)=θ~L,B,𝝁,𝒄(τ)\widetilde{\theta}_{\bm{\mu},\bm{c}}(\tau)=\widetilde{\theta}_{L,B,\bm{\mu},\bm{c}}(\tau) is defined by

θ~𝝁,𝒄(τ)=𝒏L+𝝁sgn(B(𝒏,𝒄))qQ(𝒏),\widetilde{\theta}_{\bm{\mu},\bm{c}}(\tau)=\sum_{\bm{n}\in L+\bm{\mu}}\operatorname{sgn}(B(\bm{n},\bm{c}))q^{Q(\bm{n})},

where τ={τIm(τ)>0}\tau\in\mathbb{H}=\{\tau\in\mathbb{C}\mid\operatorname{Im}(\tau)>0\}, qQ(𝒏)=e2πiQ(𝒏)τq^{Q(\bm{n})}=e^{2\pi iQ(\bm{n})\tau}, and sgn:{1,0,1}\operatorname{sgn}:\mathbb{R}\to\{-1,0,1\} is the usual sign function with sgn(0)=0\operatorname{sgn}(0)=0.

This article considers only the cases r=1,2r=1,2 and only special lattices and bilinear forms.

3.1. One dimensional case

This subsection aims to recall Hikami’s function Φ~𝒑(τ)\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau) in terms of false theta functions and re-prove Corollary 3.9. We let A=(1)A=(1) and L=ML=\sqrt{M}\mathbb{Z} for a positive integer M>0M>0. Then the bilinear form is B(x,y)=xyB(x,y)=xy, and the dual lattice is given by L=(1/M)L^{*}=(1/\sqrt{M})\mathbb{Z}. In this setting, the false theta function is defined as

θ~μ,c(τ)=nM+μsgn(n)qn22\widetilde{\theta}_{\mu,c}(\tau)=\sum_{n\in\sqrt{M}\mathbb{Z}+\mu}\operatorname{sgn}(n)q^{\frac{n^{2}}{2}}

for μ(1/M)/M/M\mu\in(1/\sqrt{M})\mathbb{Z}/\sqrt{M}\mathbb{Z}\cong\mathbb{Z}/M\mathbb{Z} and c=1c=1. To clarify necessary and unnecessary subscripts, we redefine the false theta function.

Definition 3.2.

For a positive integer M>0M>0 and μ{0,1,,M1}\mu\in\{0,1,\dots,M-1\}, we define

θ~M,μ(1)(τ)=nμ(M)sgn(n)qn22M.\widetilde{\theta}_{M,\mu}^{(1)}(\tau)=\sum_{n\equiv\mu\ (M)}\operatorname{sgn}(n)q^{\frac{n^{2}}{2M}}.

Although not directly used in this article, the following modular transformation formulas are known by Bringmann–Nazaroglu [BringmannNazaroglu2019] (see also [MatsusakaTerashima2021]).

Proposition 3.3.

For Re(τ)0\operatorname{Re}(\tau)\neq 0, we have

θ~M,μ(1)(1τ)+sgn(Re(τ))(iτ)1/2Mν=0M1e2πiμνMθ~M,ν(1)(τ)=iM0iϑM,μ(1)(z)i(z+1/τ)𝑑z,\displaystyle\widetilde{\theta}_{M,\mu}^{(1)}\left(-\frac{1}{\tau}\right)+\operatorname{sgn}(\operatorname{Re}(\tau))\frac{(-i\tau)^{1/2}}{\sqrt{M}}\sum_{\nu=0}^{M-1}e^{2\pi i\frac{\mu\nu}{M}}\widetilde{\theta}_{M,\nu}^{(1)}(\tau)=\frac{-i}{\sqrt{M}}\int_{0}^{i\infty}\frac{\vartheta_{M,\mu}^{(1)}(z)}{\sqrt{-i(z+1/\tau)}}dz,

where ϑM,μ(1)(τ)\vartheta_{M,\mu}^{(1)}(\tau) is the classical holomorphic theta function of weight 3/23/2 defined by

ϑM,μ(1)(τ)=nμ(M)nqn22M.\vartheta_{M,\mu}^{(1)}(\tau)=\sum_{n\equiv\mu\ (M)}nq^{\frac{n^{2}}{2M}}.

In the above transformation formula, a remarkable feature of the false theta functions is that an error term expressed by the integral of a modular form appears. A similar phenomenon is observed in the transformation formula for the mock theta functions presented by Watson [Watson1936]. This similarity may be a part of the “mock vs. false” phenomenon discussed by Lawrence–Zagier [LawrenceZagier1999], Zwegers [Zwegers2001, Conjecture 2.2], and Hikami [Hikami2005RCD] et al., but many mysteries remain.

Since we will use the transformation formulas for the ordinary theta functions later, we review the definition and the claim together.

Lemma 3.4.

For a positive even integer M>0M>0 and μ{0,1,,M1}\mu\in\{0,1,\dots,M-1\}, the ordinary theta function defined by

θM,μ(1)(τ)=nμ(M)qn22M\theta_{M,\mu}^{(1)}(\tau)=\sum_{n\equiv\mu\ (M)}q^{\frac{n^{2}}{2M}}

satisfies

θM,μ(1)(τ+1)\displaystyle\theta_{M,\mu}^{(1)}(\tau+1) =e2πiμ22MθM,μ(1)(τ),\displaystyle=e^{2\pi i\frac{\mu^{2}}{2M}}\theta_{M,\mu}^{(1)}(\tau),
θM,μ(1)(1τ)\displaystyle\theta_{M,\mu}^{(1)}\left(-\frac{1}{\tau}\right) =(iτ)1/2Mν=0M1e2πiμνMθM,ν(1)(τ).\displaystyle=\frac{(-i\tau)^{1/2}}{\sqrt{M}}\sum_{\nu=0}^{M-1}e^{2\pi i\frac{\mu\nu}{M}}\theta_{M,\nu}^{(1)}(\tau).

We will now rewrite the function Φ~(p1,p2,p3)(1,2,3)(τ)\widetilde{\Phi}_{(p_{1},p_{2},p_{3})}^{(\ell_{1},\ell_{2},\ell_{3})}(\tau) that Hikami considered in [Hikami2005IJM, Hikami2007] in terms of the false theta functions we have just prepared. Let 𝒑=(p1,p2,p3)\bm{p}=(p_{1},p_{2},p_{3}) be a triple of pairwise coprime positive integers, and P=p1p2p3P=p_{1}p_{2}p_{3}. For any triple =(1,2,3)3\bm{\ell}=(\ell_{1},\ell_{2},\ell_{3})\in\mathbb{Z}^{3} satisfying 0<j<pj0<\ell_{j}<p_{j}, we define an odd periodic function χ𝒑:/2P\chi_{\bm{p}}^{\bm{\ell}}:\mathbb{Z}/2P\mathbb{Z}\to\mathbb{Z} by

(3.1) χ𝒑(n)={ε1ε2ε3if nP(1+j=13εjjpj)(mod2P),0if otherwise,\displaystyle\chi_{\bm{p}}^{\bm{\ell}}(n)=\begin{cases}-\varepsilon_{1}\varepsilon_{2}\varepsilon_{3}&\text{if }n\equiv P\left(1+\sum_{j=1}^{3}\frac{\varepsilon_{j}\ell_{j}}{p_{j}}\right)\pmod{2P},\\ 0&\text{if otherwise},\end{cases}

where 𝜺=(ε1,ε2,ε3){±1}3\bm{\varepsilon}=(\varepsilon_{1},\varepsilon_{2},\varepsilon_{3})\in\{\pm 1\}^{3}. Then the function Φ~𝒑(τ)\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau) is defined by

Φ~𝒑(τ)=n=0χ𝒑(n)qn24P.\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau)=\sum_{n=0}^{\infty}\chi_{\bm{p}}^{\bm{\ell}}(n)q^{\frac{n^{2}}{4P}}.
Lemma 3.5.

The function Φ~𝐩(τ)\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau) is expressed in terms of false theta functions as

Φ~𝒑(τ)=12𝜺{±1}3ε1ε2ε3θ~2P,μ(𝜺,)(1)(τ),\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau)=-\frac{1}{2}\sum_{\bm{\varepsilon}\in\{\pm 1\}^{3}}\varepsilon_{1}\varepsilon_{2}\varepsilon_{3}\widetilde{\theta}_{2P,\mu(\bm{\varepsilon},\bm{\ell})}^{(1)}(\tau),

where we put

μ(𝜺,)P(1+j=13εjjpj)(mod2P).\mu(\bm{\varepsilon},\bm{\ell})\equiv P\left(1+\sum_{j=1}^{3}\frac{\varepsilon_{j}\ell_{j}}{p_{j}}\right)\pmod{2P}.
Proof.

Since χ𝒑(n)=χ𝒑(n)\chi_{\bm{p}}^{\bm{\ell}}(-n)=-\chi_{\bm{p}}^{\bm{\ell}}(n) holds for any nn\in\mathbb{Z}, we have

Φ~𝒑(τ)\displaystyle\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau) =12nsgn(n)χ𝒑(n)qn24P\displaystyle=\frac{1}{2}\sum_{n\in\mathbb{Z}}\operatorname{sgn}(n)\chi_{\bm{p}}^{\bm{\ell}}(n)q^{\frac{n^{2}}{4P}}
=12𝜺{±1}3nμ(𝜺,)(2P)sgn(n)(ε1ε2ε3)qn24P,\displaystyle=\frac{1}{2}\sum_{\bm{\varepsilon}\in\{\pm 1\}^{3}}\sum_{n\equiv\mu(\bm{\varepsilon},\bm{\ell})\ (2P)}\operatorname{sgn}(n)(-\varepsilon_{1}\varepsilon_{2}\varepsilon_{3})q^{\frac{n^{2}}{4P}},

which concludes the proof. ∎

In this setting, we re-prove the limit formula of Φ~𝒑(τ)\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}(\tau) as τ1/N\tau\to 1/N shown by Hikami [Hikami2005IJM, Proposition 3]. To this end, we recall the following two analytic lemmas by Lawrence–Zagier [LawrenceZagier1999].

Lemma 3.6.

Let C:C:\mathbb{Z}\to\mathbb{C} be a periodic function whose period is MM. If its mean value equals 0, that is,

m=1MC(m)=0,\sum_{m=1}^{M}C(m)=0,

then the Dirichlet series L(s,C)=m=1C(m)msL(s,C)=\sum_{m=1}^{\infty}C(m)m^{-s} defines a holomorphic function in Re(s)>1\operatorname{Re}(s)>1 and is analytically continued to the whole \mathbb{C}-plane. The special values at negative integers satisfy

L(r,C)=Mrr+1m=1MC(m)Br+1(mM),L(-r,C)=-\frac{M^{r}}{r+1}\sum_{m=1}^{M}C(m)B_{r+1}\left(\frac{m}{M}\right),

where Bm(x)B_{m}(x) is the mm-th Bernoulli polynomial defined by

m=0Bm(x)tmm!=textet1.\sum_{m=0}^{\infty}B_{m}(x)\frac{t^{m}}{m!}=\frac{te^{xt}}{e^{t}-1}.
Lemma 3.7.

