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

\DeclareTOCStyleEntry

[ beforeskip=.4em plus 1pt,pagenumberformat=]toclinesection

ceffc_{\text{eff}} for 3d 𝒩=2\mathcal{N}=2 theories

Sergei Gukova,b, Mrunmay Jagadalec
(aRichard N. Merkin Center for Pure and Applied Mathematics, California Institute of Technology, Pasadena, CA 91125, USA
bDublin Institute for Advanced Studies, 10 Burlington Rd, Dublin, Ireland
cWalter Burke Institute for Theoretical Physics, California Institute of Technology, Pasadena, CA 91125, USA
)
Abstract

Based on the observed behavior of the superconformal index in three-dimensional 𝒩=2\mathcal{N}=2 theories, we propose a quantity that can be considered as an analogue of the “effective central charge.” We discuss the general properties of this quantity and ways of computing it in a variety of different theories, including simple Lagrangian theories as well as more interesting strongly coupled examples that come from 3d-3d correspondence.

1 Introduction

Effective measures of degrees of freedom in Quantum Field Theories can be especially useful in understanding RG flows that go way outside the perturbative regime. Each time there is a candidate for such a quantity — often called a “cc-function” — one faces a variety of questions: Is it defined away from fixed points of the RG flow? If so, is the flow a gradient flow?

The answer to these questions very much depends on dd, the space-time dimension. The state of affairs is perhaps most satisfying for d=2d=2 since, in this case, RG-flows are indeed gradient flows with respect to the cc-function [1]. The next best case is in d=4d=4 where a weaker version of the cc-theorem holds, in the sense cUV>cIRc_{UV}>c_{IR}, but it is not known whether 4d RG flows are gradient flows [2, 3, 4]. (In four dimensions, the relevant function is often called aa; we still refer to it as a cc-function here for uniformity of notations.)

In both d=2d=2 and d=4d=4, these measures of degrees of freedom can be defined as conformal trace anomaly coefficients. Unfortunately, this type of definition is only limited to even values of dd and, indeed, in d=3d=3, the search for a suitable candidate of a cc-function — that is monotonic along RG flow and is stationary at the fixed points — has been considerably more challenging, see e.g. [5, 6, 7, 8]. That’s why in this paper, we wish to focus on three-dimensional QFT’s, or QFT3 for short.

The reason cc-functions measure the effective number of degrees of freedom is that, at least in d=2d=2 and d=4d=4 where they are better understood, they can be formulated directly in terms of the spectrum of the theory, namely expressing the growth of states. For example, in d=2d=2, if ana_{n} is the number of states with energy nn, then

anexp2π16ceffnasna_{n}\;\sim\;\exp 2\pi\sqrt{\frac{1}{6}c_{\text{eff}}\,n}\qquad\text{as}\quad n\to\infty (1)

for some constant ceffc_{\text{eff}} [9]. Equivalently, it can be formulated as a statement about the behavior of the generating function

χ(q)=nanqn+Δ\chi(q)\;=\;\sum_{n}a_{n}q^{n+\Delta} (2)

where Δ\Delta is the ground state energy. The starting point of this paper is a curious observation that (1) holds in theories one dimension higher, namely in SCFT3.

Conjecture 1.1

In every three-dimensional superconformal field theory with 𝒩=2\mathcal{N}=2 supersymmetry, the spectrum of supersymmetric (BPS) states obeys (1). In other words, the superconformal index or, equivalently, S2×qS1S^{2}\times_{q}S^{1} partition function,

(q):=TrS2[(1)FqR/2+J3]=Z(S2×qS1)=nanqn\mathcal{I}(q):=\Tr_{\mathcal{H}_{S^{2}}}\left[(-1)^{F}q^{R/2+J_{3}}\right]=Z(S^{2}\times_{q}S^{1})=\sum_{n}a_{n}q^{n} (3)

enjoys (1).

Although this behavior has been observed in 3d-3d correspondence (see e.g. [10, 11, 12]), to the best of our knowledge, it has never been proposed as a general property of 3d SCFTs with 𝒩=2\mathcal{N}=2 supersymmetry. What makes this observation interesting is that the number of non-BPS states in a general CFT3 grows much faster. Namely, in CFTd, we have [13]

logannd1d.\log a_{n}\;\sim\;n^{\frac{d-1}{d}}. (4)

In particular, for d=2d=2 the exponent is 12\frac{1}{2} and for d=3d=3 it is 23\frac{2}{3}. Therefore, while the growth of general (non-BPS) states is controlled by n2/3n^{2/3}, the supersymmetric (BPS) states grow only as n1/2n^{1/2} for large values of nn. Based on this observation, we can make the following definition.