The following asymptotic expansion holds.

m=1C(m)em2tm=0L(2m,C)(t)mm!(t0+).\sum_{m=1}^{\infty}C(m)e^{-m^{2}t}\sim\sum_{m=0}^{\infty}L(-2m,C)\frac{(-t)^{m}}{m!}\quad(t\to 0+).
Proposition 3.8.

For positive integers M,N>0M,N>0 and μ{0,1,,M1}\mu\in\{0,1,\dots,M-1\} with 2μ0(modM)2\mu\not\equiv 0\pmod{M}, we have

limt0+θ~M,μ(1)(1N+it)=12MNn=12MNnψM,μ(n)eπin2MN,\lim_{t\to 0+}\widetilde{\theta}_{M,\mu}^{(1)}\left(\frac{1}{N}+it\right)=-\frac{1}{2MN}\sum_{n=1}^{2MN}n\psi_{M,\mu}(n)e^{\pi i\frac{n^{2}}{MN}},

where we put

ψM,μ(n)={εif nεμ(modM),0if otherwise,\psi_{M,\mu}(n)=\begin{cases}\varepsilon&\text{if }n\equiv\varepsilon\mu\pmod{M},\\ 0&\text{if otherwise},\end{cases}

for ε{±1}\varepsilon\in\{\pm 1\}.

Proof.

By the definition, we have

θ~M,μ(1)(1N+it)\displaystyle\widetilde{\theta}_{M,\mu}^{(1)}\left(\frac{1}{N}+it\right) =nμ(M)sgn(n)eπin2MNen2πtM=n=1ψM,μ(n)eπin2MNen2πtM.\displaystyle=\sum_{n\equiv\mu\ (M)}\operatorname{sgn}(n)e^{\pi i\frac{n^{2}}{MN}}e^{-n^{2}\frac{\pi t}{M}}=\sum_{n=1}^{\infty}\psi_{M,\mu}(n)e^{\pi i\frac{n^{2}}{MN}}e^{-n^{2}\frac{\pi t}{M}}.

Since the function CM,μ,N(n)=ψM,μ(n)eπin2MNC_{M,\mu,N}(n)=\psi_{M,\mu}(n)e^{\pi i\frac{n^{2}}{MN}} has a period 2MN2MN and its mean value equals 0, we can apply Lawrence–Zagier’s lemmas. Then we have

limt0+θ~M,μ(1)(1N+it)\displaystyle\lim_{t\to 0+}\widetilde{\theta}_{M,\mu}^{(1)}\left(\frac{1}{N}+it\right) =L(0,CM,μ,N)=n=12MNCM,μ,N(n)B1(n2MN).\displaystyle=L(0,C_{M,\mu,N})=-\sum_{n=1}^{2MN}C_{M,\mu,N}(n)B_{1}\left(\frac{n}{2MN}\right).

The fact that B1(x)=x1/2B_{1}(x)=x-1/2 concludes the proof. ∎

Corollary 3.9 ([Hikami2005IJM, Proposition 3]).
limt0+Φ~𝒑(1N+it)=12PNn=12PNnχ𝒑(n)eπin22PN.\lim_{t\to 0+}\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}\left(\frac{1}{N}+it\right)=-\frac{1}{2PN}\sum_{n=1}^{2PN}n\chi_{\bm{p}}^{\bm{\ell}}(n)e^{\pi i\frac{n^{2}}{2PN}}.
Proof.

By Lemma 3.5 and Proposition 3.8, we have

limt0+Φ~𝒑(1N+it)=18PN𝜺{±1}3ε1ε2ε3n=14PNnψ2P,μ(𝜺,)(n)eπin22PN.\lim_{t\to 0+}\widetilde{\Phi}_{\bm{p}}^{\bm{\ell}}\left(\frac{1}{N}+it\right)=\frac{1}{8PN}\sum_{\bm{\varepsilon}\in\{\pm 1\}^{3}}\varepsilon_{1}\varepsilon_{2}\varepsilon_{3}\sum_{n=1}^{4PN}n\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)e^{\pi i\frac{n^{2}}{2PN}}.

First, since ψ2P,μ(𝜺,)(n)eπin22PN\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)e^{\pi i\frac{n^{2}}{2PN}} has a period of 2PN2PN and its mean value equals 0, the inner sum is reduced to

n=14PNnψ2P,μ(𝜺,)(n)eπin22PN\displaystyle\sum_{n=1}^{4PN}n\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)e^{\pi i\frac{n^{2}}{2PN}} =n=12PNnψ2P,μ(𝜺,)(n)eπin22PN+n=12PN(2PN+n)ψ2P,μ(𝜺,)(n)eπin22PN\displaystyle=\sum_{n=1}^{2PN}n\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)e^{\pi i\frac{n^{2}}{2PN}}+\sum_{n=1}^{2PN}(2PN+n)\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)e^{\pi i\frac{n^{2}}{2PN}}
=2n=12PNnψ2P,μ(𝜺,)(n)eπin22PN.\displaystyle=2\sum_{n=1}^{2PN}n\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)e^{\pi i\frac{n^{2}}{2PN}}.

Second, we can check that

𝜺{±1}3ε1ε2ε3ψ2P,μ(𝜺,)(n)=2χ𝒑(n),\sum_{\bm{\varepsilon}\in\{\pm 1\}^{3}}\varepsilon_{1}\varepsilon_{2}\varepsilon_{3}\psi_{2P,\mu(\bm{\varepsilon},\bm{\ell})}(n)=-2\chi_{\bm{p}}^{\bm{\ell}}(n),

which finishes the proof. ∎

3.2. False theta decompositions in two dimensional case

In the case of r=2r=2, we consider L=22L=2\mathbb{Z}^{2} and a bilinear form associated with the matrix

A=(2p+1223)A=\begin{pmatrix}2p+1&2\\ 2&3\end{pmatrix}

for a positive integer p1p\geq 1. In other words, B(𝒙,𝒚)=𝒙TA𝒚B(\bm{x},\bm{y})=\bm{x}^{T}A\bm{y} and Q(𝒙)=(p+1/2)x12+2x1x2+3/2x22Q(\bm{x})=(p+1/2)x_{1}^{2}+2x_{1}x_{2}+3/2x_{2}^{2}. The dual lattice is given by L=12A12L^{*}=\frac{1}{2}A^{-1}\mathbb{Z}^{2}. As a vector 𝒄2\bm{c}\in\mathbb{R}^{2} satisfying 2Q(𝒄)=12Q(\bm{c})=1, we choose here

𝒄𝟏=13(6p1)(32),𝒄𝟐=1(2p+1)(6p1)(22p+1).\bm{c_{1}}=\frac{1}{\sqrt{3(6p-1)}}\begin{pmatrix}3\\ -2\end{pmatrix},\quad\bm{c_{2}}=\frac{1}{\sqrt{(2p+1)(6p-1)}}\begin{pmatrix}-2\\ 2p+1\end{pmatrix}.

Then for 𝝁L\bm{\mu}\in L^{*} and 𝒄{𝒄𝟏,𝒄𝟐}\bm{c}\in\{\bm{c_{1}},\bm{c_{2}}\}, we consider the false theta function

(3.2) θ~𝝁,𝒄(2)(τ)=𝒏L+𝝁sgn(B(𝒏,𝒄))qQ(𝒏)\displaystyle\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau)=\sum_{\bm{n}\in L+\bm{\mu}}\operatorname{sgn}(B(\bm{n},\bm{c}))q^{Q(\bm{n})}

defined in Definition 3.1. To distinguish it from the one-dimensional case, we intentionally add superscript here. In the next subsection, we reformulate the Hecke-type series shown in Section 2.3 using the above false theta functions. Before we do so, we prepare the decomposition formula to compute the limit values of false theta functions.

Lemma 3.10.

We define Q𝐜(𝐱)=Q(𝐱)12B(𝐱,𝐜)2Q_{\bm{c}}(\bm{x})=Q(\bm{x})-\frac{1}{2}B(\bm{x},\bm{c})^{2}, 𝐞1=(10)\bm{e}_{1}=\bigl{(}\begin{smallmatrix}1\\ 0\end{smallmatrix}\bigr{)}, and 𝐞2=(01)\bm{e}_{2}=\bigl{(}\begin{smallmatrix}0\\ 1\end{smallmatrix}\bigr{)}. For 𝐱=(x,y)\bm{x}=(x,y), we have

Q𝒄𝟏(𝒙)\displaystyle Q_{\bm{c_{1}}}(\bm{x}) =16(2x+3y)2=16B(𝒆2,𝒙)2,\displaystyle=\frac{1}{6}(2x+3y)^{2}=\frac{1}{6}B(\bm{e}_{2},\bm{x})^{2},
Q𝒄𝟐(𝒙)\displaystyle Q_{\bm{c_{2}}}(\bm{x}) =12(2p+1)((2p+1)x+2y)2=12(2p+1)B(𝒆1,𝒙)2\displaystyle=\frac{1}{2(2p+1)}((2p+1)x+2y)^{2}=\frac{1}{2(2p+1)}B(\bm{e}_{1},\bm{x})^{2}

and

B(𝒙,𝒄𝟏)=6p13x,B(𝒙,𝒄𝟐)=6p12p+1y.\displaystyle B(\bm{x},\bm{c_{1}})=\sqrt{\frac{6p-1}{3}}x,\qquad B(\bm{x},\bm{c_{2}})=\sqrt{\frac{6p-1}{2p+1}}y.
Proof.

It follows immediately from a direct calculation. ∎

Lemma 3.11.

For 𝐜{𝐜𝟏,𝐜𝟐}\bm{c}\in\{\bm{c_{1}},\bm{c_{2}}\} and 𝛍=12A1(m1m2)L\bm{\mu}=\frac{1}{2}A^{-1}\bigl{(}\begin{smallmatrix}m_{1}\\ m_{2}\end{smallmatrix}\bigr{)}\in L^{*} with (m1m2)2\bigl{(}\begin{smallmatrix}m_{1}\\ m_{2}\end{smallmatrix}\bigr{)}\in\mathbb{Z}^{2}, we have

Q𝒄𝟏(L+𝝁)=124(4+m2)2,Q𝒄𝟐(L+𝝁)=18(2p+1)(4+m1)2.Q_{\bm{c_{1}}}(L+\bm{\mu})=\frac{1}{24}(4\mathbb{Z}+m_{2})^{2},\qquad Q_{\bm{c_{2}}}(L+\bm{\mu})=\frac{1}{8(2p+1)}(4\mathbb{Z}+m_{1})^{2}.
Proof.

Lemma 3.10 and the equations

B(𝒆2,L+𝝁)\displaystyle B(\bm{e}_{2},L+\bm{\mu}) =B(𝒆2,L)+12𝒆2TAA1(m1m2)=2+m22,\displaystyle=B(\bm{e}_{2},L)+\frac{1}{2}\bm{e}_{2}^{T}AA^{-1}\begin{pmatrix}m_{1}\\ m_{2}\end{pmatrix}=2\mathbb{Z}+\frac{m_{2}}{2},
B(𝒆1,L+𝝁)\displaystyle B(\bm{e}_{1},L+\bm{\mu}) =2+m12\displaystyle=2\mathbb{Z}+\frac{m_{1}}{2}

yield the result. ∎

With these preparations, the false theta functions θ~𝝁,𝒄(2)(τ)\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau) can be decomposed into a sum of products of the one-dimensional false theta functions and the ordinary theta functions.

Theorem 3.12.

For 𝛍=12A1(m1m2)L\bm{\mu}=\frac{1}{2}A^{-1}\bigl{(}\begin{smallmatrix}m_{1}\\ m_{2}\end{smallmatrix}\bigr{)}\in L^{*} with odd integers m1,m2m_{1},m_{2}, we have

θ~𝝁,𝒄𝟏(2)(τ)=j=02θ12,m2+4j(1)(τ)θ~12(6p1),3m12m24(6p1)j(1)(τ).\widetilde{\theta}_{\bm{\mu},\bm{c_{1}}}^{(2)}(\tau)=\sum_{j=0}^{2}\theta_{12,m_{2}+4j}^{(1)}(\tau)\widetilde{\theta}_{12(6p-1),3m_{1}-2m_{2}-4(6p-1)j}^{(1)}(\tau).
Proof.

By Lemma 3.11,

θ~𝝁,𝒄𝟏(2)(τ)\displaystyle\widetilde{\theta}_{\bm{\mu},\bm{c_{1}}}^{(2)}(\tau) =𝒏L+𝝁sgn(B(𝒏,𝒄𝟏))q12B(𝒏,𝒄𝟏)2qQ𝒄𝟏(𝒏)\displaystyle=\sum_{\bm{n}\in L+\bm{\mu}}\operatorname{sgn}(B(\bm{n},\bm{c_{1}}))q^{\frac{1}{2}B(\bm{n},\bm{c_{1}})^{2}}q^{Q_{\bm{c_{1}}}(\bm{n})}
=Nq(4N+m2)224𝒏22Q𝒄𝟏(𝒏+𝝁)=(4N+m2)224sgn(B(𝒏+𝝁,𝒄𝟏))q12B(𝒏+𝝁,𝒄𝟏)2.\displaystyle=\sum_{N\in\mathbb{Z}}q^{\frac{(4N+m_{2})^{2}}{24}}\sum_{\begin{subarray}{c}\bm{n}\in 2\mathbb{Z}^{2}\\ Q_{\bm{c_{1}}}(\bm{n}+\bm{\mu})=\frac{(4N+m_{2})^{2}}{24}\end{subarray}}\operatorname{sgn}(B(\bm{n}+\bm{\mu},\bm{c_{1}}))q^{\frac{1}{2}B(\bm{n}+\bm{\mu},\bm{c_{1}})^{2}}.

By Lemma 3.10, the condition Q𝒄𝟏(𝒏+𝝁)=(4N+m2)224Q_{\bm{c_{1}}}(\bm{n}+\bm{\mu})=\frac{(4N+m_{2})^{2}}{24} for 𝒏=2(n1n2)22\bm{n}=2\bigl{(}\begin{smallmatrix}n_{1}\\ n_{2}\end{smallmatrix}\bigr{)}\in 2\mathbb{Z}^{2} is reduced to

2B(𝒆2,𝒏+𝝁)=±(4N+m2),\displaystyle 2B(\bm{e}_{2},\bm{n}+\bm{\mu})=\pm(4N+m_{2}),

that is, 2n1+3n2=N2n_{1}+3n_{2}=N or 2n1+3n2=Nm222n_{1}+3n_{2}=-N-\frac{m_{2}}{2}. However, since m2m_{2} is odd, the second possibility is eliminated. Therefore, we have

(n1n2)=n(32)+N(11)\begin{pmatrix}n_{1}\\ n_{2}\end{pmatrix}=n\begin{pmatrix}3\\ -2\end{pmatrix}+N\begin{pmatrix}-1\\ 1\end{pmatrix}

for any nn\in\mathbb{Z}. For this 𝒏\bm{n}, the value of the bilinear form is given by

B(𝒏+𝝁,𝒄𝟏)\displaystyle B(\bm{n}+\bm{\mu},\bm{c_{1}}) =6p13(6n2N+3m12m22(6p1))\displaystyle=\sqrt{\frac{6p-1}{3}}\left(6n-2N+\frac{3m_{1}-2m_{2}}{2(6p-1)}\right)
=12(6p1)n+(3m12m24(6p1)N)12(6p1).\displaystyle=\frac{12(6p-1)n+(3m_{1}-2m_{2}-4(6p-1)N)}{\sqrt{12(6p-1)}}.

Therefore, we have

θ~𝝁,𝒄𝟏(2)(τ)\displaystyle\widetilde{\theta}_{\bm{\mu},\bm{c_{1}}}^{(2)}(\tau) =Nq(4N+m2)224n3m12m24(6p1)N(mod12(6p1))sgn(n)qn224(6p1).\displaystyle=\sum_{N\in\mathbb{Z}}q^{\frac{(4N+m_{2})^{2}}{24}}\sum_{n\equiv 3m_{1}-2m_{2}-4(6p-1)N\pmod{12(6p-1)}}\operatorname{sgn}(n)q^{\frac{n^{2}}{24(6p-1)}}.

If we classify NN modulo 33, we conclude the claim. ∎

Theorem 3.13.

For 𝛍=12A1(m1m2)L\bm{\mu}=\frac{1}{2}A^{-1}\bigl{(}\begin{smallmatrix}m_{1}\\ m_{2}\end{smallmatrix}\bigr{)}\in L^{*} with odd integers m1,m2m_{1},m_{2}, we have

θ~𝝁,𝒄𝟐(2)(τ)=j=02pθ4(2p+1),m1+4j(1)(τ)θ~4(2p+1)(6p1),2m1+(2p+1)m24p(6p1)j(1)(τ).\widetilde{\theta}_{\bm{\mu},\bm{c_{2}}}^{(2)}(\tau)=\sum_{j=0}^{2p}\theta_{4(2p+1),m_{1}+4j}^{(1)}(\tau)\widetilde{\theta}_{4(2p+1)(6p-1),-2m_{1}+(2p+1)m_{2}-4p(6p-1)j}^{(1)}(\tau).
Proof.

The idea of the proof is entirely the same as that of Theorem 3.12. By Lemma 3.11,

θ~𝝁,𝒄𝟐(2)(τ)\displaystyle\widetilde{\theta}_{\bm{\mu},\bm{c_{2}}}^{(2)}(\tau) =𝒏L+𝝁sgn(B(𝒏,𝒄𝟐))q12B(𝒏,𝒄𝟐)2qQ𝒄𝟐(𝒏)\displaystyle=\sum_{\bm{n}\in L+\bm{\mu}}\operatorname{sgn}(B(\bm{n},\bm{c_{2}}))q^{\frac{1}{2}B(\bm{n},\bm{c_{2}})^{2}}q^{Q_{\bm{c_{2}}}(\bm{n})}
=Nq(4N+m1)28(2p+1)𝒏22Q𝒄𝟐(𝒏+𝝁)=(4N+m1)28(2p+1)sgn(B(𝒏+𝝁,𝒄𝟐))q12B(𝒏+𝝁,𝒄𝟐)2.\displaystyle=\sum_{N\in\mathbb{Z}}q^{\frac{(4N+m_{1})^{2}}{8(2p+1)}}\sum_{\begin{subarray}{c}\bm{n}\in 2\mathbb{Z}^{2}\\ Q_{\bm{c_{2}}}(\bm{n}+\bm{\mu})=\frac{(4N+m_{1})^{2}}{8(2p+1)}\end{subarray}}\operatorname{sgn}(B(\bm{n}+\bm{\mu},\bm{c_{2}}))q^{\frac{1}{2}B(\bm{n}+\bm{\mu},\bm{c_{2}})^{2}}.

By Lemma 3.10, the condition Q𝒄𝟐(𝒏+𝝁)=(4N+m1)28(2p+1)Q_{\bm{c_{2}}}(\bm{n}+\bm{\mu})=\frac{(4N+m_{1})^{2}}{8(2p+1)} for 𝒏=2(n1n2)22\bm{n}=2\bigl{(}\begin{smallmatrix}n_{1}\\ n_{2}\end{smallmatrix}\bigr{)}\in 2\mathbb{Z}^{2} is reduced to

2B(𝒆1,𝒏+𝝁)=±(4N+m1),\displaystyle 2B(\bm{e}_{1},\bm{n}+\bm{\mu})=\pm(4N+m_{1}),

that is, (2p+1)n1+2n2=N(2p+1)n_{1}+2n_{2}=N or (2p+1)n1+2n2=Nm12(2p+1)n_{1}+2n_{2}=-N-\frac{m_{1}}{2}. However, since m1m_{1} is odd, the second possibility is eliminated. Therefore, we have

(n1n2)=n(22p+1)+N(1p)\begin{pmatrix}n_{1}\\ n_{2}\end{pmatrix}=n\begin{pmatrix}-2\\ 2p+1\end{pmatrix}+N\begin{pmatrix}1\\ -p\end{pmatrix}

for any nn\in\mathbb{Z}. For this 𝒏\bm{n}, the value of the bilinear form is given by

B(𝒏+𝝁,𝒄𝟐)\displaystyle B(\bm{n}+\bm{\mu},\bm{c_{2}}) =6p12p+1(2(2p+1)n2pN+2m1+(2p+1)m22(6p1))\displaystyle=\sqrt{\frac{6p-1}{2p+1}}\left(2(2p+1)n-2pN+\frac{-2m_{1}+(2p+1)m_{2}}{2(6p-1)}\right)
=4(2p+1)(6p1)n+(2m1+(2p+1)m24p(6p1)N)4(2p+1)(6p1).\displaystyle=\frac{4(2p+1)(6p-1)n+(-2m_{1}+(2p+1)m_{2}-4p(6p-1)N)}{\sqrt{4(2p+1)(6p-1)}}.

Therefore, we have

θ~𝝁,𝒄𝟐(2)(τ)\displaystyle\widetilde{\theta}_{\bm{\mu},\bm{c_{2}}}^{(2)}(\tau) =Nq(4N+m1)28(2p+1)n2m1+(2p+1)m24p(6p1)N(mod4(2p+1)(6p1))sgn(n)qn28(2p+1)(6p1).\displaystyle=\sum_{N\in\mathbb{Z}}q^{\frac{(4N+m_{1})^{2}}{8(2p+1)}}\sum_{n\equiv-2m_{1}+(2p+1)m_{2}-4p(6p-1)N\pmod{4(2p+1)(6p-1)}}\operatorname{sgn}(n)q^{\frac{n^{2}}{8(2p+1)(6p-1)}}.

If we classify NN modulo 2p+12p+1, we conclude the claim. ∎

3.3. Limit values

Under the decomposition formulas given in Theorem 3.12 and Theorem 3.13, we compute the limit values of Habiro-type series as qe2πi/Nq\to e^{2\pi i/N} radially from within the unit disc. First, we transform Hecke-type series into false theta functions. The notations are the same as in Section 3.2.

Lemma 3.14.

For any pair of integers (m1,m2)2(m_{1},m_{2})\in\mathbb{Z}^{2}, we put

𝝁=(μ1μ2):=12A1(m1m2).\bm{\mu}=\begin{pmatrix}\mu_{1}\\ \mu_{2}\end{pmatrix}:=\frac{1}{2}A^{-1}\begin{pmatrix}m_{1}\\ m_{2}\end{pmatrix}.

If 0<μ1<10<\mu_{1}<1 and 0<μ2<10<\mu_{2}<1, we have

1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+m12a+m22b\displaystyle\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{m_{1}}{2}a+\frac{m_{2}}{2}b}
=q124Q(𝝁)2η(τ)𝜺=(ε1ε2){0,1}2(1)ε1+ε2(θ~𝝁+𝜺,𝒄𝟏(2)(τ)+θ~𝝁+𝜺,𝒄𝟐(2)(τ)),\displaystyle=\frac{q^{\frac{1}{24}-Q(\bm{\mu})}}{2\eta(\tau)}\sum_{\bm{\varepsilon}=\bigl{(}\begin{smallmatrix}\varepsilon_{1}\\ \varepsilon_{2}\end{smallmatrix}\bigr{)}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\bigg{(}\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c_{2}}}^{(2)}(\tau)\bigg{)},

where η(τ)=q1/24(q)\eta(\tau)=q^{1/24}(q)_{\infty} is the Dedekind eta function.

Proof.

By the definition,

𝜺{0,1}2(1)ε1+ε2(θ~𝝁+𝜺,𝒄𝟏(2)(τ)+θ~𝝁+𝜺,𝒄𝟐(2)(τ))\displaystyle\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\bigg{(}\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c_{2}}}^{(2)}(\tau)\bigg{)}
=𝜺{0,1}2(1)ε1+ε2𝒏22(sgn(B(𝒏+𝝁+𝜺,𝒄𝟏))+sgn(B(𝒏+𝝁+𝜺,𝒄𝟐)))qQ(𝒏+𝝁+𝜺).\displaystyle=\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\sum_{\bm{n}\in 2\mathbb{Z}^{2}}(\operatorname{sgn}(B(\bm{n}+\bm{\mu}+\bm{\varepsilon},\bm{c_{1}}))+\operatorname{sgn}(B(\bm{n}+\bm{\mu}+\bm{\varepsilon},\bm{c_{2}})))q^{Q(\bm{n}+\bm{\mu}+\bm{\varepsilon})}.

Since (1)ε1+ε2=(1)(n1+ε1)+(n2+ε2)(-1)^{\varepsilon_{1}+\varepsilon_{2}}=(-1)^{(n_{1}+\varepsilon_{1})+(n_{2}+\varepsilon_{2})} for any 𝒏22\bm{n}\in 2\mathbb{Z}^{2}, the above is equal to

𝒏2(1)n1+n2(sgn(B(𝒏+𝝁,𝒄𝟏))+sgn(B(𝒏+𝝁,𝒄𝟐)))qQ(𝒏+𝝁).\displaystyle\sum_{\bm{n}\in\mathbb{Z}^{2}}(-1)^{n_{1}+n_{2}}(\operatorname{sgn}(B(\bm{n}+\bm{\mu},\bm{c_{1}}))+\operatorname{sgn}(B(\bm{n}+\bm{\mu},\bm{c_{2}})))q^{Q(\bm{n}+\bm{\mu})}.

By Lemma 3.10 and our assumption 0<μ1<10<\mu_{1}<1, we have

sgn(B(𝒏+𝝁,𝒄𝟏))=sgn(n1+μ1)=sgn0(n1),\displaystyle\operatorname{sgn}(B(\bm{n}+\bm{\mu},\bm{c_{1}}))=\operatorname{sgn}(n_{1}+\mu_{1})=\operatorname{sgn}_{0}(n_{1}),

where sgn0\operatorname{sgn}_{0} is the sign function with sgn0(0)=1\operatorname{sgn}_{0}(0)=1. Similarly sgn(B(𝒏+𝝁,𝒄𝟐))=sgn0(n2)\operatorname{sgn}(B(\bm{n}+\bm{\mu},\bm{c_{2}}))=\operatorname{sgn}_{0}(n_{2}) holds. Thus we obtain

𝜺{0,1}2(1)ε1+ε2(θ~𝝁+𝜺,𝒄𝟏(2)(τ)+θ~𝝁+𝜺,𝒄𝟐(2)(τ))=2(a,b0a,b<0)(1)a+bqQ(a+μ1,b+μ2).\displaystyle\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\bigg{(}\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c_{2}}}^{(2)}(\tau)\bigg{)}=2\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{Q(a+\mu_{1},b+\mu_{2})}.

Using the relation Q(𝒙+𝝁)=Q(𝒙)+Q(𝝁)+B(𝒙,𝝁)Q(\bm{x}+\bm{\mu})=Q(\bm{x})+Q(\bm{\mu})+B(\bm{x},\bm{\mu}) concludes the proof. ∎

Theorem 3.15.

Let

𝝁𝟏\displaystyle\bm{\mu_{1}} =12A1(2p+15)=12(6p1)(6p76p+3),\displaystyle=\frac{1}{2}A^{-1}\begin{pmatrix}2p+1\\ 5\end{pmatrix}=\frac{1}{2(6p-1)}\begin{pmatrix}6p-7\\ 6p+3\end{pmatrix},
𝝁𝟐\displaystyle\bm{\mu_{2}} =12A1(11)=12(6p1)(12p1),\displaystyle=\frac{1}{2}A^{-1}\begin{pmatrix}1\\ 1\end{pmatrix}=\frac{1}{2(6p-1)}\begin{pmatrix}1\\ 2p-1\end{pmatrix},
𝝁𝟑\displaystyle\bm{\mu_{3}} =12A1(2p+33)=12(6p1)(6p+32p3),\displaystyle=\frac{1}{2}A^{-1}\begin{pmatrix}2p+3\\ 3\end{pmatrix}=\frac{1}{2(6p-1)}\begin{pmatrix}6p+3\\ 2p-3\end{pmatrix},
𝝁𝟒\displaystyle\bm{\mu_{4}} =12A1(2p+11)=12(6p1)(6p+12p1),\displaystyle=\frac{1}{2}A^{-1}\begin{pmatrix}2p+1\\ 1\end{pmatrix}=\frac{1}{2(6p-1)}\begin{pmatrix}6p+1\\ -2p-1\end{pmatrix},
𝝁𝟓\displaystyle\bm{\mu_{5}} =12A1(13)=12(6p1)(36p+1),\displaystyle=\frac{1}{2}A^{-1}\begin{pmatrix}1\\ 3\end{pmatrix}=\frac{1}{2(6p-1)}\begin{pmatrix}-3\\ 6p+1\end{pmatrix},

and

e1=(6p+5)2,e2=1,e3=36p2+84p+1,e4=(6p+1)2,e5=48p+1.\displaystyle e_{1}=(6p+5)^{2},\quad e_{2}=1,\quad e_{3}=36p^{2}+84p+1,\quad e_{4}=(6p+1)^{2},\quad e_{5}=48p+1.

The five families of the Habiro-type series defined in Definition 1.2 have the following expressions in terms of false theta functions except for the cases of p=1p=1 with k=1,3k=1,3.

Hp(k)(q)\displaystyle H_{p}^{(k)}(q) =qek24(6p1)2η(τ)𝜺{0,1}2(1)ε1+ε2(θ~𝝁𝒌+𝜺,𝒄𝟏(2)(τ)+θ~𝝁𝒌+𝜺,𝒄𝟐(2)(τ)).\displaystyle=\frac{q^{-\frac{e_{k}}{24(6p-1)}}}{2\eta(\tau)}\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\bigg{(}\widetilde{\theta}_{\bm{\mu_{k}}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\theta}_{\bm{\mu_{k}}+\bm{\varepsilon},\bm{c_{2}}}^{(2)}(\tau)\bigg{)}.
Proof.

The cases of k=1,2,3k=1,2,3 immediately follow from Theorem 2.13, Theorem 2.15, Theorem 2.16, and Lemma 3.14. As for the cases of k=4,5k=4,5, Lemma 3.14 can not be applied directly because 𝝁𝒌\bm{\mu_{k}} does not satisfy the required conditions.

First, we consider the case k=4k=4. The difference in proofs for this case comes from

sgn(B(𝒏+𝝁𝟒,𝒄𝟐))=sgn(n2+μ4,2)={+1if n2>0,1if n20,\operatorname{sgn}(B(\bm{n}+\bm{\mu_{4}},\bm{c_{2}}))=\operatorname{sgn}(n_{2}+\mu_{4,2})=\begin{cases}+1&\text{if }n_{2}>0,\\ -1&\text{if }n_{2}\leq 0,\end{cases}

which is not equal to sgn0(n2)\operatorname{sgn}_{0}(n_{2}). We recall that, by Theorem 2.17,

Hp(4)(q)=1(q)(a,b0a,b<0)(1)a+bq(p+12)a2+2ab+32b2+2p+12a+12b.H_{p}^{(4)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{a,b\geq 0}-\sum_{a,b<0}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{2p+1}{2}a+\frac{1}{2}b}.

To modify the proof of Lemma 3.14, we add the trivial term

1(q)a(1)aq(p+12)a2+(p+12)a=0-\frac{1}{(q)_{\infty}}\sum_{a\in\mathbb{Z}}(-1)^{a}q^{\left(p+\frac{1}{2}\right)a^{2}+\left(p+\frac{1}{2}\right)a}=0

to the right-hand side. Then, we get

Hp(4)(q)=1(q)(a0b>0a<0b0)(1)a+bq(p+12)a2+2ab+32b2+2p+12a+12b.H_{p}^{(4)}(q)=\frac{1}{(q)_{\infty}}\left(\sum_{\begin{subarray}{c}a\geq 0\\ b>0\end{subarray}}-\sum_{\begin{subarray}{c}a<0\\ b\leq 0\end{subarray}}\right)(-1)^{a+b}q^{\left(p+\frac{1}{2}\right)a^{2}+2ab+\frac{3}{2}b^{2}+\frac{2p+1}{2}a+\frac{1}{2}b}.

The rest is similar. As for the case of k=5k=5, we also obtain the desired formula by adding the trivial term

1(q)b(1)bq32b2+32b=0-\frac{1}{(q)_{\infty}}\sum_{b\in\mathbb{Z}}(-1)^{b}q^{\frac{3}{2}b^{2}+\frac{3}{2}b}=0

to the expression of Hp(5)(q)H_{p}^{(5)}(q) in Theorem 2.14. ∎

We note that for the two excluded cases, the equality no longer holds either. However, the difference seems to be just 1/q1/q.

Finally, we compute the limit values. For simplicity, we put

Θ~M,μ(1)(τ)\displaystyle\widetilde{\Theta}_{M,\mu}^{(1)}(\tau) =θ~M,μ(1)(τ)θ~M,μ+M/2(1)(τ),\displaystyle=\widetilde{\theta}_{M,\mu}^{(1)}(\tau)-\widetilde{\theta}_{M,\mu+M/2}^{(1)}(\tau),
ΘM,μ(1)(τ)\displaystyle\Theta_{M,\mu}^{(1)}(\tau) =θM,μ(1)(τ)θM,μ+M/2(1)(τ)\displaystyle=\theta_{M,\mu}^{(1)}(\tau)-\theta_{M,\mu+M/2}^{(1)}(\tau)

and

Θ~𝝁,𝒄(2)(τ)=𝜺{0,1}2(1)ε1+ε2θ~𝝁+𝜺,𝒄(2)(τ)\displaystyle\widetilde{\Theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau)=\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\widetilde{\theta}_{\bm{\mu}+\bm{\varepsilon},\bm{c}}^{(2)}(\tau)

for a positive even integer M>0M>0. By the false theta decompositions given in Theorem 3.12 and Theorem 3.13, we have the following.

Lemma 3.16.

The notations are the same as in Theorem 3.12 and Theorem 3.13. Then we have

Θ~𝝁,𝒄𝟏(2)(τ)\displaystyle\widetilde{\Theta}_{\bm{\mu},\bm{c_{1}}}^{(2)}(\tau) =j=02Θ12,m2+4j(1)(τ)Θ~12(6p1),3m12m24(6p1)j(1)(τ),\displaystyle=\sum_{j=0}^{2}\Theta_{12,m_{2}+4j}^{(1)}(\tau)\widetilde{\Theta}_{12(6p-1),3m_{1}-2m_{2}-4(6p-1)j}^{(1)}(\tau),
Θ~𝝁,𝒄𝟐(2)(τ)\displaystyle\widetilde{\Theta}_{\bm{\mu},\bm{c_{2}}}^{(2)}(\tau) =j=02pΘ4(2p+1),m1+4j(1)(τ)Θ~4(2p+1)(6p1),2m1+(2p+1)m24p(6p1)j(1)(τ).\displaystyle=\sum_{j=0}^{2p}\Theta_{4(2p+1),m_{1}+4j}^{(1)}(\tau)\widetilde{\Theta}_{4(2p+1)(6p-1),-2m_{1}+(2p+1)m_{2}-4p(6p-1)j}^{(1)}(\tau).
Proof.

By Theorem 3.12,

Θ~𝝁,𝒄𝟏(2)(τ)\displaystyle\widetilde{\Theta}_{\bm{\mu},\bm{c_{1}}}^{(2)}(\tau) =𝜺{0,1}2(1)ε1+ε2\displaystyle=\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}
×j=02θ12,m2+4ε1+6ε2+4j(1)(τ)θ~12(6p1),3m12m2+2(6p1)ε14(6p1)j(1)(τ).\displaystyle\quad\times\sum_{j=0}^{2}\theta_{12,m_{2}+4\varepsilon_{1}+6\varepsilon_{2}+4j}^{(1)}(\tau)\widetilde{\theta}_{12(6p-1),3m_{1}-2m_{2}+2(6p-1)\varepsilon_{1}-4(6p-1)j}^{(1)}(\tau).

Here we use the fact that

(10)=12A1(4p+24),(01)=12A1(46).\begin{pmatrix}1\\ 0\end{pmatrix}=\frac{1}{2}A^{-1}\begin{pmatrix}4p+2\\ 4\end{pmatrix},\quad\begin{pmatrix}0\\ 1\end{pmatrix}=\frac{1}{2}A^{-1}\begin{pmatrix}4\\ 6\end{pmatrix}.

The calculation

(ε1,ε2)(\varepsilon_{1},\varepsilon_{2}) m2+4ε1+6ε2+4jm_{2}+4\varepsilon_{1}+6\varepsilon_{2}+4j 3m12m2+2(6p1)ε14(6p1)j3m_{1}-2m_{2}+2(6p-1)\varepsilon_{1}-4(6p-1)j
(1,0)(1,0) m2+4(j+1)m_{2}+4(j+1) 3m12m2+6(6p1)4(6p1)(j+1)3m_{1}-2m_{2}+6(6p-1)-4(6p-1)(j+1)
(0,1)(0,1) m2+6+4jm_{2}+6+4j 3m12m24(6p1)j3m_{1}-2m_{2}-4(6p-1)j

yields the result. The same calculation works for the second claim. ∎

The expressions in Theorem 3.15 is rewritten as

Hp(k)(q)=qek24(6p1)2η(τ)(Θ~𝝁𝒌,𝒄𝟏(2)(τ)+Θ~𝝁𝒌,𝒄𝟐(2)(τ)).H_{p}^{(k)}(q)=\frac{q^{-\frac{e_{k}}{24(6p-1)}}}{2\eta(\tau)}\bigg{(}\widetilde{\Theta}_{\bm{\mu_{k}},\bm{c_{1}}}^{(2)}(\tau)+\widetilde{\Theta}_{\bm{\mu_{k}},\bm{c_{2}}}^{(2)}(\tau)\bigg{)}.

Then the second term

qek24(6p1)2η(τ)Θ~𝝁𝒌,𝒄𝟐(2)(τ)\frac{q^{-\frac{e_{k}}{24(6p-1)}}}{2\eta(\tau)}\widetilde{\Theta}_{\bm{\mu_{k}},\bm{c_{2}}}^{(2)}(\tau)

diverges in the vertical limit τ1/N\tau\to 1/N because of the Dedekind eta function in the denominator. On the other hand, the remaining first term converges in the same limit since the divergence is canceled out by the decay of Θ~𝝁𝒌,𝒄𝟏(2)(τ)\widetilde{\Theta}_{\bm{\mu_{k}},\bm{c_{1}}}^{(2)}(\tau). The difference in convergence comes from the difference of 1212 and 4(2p+1)4(2p+1) in the subscripts of Θ(1)\Theta^{(1)}. To confirm it, we recall the modular transformation formula for the Dedekind eta function,

η(τ+1)\displaystyle\eta(\tau+1) =eπi12η(τ),\displaystyle=e^{\frac{\pi i}{12}}\eta(\tau),
η(1τ)\displaystyle\eta\left(-\frac{1}{\tau}\right) =(iτ)1/2η(τ).\displaystyle=(-i\tau)^{1/2}\eta(\tau).

By combining Lemma 3.4, for an even MM, we have

(3.3) θM,μ(1)η(τ+1)=e2πi(μ22M124)θM,μ(1)η(τ),θM,μ(1)η(1τ)=1Mν=0M1e2πiμνMθM,ν(1)η(τ).\displaystyle\begin{split}\frac{\theta_{M,\mu}^{(1)}}{\eta}(\tau+1)&=e^{2\pi i\left(\frac{\mu^{2}}{2M}-\frac{1}{24}\right)}\frac{\theta_{M,\mu}^{(1)}}{\eta}(\tau),\\ \frac{\theta_{M,\mu}^{(1)}}{\eta}\left(-\frac{1}{\tau}\right)&=\frac{1}{\sqrt{M}}\sum_{\nu=0}^{M-1}e^{2\pi i\frac{\mu\nu}{M}}\frac{\theta_{M,\nu}^{(1)}}{\eta}(\tau).\end{split}

These transformations imply the following.

Lemma 3.17.

For an even integer M>0M>0 and a positive integer N>0N>0, we have

θM,μ(1)η(1N+it)=1Mν=0M1(ν=0M1e2πi(μ+ν)νMe2πiN(ν22M124))θM,ν(1)η(1N+iN2t).\frac{\theta_{M,\mu}^{(1)}}{\eta}\left(\frac{1}{N}+it\right)=\frac{1}{M}\sum_{\nu^{\prime}=0}^{M-1}\left(\sum_{\nu=0}^{M-1}e^{2\pi i\frac{(\mu+\nu^{\prime})\nu}{M}}e^{-2\pi iN\left(\frac{\nu^{2}}{2M}-\frac{1}{24}\right)}\right)\frac{\theta_{M,\nu^{\prime}}^{(1)}}{\eta}\left(-\frac{1}{N}+\frac{i}{N^{2}t}\right).
Proof.

It immediately follows from the identity

1N+it=1N11N+iN2t\frac{1}{N}+it=-\cfrac{1}{-N-\cfrac{1}{-\cfrac{1}{N}+\cfrac{i}{N^{2}t}}}

and the transformation formulas in (3.3). ∎

This lemma implies that

ΘM,μ(1)η(1N+it)\displaystyle\frac{\Theta_{M,\mu}^{(1)}}{\eta}\left(\frac{1}{N}+it\right) =2Mν=0M1(0νM1ν:odde2πi(μ+ν)νMe2πiN(ν22M124))θM,ν(1)η(1N+iN2t)\displaystyle=\frac{2}{M}\sum_{\nu^{\prime}=0}^{M-1}\left(\sum_{\begin{subarray}{c}0\leq\nu\leq M-1\\ \nu:\text{odd}\end{subarray}}e^{2\pi i\frac{(\mu+\nu^{\prime})\nu}{M}}e^{-2\pi iN\left(\frac{\nu^{2}}{2M}-\frac{1}{24}\right)}\right)\frac{\theta_{M,\nu^{\prime}}^{(1)}}{\eta}\left(-\frac{1}{N}+\frac{i}{N^{2}t}\right)
=2Mν=0M1(e2πi(μ+νMN2M+N24)ν=0M/21e2πiNν2+(μ+νN)νM/2)θM,ν(1)η(1N+iN2t).\displaystyle=\frac{2}{M}\sum_{\nu^{\prime}=0}^{M-1}\left(e^{2\pi i\left(\frac{\mu+\nu^{\prime}}{M}-\frac{N}{2M}+\frac{N}{24}\right)}\sum_{\nu=0}^{M/2-1}e^{2\pi i\frac{-N\nu^{2}+(\mu+\nu^{\prime}-N)\nu}{M/2}}\right)\frac{\theta_{M,\nu^{\prime}}^{(1)}}{\eta}\left(-\frac{1}{N}+\frac{i}{N^{2}t}\right).

Furthermore, we take M=4(2p+1)M=4(2p+1) and put b=μ+νb=\mu+\nu^{\prime}. Then the inner sum becomes

ν=0M/21e2πiNν2+(μ+νN)νM/2\displaystyle\sum_{\nu=0}^{M/2-1}e^{2\pi i\frac{-N\nu^{2}+(\mu+\nu^{\prime}-N)\nu}{M/2}} =ν=04p+1e2πiN(ν+2p+1)2+(bN)(ν+2p+1)2(2p+1)=(1)bν=04p+1e2πiNν2+(bN)ν2(2p+1).\displaystyle=\sum_{\nu=0}^{4p+1}e^{2\pi i\frac{-N(\nu+2p+1)^{2}+(b-N)(\nu+2p+1)}{2(2p+1)}}=(-1)^{b}\sum_{\nu=0}^{4p+1}e^{2\pi i\frac{-N\nu^{2}+(b-N)\nu}{2(2p+1)}}.

The equation shows that if b=μ+νb=\mu+\nu^{\prime} is odd, the inner sum equals 0.

Lemma 3.18.

For a positive integer p1p\geq 1 and an odd μ\mu, we have

Θ4(2p+1),μ(1)η(1N+it)\displaystyle\frac{\Theta_{4(2p+1),\mu}^{(1)}}{\eta}\left(\frac{1}{N}+it\right) =12(2p+1)ν=02(2p+1)1e2πi(2ν+μ+14(2p+1)N8(2p+1)+N24)\displaystyle=\frac{1}{2(2p+1)}\sum_{\nu^{\prime}=0}^{2(2p+1)-1}e^{2\pi i\left(\frac{2\nu^{\prime}+\mu+1}{4(2p+1)}-\frac{N}{8(2p+1)}+\frac{N}{24}\right)}
×ν=02(2p+1)1e2πiNν2+(2ν+μ+1N)ν2(2p+1)θ4(2p+1),2ν+1(1)η(1N+iN2t).\displaystyle\qquad\times\sum_{\nu=0}^{2(2p+1)-1}e^{2\pi i\frac{-N\nu^{2}+(2\nu^{\prime}+\mu+1-N)\nu}{2(2p+1)}}\frac{\theta_{4(2p+1),2\nu^{\prime}+1}^{(1)}}{\eta}\left(-\frac{1}{N}+\frac{i}{N^{2}t}\right).

In the particular case of p=1p=1, we obtain the following converging limit formula.

Corollary 3.19.

For an odd μ\mu, we have

limt0Θ12,μ(1)η(1N+it)={1if μ1,11(mod12),1if μ5,7(mod12),0if otherwise.\lim_{t\to 0}\frac{\Theta_{12,\mu}^{(1)}}{\eta}\left(\frac{1}{N}+it\right)=\begin{cases}1&\text{if }\mu\equiv 1,11\pmod{12},\\ -1&\text{if }\mu\equiv 5,7\pmod{12},\\ 0&\text{if otherwise}.\end{cases}
Proof.

Since μ\mu is odd, by applying Lemma 3.18, we obtain

Θ12,μ(1)η(1N+it)\displaystyle\frac{\Theta_{12,\mu}^{(1)}}{\eta}\left(\frac{1}{N}+it\right) =16ν=05e2πi(2ν+μ+112)\displaystyle=\frac{1}{6}\sum_{\nu^{\prime}=0}^{5}e^{2\pi i\left(\frac{2\nu^{\prime}+\mu+1}{12}\right)}
×ν=05e2πiNν2+(2ν+μ+1N)ν6θ12,2ν+1(1)η(1N+iN2t).\displaystyle\qquad\times\sum_{\nu=0}^{5}e^{2\pi i\frac{-N\nu^{2}+(2\nu^{\prime}+\mu+1-N)\nu}{6}}\frac{\theta_{12,2\nu^{\prime}+1}^{(1)}}{\eta}\left(-\frac{1}{N}+\frac{i}{N^{2}t}\right).

By the definition of the theta function θM,μ(1)(τ)\theta_{M,\mu}^{(1)}(\tau) in Lemma 3.4, we see that

limτiq1/24θ12,2ν+1(1)(τ)={1if 2ν+11,11(mod12),0if otherwise.\lim_{\tau\to i\infty}q^{-1/24}\theta_{12,2\nu^{\prime}+1}^{(1)}(\tau)=\begin{cases}1&\text{if }2\nu^{\prime}+1\equiv 1,11\pmod{12},\\ 0&\text{if otherwise}.\end{cases}

Therefore,

limt0Θ12,μ(1)η(1N+it)\displaystyle\lim_{t\to 0}\frac{\Theta_{12,\mu}^{(1)}}{\eta}\left(\frac{1}{N}+it\right) =16e2πiμ+112ν=05e2πiNν2+(μ+1N)ν6+16e2πiμ112ν=05e2πiNν2+(μ1N)ν6\displaystyle=\frac{1}{6}e^{2\pi i\frac{\mu+1}{12}}\sum_{\nu=0}^{5}e^{2\pi i\frac{-N\nu^{2}+(\mu+1-N)\nu}{6}}+\frac{1}{6}e^{2\pi i\frac{\mu-1}{12}}\sum_{\nu=0}^{5}e^{2\pi i\frac{-N\nu^{2}+(\mu-1-N)\nu}{6}}
={1if μ1,11(mod12),1if μ5,7(mod12),0if otherwise,\displaystyle=\begin{cases}1&\text{if }\mu\equiv 1,11\pmod{12},\\ -1&\text{if }\mu\equiv 5,7\pmod{12},\\ 0&\text{if otherwise},\end{cases}

which concludes the proof. ∎

Corollary 3.20.

For any pair of odd integers (m1,m2)2(m_{1},m_{2})\in\mathbb{Z}^{2}, we put 𝛍=12A1(m1m2)\bm{\mu}=\frac{1}{2}A^{-1}\bigl{(}\begin{smallmatrix}m_{1}\\ m_{2}\end{smallmatrix}\bigr{)}. Then we have

limτ1/NΘ~𝝁,𝒄𝟏(2)(τ)η(τ)\displaystyle\lim_{\tau\to 1/N}\frac{\widetilde{\Theta}_{\bm{\mu},\bm{c_{1}}}^{(2)}(\tau)}{\eta(\tau)}
=θ~12(6p1),3m12m24(6p1)j1(1)(1/N)θ~12(6p1),3m12m24(6p1)j1+6(6p1)(1)(1/N)\displaystyle=\widetilde{\theta}_{12(6p-1),3m_{1}-2m_{2}-4(6p-1)j_{1}}^{(1)}(1/N)-\widetilde{\theta}_{12(6p-1),3m_{1}-2m_{2}-4(6p-1)j_{1}+6(6p-1)}^{(1)}(1/N)
θ~12(6p1),3m12m24(6p1)j2(1)(1/N)+θ~12(6p1),3m12m24(6p1)j2+6(6p1)(1)(1/N),\displaystyle\qquad-\widetilde{\theta}_{12(6p-1),3m_{1}-2m_{2}-4(6p-1)j_{2}}^{(1)}(1/N)+\widetilde{\theta}_{12(6p-1),3m_{1}-2m_{2}-4(6p-1)j_{2}+6(6p-1)}^{(1)}(1/N),

where we put

(j1,j2)={(0,1)if m21(mod12),(2,1)if m23(mod12),(2,0)if m25(mod12),(1,0)if m27(mod12),(1,2)if m29(mod12),(0,2)if m211(mod12).\displaystyle(j_{1},j_{2})=\begin{cases}(0,1)&\text{if }m_{2}\equiv 1\pmod{12},\\ (2,1)&\text{if }m_{2}\equiv 3\pmod{12},\\ (2,0)&\text{if }m_{2}\equiv 5\pmod{12},\\ (1,0)&\text{if }m_{2}\equiv 7\pmod{12},\\ (1,2)&\text{if }m_{2}\equiv 9\pmod{12},\\ (0,2)&\text{if }m_{2}\equiv 11\pmod{12}.\end{cases}

Here we put θ~M,μ(1)(1/N)=limτ1/Nθ~M,μ(1)(τ).\widetilde{\theta}_{M,\mu}^{(1)}(1/N)=\lim_{\tau\to 1/N}\widetilde{\theta}_{M,\mu}^{(1)}(\tau).

Proof.

It immediately follows from Lemma 3.16 and Corollary 3.19. ∎

Theorem 3.21.

The limit of the first half of the expression of each of five Habiro-type series given in Theorem 3.15 converges as τ1/N\tau\to 1/N. More precisely, we have

limτ1/Nqek24(6p1)2η(τ)𝜺{0,1}2(1)ε1+ε2θ~𝝁𝒌+𝜺,𝒄𝟏(2)(τ)\displaystyle\lim_{\tau\to 1/N}\frac{q^{-\frac{e_{k}}{24(6p-1)}}}{2\eta(\tau)}\sum_{\bm{\varepsilon}\in\{0,1\}^{2}}(-1)^{\varepsilon_{1}+\varepsilon_{2}}\widetilde{\theta}_{\bm{\mu_{k}}+\bm{\varepsilon},\bm{c_{1}}}^{(2)}(\tau) =12limτ1/Nqek24(6p1)Φ~(2,3,6p1)(1,1,k)(τ),\displaystyle=-\frac{1}{2}\lim_{\tau\to 1/N}q^{-\frac{e_{k}}{24(6p-1)}}\widetilde{\Phi}_{(2,3,6p-1)}^{(1,1,\ell_{k})}(\tau),

where (k)1k5=(1,p,2p1,2p,3p1)(\ell_{k})_{1\leq k\leq 5}=(1,p,2p-1,2p,3p-1).

Proof.

By Corollary 3.20, the limit on the left-hand side converges. The subscripts of the resulting false theta functions are listed below.

(m1,m2)(m_{1},m_{2}) 3m12m24(6p1)j13m_{1}-2m_{2}-4(6p-1)j_{1} 3m12m24(6p1)j23m_{1}-2m_{2}-4(6p-1)j_{2}
(2p+1,5)(2p+1,5) 42p+1-42p+1 6p76p-7
(1,1)(1,1) 11 24p+5-24p+5
(2p+3,3)(2p+3,3) 42p+11-42p+11 18p+7-18p+7
(2p+1,1)(2p+1,1) 6p+16p+1 18p+5-18p+5
(1,3)(1,3) 48p+5-48p+5 24p+1-24p+1

For 𝒑=(2,3,6p1)\bm{p}=(2,3,6p-1), the values of

μ(𝜺,)=6(6p1)(1+ε112+ε223+ε336p1)mod12(6p1)\mu(\bm{\varepsilon},\bm{\ell})=6(6p-1)\left(1+\frac{\varepsilon_{1}\ell_{1}}{2}+\frac{\varepsilon_{2}\ell_{2}}{3}+\frac{\varepsilon_{3}\ell_{3}}{6p-1}\right)\mod{12(6p-1)}

defined in Lemma 3.5 coincide with the values in the above list for some (𝜺,)(\bm{\varepsilon},\bm{\ell})’s as follows.

=(1,2,3)\bm{\ell}=(\ell_{1},\ell_{2},\ell_{3}) 𝜺\bm{\varepsilon} μ(𝜺,)\mu(\bm{\varepsilon},\bm{\ell}) 𝜺\bm{\varepsilon} μ(𝜺,)\mu(\bm{\varepsilon},\bm{\ell})
𝟏=(1,1,1)\bm{\ell_{1}}=(1,1,1) (1,1,1)(-1,1,-1) 42p+1-42p+1 (1,1,1)(-1,-1,-1) 6p76p-7
𝟐=(1,1,p)\bm{\ell_{2}}=(1,1,p) (1,1,1)(1,1,1) 11 (1,1,1)(1,-1,1) 24p+5-24p+5
𝟑=(1,1,2p1)\bm{\ell_{3}}=(1,1,2p-1) (1,1,1)(1,-1,-1) 42p+11-42p+11 (1,1,1)(1,1,-1) 18p+7-18p+7
𝟒=(1,1,2p)\bm{\ell_{4}}=(1,1,2p) (1,1,1)(1,1,1) 6p+16p+1 (1,1,1)(1,-1,1) 18p+5-18p+5
𝟓=(1,1,3p1)\bm{\ell_{5}}=(1,1,3p-1) (1,1,1)(-1,-1,1) 48p+5-48p+5 (1,1,1)(-1,1,1) 24p+1-24p+1

By the relations μ(𝜺,)μ(𝜺,)\mu(-\bm{\varepsilon},\bm{\ell})\equiv-\mu(\bm{\varepsilon},\bm{\ell}) and μ((ε1,ε2,ε3),)μ((ε1,ε2,ε3),)+6(6p1)\mu((-\varepsilon_{1},\varepsilon_{2},\varepsilon_{3}),\bm{\ell})\equiv\mu((\varepsilon_{1},\varepsilon_{2},\varepsilon_{3}),\bm{\ell})+6(6p-1) for 1=1\ell_{1}=1, we have

Φ~𝒑𝟏(τ)\displaystyle\widetilde{\Phi}_{\bm{p}}^{\bm{\ell_{1}}}(\tau) =θ~12(6p1),6p7(1)(τ)θ~12(6p1),42p+1(1)(τ)\displaystyle=\widetilde{\theta}_{12(6p-1),6p-7}^{(1)}(\tau)-\widetilde{\theta}_{12(6p-1),-42p+1}^{(1)}(\tau)
θ~12(6p1),6p7+6(6p1)(1)(τ)+θ~12(6p1),42p+1+6(6p1)(1)(τ),\displaystyle\qquad-\widetilde{\theta}_{12(6p-1),6p-7+6(6p-1)}^{(1)}(\tau)+\widetilde{\theta}_{12(6p-1),-42p+1+6(6p-1)}^{(1)}(\tau),

which implies that

limτ1/NΘ~𝝁𝟏,𝒄𝟏(2)(τ)η(τ)=limτ1/NΦ~𝒑𝟏(τ).\lim_{\tau\to 1/N}\frac{\widetilde{\Theta}_{\bm{\mu_{1}},\bm{c_{1}}}^{(2)}(\tau)}{\eta(\tau)}=-\lim_{\tau\to 1/N}\widetilde{\Phi}_{\bm{p}}^{\bm{\ell_{1}}}(\tau).

The same calculation works for the remaining four cases. ∎

As for the second half of the expressions of the Habiro-type series, the corresponding limit

limt0Θ4(2p+1),m1+4j(1)η(1N+it)\lim_{t\to 0}\frac{\Theta_{4(2p+1),m_{1}+4j}^{(1)}}{\eta}\left(\frac{1}{N}+it\right)

to Corollary 3.19 diverges in general because of the Dedekind eta function in the denominator. In other words, the limit limqe2πi/NHp(k)(q)\lim_{q\to e^{2\pi i/N}}H_{p}^{(k)}(q) from within the unit disc diverges in general.

3.4. Hikami’s question, revisited

In [Hikami2007, Concluding remarks], Hikami left the question on the modular transformation theory of the Hecke-type series expression of the Habiro-type series as a future study. Hikami’s question locates in the counterpart of the transformation theory of the indefinite-theta function expressions of Ramanujan’s mock theta functions developed by Zwegers [Zwegers2002]. As mentioned at the beginning of Section 3, the work of Bringmann–Nazaroglu [BringmannNazaroglu2019] on the transformation theory of false theta functions is one answer to this question. In this last subsection, we will briefly review it.

We now consider the false theta function

θ~𝝁,𝒄(2)(τ)=𝒏L+𝝁sgn(B(𝒏,𝒄))qQ(𝒏)\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau)=\sum_{\bm{n}\in L+\bm{\mu}}\operatorname{sgn}(B(\bm{n},\bm{c}))q^{Q(\bm{n})}

for the general setting in Definition 3.1 with rankL=2\mathrm{rank}L=2. Recalling the definition of Q𝒄(𝒙)Q_{\bm{c}}(\bm{x}) defined in Lemma 3.10 and the expression

θ~𝝁,𝒄(2)(τ)=𝒏L+𝝁sgn(B(𝒏,𝒄))q12B(𝒏,𝒄)2qQ𝒄(𝒏),\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau)=\sum_{\bm{n}\in L+\bm{\mu}}\operatorname{sgn}(B(\bm{n},\bm{c}))q^{\frac{1}{2}B(\bm{n},\bm{c})^{2}}q^{Q_{\bm{c}}(\bm{n})},

we introduce the function

fτ,z(𝒙)=B(𝒙,𝒄)eπizB(𝒙,𝒄)2e2πiτQ𝒄(𝒙)=B(𝒙,𝒄)e2πiQ(𝒙)τeπi(zτ)B(𝒙,𝒄)2f_{\tau,z}(\bm{x})=B(\bm{x},\bm{c})e^{\pi izB(\bm{x},\bm{c})^{2}}e^{2\pi i\tau Q_{\bm{c}}(\bm{x})}=B(\bm{x},\bm{c})e^{2\pi iQ(\bm{x})\tau}e^{\pi i(z-\tau)B(\bm{x},\bm{c})^{2}}

for τ,z\tau,z\in\mathbb{H} and 𝒙2\bm{x}\in\mathbb{R}^{2}.

Lemma 3.22.

We have

(fτ,z)(𝒙)=i(iτ)1/2(iz)3/2detAf1τ,1z(𝒙),\mathcal{F}(f_{\tau,z})(\bm{x})=\frac{-i(-i\tau)^{-1/2}(-iz)^{-3/2}}{\sqrt{\det A}}f_{-\frac{1}{\tau},-\frac{1}{z}}(\bm{x}),

where

(f)(𝒙)=2f(𝒚)e2πiB(𝒙,𝒚)𝑑𝒚\mathcal{F}(f)(\bm{x})=\int_{\mathbb{R}^{2}}f(\bm{y})e^{-2\pi iB(\bm{x},\bm{y})}d\bm{y}

is the Fourier transform of ff.

Proof.

The idea of the proof is based on Bringmann–Nazaroglu [BringmannNazaroglu2019]. By the definition,

(fτ,z)(𝒙)=2B(𝒚,𝒄)e2πiQ(𝒚)τeπi(zτ)B(𝒚,𝒄)2e2πiB(𝒙,𝒚)𝑑𝒚.\displaystyle\mathcal{F}(f_{\tau,z})(\bm{x})=\int_{\mathbb{R}^{2}}B(\bm{y},\bm{c})e^{2\pi iQ(\bm{y})\tau}e^{\pi i(z-\tau)B(\bm{y},\bm{c})^{2}}e^{-2\pi iB(\bm{x},\bm{y})}d\bm{y}.

Let A=(abbd)A=\bigl{(}\begin{smallmatrix}a&b\\ b&d\end{smallmatrix}\bigr{)} and 𝒄=(c1c2)\bm{c}=\bigl{(}\begin{smallmatrix}c_{1}\\ c_{2}\end{smallmatrix}\bigr{)}, we put

C=12Q(𝒄)(ac1+bc2c2detAbc1+dc2c1detA).C=\frac{1}{\sqrt{2Q(\bm{c})}}\begin{pmatrix}ac_{1}+bc_{2}&c_{2}\sqrt{\det A}\\ bc_{1}+dc_{2}&-c_{1}\sqrt{\det A}\end{pmatrix}.

Then we have A=CCTA=CC^{T} and CT𝒄=(2Q(𝒄)0)C^{T}\bm{c}=\bigl{(}\begin{smallmatrix}\sqrt{2Q(\bm{c})}\\ 0\end{smallmatrix}\bigr{)}. Since 2Q(𝒄)=12Q(\bm{c})=1, we have CT𝒄=(10)C^{T}\bm{c}=\bigl{(}\begin{smallmatrix}1\\ 0\end{smallmatrix}\bigr{)}. By putting (y1y2)=CT𝒚\bigl{(}\begin{smallmatrix}y_{1}\\ y_{2}\end{smallmatrix}\bigr{)}=C^{T}\bm{y} and (x1x2)=CT𝒙\bigl{(}\begin{smallmatrix}x_{1}\\ x_{2}\end{smallmatrix}\bigr{)}=C^{T}\bm{x},

(fτ,z)(𝒙)\displaystyle\mathcal{F}(f_{\tau,z})(\bm{x}) =2y1eπi(y12+y22)τeπi(zτ)y12e2πi(x1y1+x2y2)dy1dy2detA\displaystyle=\int_{\mathbb{R}^{2}}y_{1}e^{\pi i(y_{1}^{2}+y_{2}^{2})\tau}e^{\pi i(z-\tau)y_{1}^{2}}e^{-2\pi i(x_{1}y_{1}+x_{2}y_{2})}\frac{dy_{1}dy_{2}}{\sqrt{\det A}}
=1detAy1eπiy12z2πix1y1𝑑y1eπiy22τ2πix2y2𝑑y2.\displaystyle=\frac{1}{\sqrt{\det A}}\int_{-\infty}^{\infty}y_{1}e^{\pi iy_{1}^{2}z-2\pi ix_{1}y_{1}}dy_{1}\int_{-\infty}^{\infty}e^{\pi iy_{2}^{2}\tau-2\pi ix_{2}y_{2}}dy_{2}.

Each integral is well-known and is equal to

=idetA(iτ)1/2(iz)3/2x1eπix121zeπix221τ\displaystyle=\frac{-i}{\sqrt{\det A}}(-i\tau)^{-1/2}(-iz)^{-3/2}x_{1}e^{\pi ix_{1}^{2}\frac{-1}{z}}e^{\pi ix_{2}^{2}\frac{-1}{\tau}}
=idetA(iτ)1/2(iz)3/2B(𝒙,𝒄)e2πiQ(𝒙)1τeπi(1z1τ)B(𝒙,𝒄)2,\displaystyle=\frac{-i}{\sqrt{\det A}}(-i\tau)^{-1/2}(-iz)^{-3/2}B(\bm{x},\bm{c})e^{2\pi iQ(\bm{x})\frac{-1}{\tau}}e^{\pi i(\frac{-1}{z}-\frac{-1}{\tau})B(\bm{x},\bm{c})^{2}},

which equals the right-hand side of the desired equation. ∎

Lemma 3.23.

We define the bivariate theta function g𝛍,𝐜(τ,z)g_{\bm{\mu},\bm{c}}(\tau,z) by

g𝝁,𝒄(τ,z)=𝒏L+𝝁fτ,z(𝒏)=𝒏L+𝝁B(𝒏,𝒄)eπizB(𝒏,𝒄)2qQ𝒄(𝒏).g_{\bm{\mu},\bm{c}}(\tau,z)=\sum_{\bm{n}\in L+\bm{\mu}}f_{\tau,z}(\bm{n})=\sum_{\bm{n}\in L+\bm{\mu}}B(\bm{n},\bm{c})e^{\pi izB(\bm{n},\bm{c})^{2}}q^{Q_{\bm{c}}(\bm{n})}.

Then we have

g𝝁,𝒄(1τ,1z)=i(iτ)1/2(iz)3/2vol(2/L)detA𝝂L/Le2πiB(𝝂,𝝁)g𝝂,𝒄(τ,z).g_{\bm{\mu},\bm{c}}\left(-\frac{1}{\tau},-\frac{1}{z}\right)=\frac{-i(-i\tau)^{1/2}(-iz)^{3/2}}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\nu},\bm{\mu})}g_{\bm{\nu},\bm{c}}(\tau,z).
Proof.

It follows from Poisson’s summation formula

vol(2/L)𝒏Lf(𝒏+𝒙)=𝒏L(f)(𝒏)e2πiB(𝒏,𝒙).\operatorname{vol}(\mathbb{R}^{2}/L)\sum_{\bm{n}\in L}f(\bm{n}+\bm{x})=\sum_{\bm{n}\in L^{*}}\mathcal{F}(f)(\bm{n})e^{2\pi iB(\bm{n},\bm{x})}.

Lemma 3.24.
θ~𝝁,𝒄(2)(τ)=iτig𝝁,𝒄(τ,z)i(zτ)𝑑z.\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau)=-i\int_{\tau}^{i\infty}\frac{g_{\bm{\mu},\bm{c}}(\tau,z)}{\sqrt{-i(z-\tau)}}dz.
Proof.

Since

τieπia2zdzi(zτ)=ieπia2τ0eπa2tdtt=ieπia2τ|a|,\int_{\tau}^{i\infty}\frac{e^{\pi ia^{2}z}dz}{\sqrt{-i(z-\tau)}}=ie^{\pi ia^{2}\tau}\int_{0}^{\infty}\frac{e^{-\pi a^{2}t}dt}{\sqrt{t}}=\frac{ie^{\pi ia^{2}\tau}}{|a|},

we have

θ~𝝁,𝒄(2)(τ)\displaystyle\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}(\tau) =i𝒏L+𝝁B(𝒏,𝒄)0B(𝒏,𝒄)ieπiB(𝒏,𝒄)2τ|B(𝒏,𝒄)|qQ𝒄(𝒏)\displaystyle=-i\sum_{\begin{subarray}{c}\bm{n}\in L+\bm{\mu}\\ B(\bm{n},\bm{c})\neq 0\end{subarray}}B(\bm{n},\bm{c})\frac{ie^{\pi iB(\bm{n},\bm{c})^{2}\tau}}{|B(\bm{n},\bm{c})|}q^{Q_{\bm{c}}(\bm{n})}
=i𝒏L+𝝁B(𝒏,𝒄)τieπizB(𝒏,𝒄)2dzi(zτ)qQ𝒄(𝒏),\displaystyle=-i\sum_{\bm{n}\in L+\bm{\mu}}B(\bm{n},\bm{c})\int_{\tau}^{i\infty}\frac{e^{\pi izB(\bm{n},\bm{c})^{2}}dz}{\sqrt{-i(z-\tau)}}q^{Q_{\bm{c}}(\bm{n})},

which finishes the proof. ∎

The above integral expression and the modular transformation of g𝝁,𝒄(τ,z)g_{\bm{\mu},\bm{c}}(\tau,z) yield the following SS-transformation formula.

Theorem 3.25.

We assume that Re(τ)0\operatorname{Re}(\tau)\neq 0. Then we have

(iτ)1θ~𝝁,𝒄(2)(1τ)=isgn(Re(τ))vol(2/L)detA𝝂L/Le2πiB(𝝁,𝝂)0τg𝝂,𝒄(τ,z)i(zτ)𝑑z.\displaystyle(-i\tau)^{-1}\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}\left(-\frac{1}{\tau}\right)=-\frac{i\operatorname{sgn}(\operatorname{Re}(\tau))}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\mu},\bm{\nu})}\int_{0}^{\tau}\frac{g_{\bm{\nu},\bm{c}}(\tau,z)}{\sqrt{-i(z-\tau)}}dz.
Proof.

By changing a variable via z=1/zz=-1/z^{\prime}, we have

(iτ)1θ~𝝁,𝒄(2)(1τ)=i(iτ)10τg𝝁,𝒄(1τ,1z)i(1z1τ)dz(iz)2.\displaystyle(-i\tau)^{-1}\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}\left(-\frac{1}{\tau}\right)=-i(-i\tau)^{-1}\int_{0}^{\tau}\frac{g_{\bm{\mu},\bm{c}}\left(-\frac{1}{\tau},-\frac{1}{z^{\prime}}\right)}{\sqrt{-i\left(-\frac{1}{z^{\prime}}-\frac{-1}{\tau}\right)}}\frac{dz^{\prime}}{(-iz^{\prime})^{2}}.

By Lemma 3.23, it becomes

=1vol(2/L)detA𝝂L/Le2πiB(𝝂,𝝁)0τg𝝂,𝒄(τ,z)i(zτ)𝑑z.=-\frac{1}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\nu},\bm{\mu})}\int_{0}^{\tau}\frac{g_{\bm{\nu},\bm{c}}(\tau,z)}{\sqrt{i(z-\tau)}}dz.

Since i(zτ)=isgn(Re(τ))i(zτ)\sqrt{i(z-\tau)}=-i\operatorname{sgn}(\operatorname{Re}(\tau))\sqrt{-i(z-\tau)} holds for zz\in\mathbb{H} on the line segment connecting 0 and τ\tau, we have the desired result. ∎

By adding

sgn(Re(τ))vol(2/L)detA𝝂L/Le2πiB(𝝁,𝝂)θ~𝝂,𝒄(2)(τ)\displaystyle\frac{\operatorname{sgn}(\operatorname{Re}(\tau))}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\mu},\bm{\nu})}\widetilde{\theta}_{\bm{\nu},\bm{c}}^{(2)}(\tau)
=isgn(Re(τ))vol(2/L)detA𝝂L/Le2πiB(𝝁,𝝂)τig𝝂,𝒄(τ,z)i(zτ)𝑑z\displaystyle=-\frac{i\operatorname{sgn}(\operatorname{Re}(\tau))}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\mu},\bm{\nu})}\int_{\tau}^{i\infty}\frac{g_{\bm{\nu},\bm{c}}(\tau,z)}{\sqrt{-i(z-\tau)}}dz

to the both sides of Theorem 3.25, we also obtain the following.

Corollary 3.26.

We assume that Re(τ)0\operatorname{Re}(\tau)\neq 0. Then we have

(iτ)1θ~𝝁,𝒄(2)(1τ)+sgn(Re(τ))vol(2/L)detA𝝂L/Le2πiB(𝝁,𝝂)θ~𝝂,𝒄(2)(τ)\displaystyle(-i\tau)^{-1}\widetilde{\theta}_{\bm{\mu},\bm{c}}^{(2)}\left(-\frac{1}{\tau}\right)+\frac{\operatorname{sgn}(\operatorname{Re}(\tau))}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\mu},\bm{\nu})}\widetilde{\theta}_{\bm{\nu},\bm{c}}^{(2)}(\tau)
=isgn(Re(τ))vol(2/L)detA𝝂L/Le2πiB(𝝁,𝝂)0ig𝝂,𝒄(τ,z)i(zτ)𝑑z,\displaystyle=-\frac{i\operatorname{sgn}(\operatorname{Re}(\tau))}{\operatorname{vol}(\mathbb{R}^{2}/L)\sqrt{\det A}}\sum_{\bm{\nu}\in L^{*}/L}e^{2\pi iB(\bm{\mu},\bm{\nu})}\int_{0}^{i\infty}\frac{g_{\bm{\nu},\bm{c}}(\tau,z)}{\sqrt{-i(z-\tau)}}dz,

where the integration path avoids the branch cut defined by i(zτ)\sqrt{-i(z-\tau)}, that is, {z=τiuu>0}\{z=\tau-iu\in\mathbb{C}\mid u>0\}.

If a shape similar to Proposition 3.3 is desired, we can re-apply Lemma 3.23 to the right-hand side.

Acknowledgements

The author would like to express his sincere gratitude to Kazuhiro Hikami for his introduction to the theory of quantum invariants in a series of lectures and seminars. The author is also grateful to Yuya Murakami, Shin-ichiro Seki, and Shoma Sugimoto for continuous helpful communication. The work was supported by JSPS KAKENHI Grant Number JP20K14292 and JP21K18141.

References