Definition 1

Assuming Conjecture 1.1, to any 3d 𝒩=2\mathcal{N}=2 SCFT we associate a quantity ceffc_{\text{eff}} defined via the asymptotic behavior of superconformal index (3):

ceff:=32π2limn(log|an|)2nc_{\text{eff}}\;:=\;\frac{3}{2\pi^{2}}\lim_{n\to\infty}\frac{(\log|a_{n}|)^{2}}{n} (5)

Because the superconformal index is invariant along RG flows, we can’t conjecture that ceffc_{\text{eff}} defined in this way is decreasing along RG flows. Nevertheless, we hope it can still be a useful measure of the number of degrees of freedom in a 3d 𝒩=2\mathcal{N}=2 SCFT.

The reason the absolute value in (5) is used is that, in the supersymmetric setting, the coefficients ana_{n} may not be all positive. After all, they are counting BPS states with signs.111This aspect has important consequences, namely the 3d superconformal index is expected to admit a categorification [14]. In part for this reason, below we will need a slightly refined version of Conjecture 1.1 and Definition 1 of ceffc_{\text{eff}}.

Conjecture 1.2

The growth of supersymmetric (BPS) states in a three-dimensional 𝒩=2\mathcal{N}=2 theory is given by the following asymptotic formula,

anRe[exp(2π23ceffn+2πirn)],a_{n}\sim\mathrm{Re}\left[\exp(\sqrt{\frac{2\pi^{2}}{3}c_{\text{eff}}\hskip 4.26773ptn}+2\pi irn)\right], (6)

where r/r\in{\mathbb{Q}}/{\mathbb{Z}}, and ceffc_{\text{eff}}\in{\mathbb{C}}.

This asymptotic formula for the density of BPS states is very reminiscent of the Cardy formula (1) for the density of states in two-dimensional conformal field theories. In fact, when rr is zero, and ceffc_{\text{eff}} is real and positive, it is exactly the same as the Cardy formula. When rr is non-zero, or when ceffc_{\text{eff}} is a generic complex number, the formula (6) captures two curious features of the density of BPS states that we call “branching” and “oscillations.” When r=k0r=\frac{\ell}{k}\neq 0 with \ell and kk relatively prime integers, the density of BPS states branches into kk branches. On the other hand, when the imaginary part of ceff\sqrt{c_{\text{eff}}} is non-zero, the density of BPS states oscillates. These two curious features can be seen already in one of the simplest examples of three-dimensional 𝒩=2\mathcal{N}=2 SCFTs, namely in the theory of a chiral superfield Φ\Phi with the cubic superpotential W=Φ3W=\Phi^{3}. As illustrated in Figure 1, the coefficients ana_{n} in this theory exhibit oscillations and form three distinct branches.

Refer to caption
Figure 1: Coefficients of the supersymmetric index in the 𝒩=2\mathcal{N}=2 Landau-Ginzburg theory with the superpotential W=Φ3W=\Phi^{3}.

In the original derivation of (1) by Cardy [9], the SL(2,)SL(2,\mathbb{Z}) modular symmetry group of the underlying torus T2T^{2} played a key role. If one had a similar symmetry in three dimensions, the derivation of (6) would be much easier, and we would not need to state it as a conjecture. As mentioned above, 3d superconfomal index (3) can be understood as a supersymmetric partition function on S2×qS1S^{2}\times_{q}S^{1}. And, although it produces a qq-series reminiscent of 2d CFT characters or elliptic genera, it lacks any obvious modular properties because SL(2,)SL(2,\mathbb{Z}) is not a symmetry of S2×qS1S^{2}\times_{q}S^{1}. Nevertheless, the modular group does play a more subtle role in 3d 𝒩=2\mathcal{N}=2 theories,222It controls the spectrum of supersymmetric line operators and twisted partition functions [15] in a way that all such relations are mutually compatible, and also compatible with the modular properties of BPS partition functions [10]. which can be traced to embedding of T2T^{2} into S2×S1S^{2}\times S^{1} as a central surface in a genus-1 Heegaard splitting. A similar property is shared by D2×qS1D^{2}\times_{q}S^{1} supersymmetric partition function, where T2T^{2} is the boundary and its SL(2,)SL(2,\mathbb{Z}) is also manifest.

As a result, some supersymmetric partition functions of 3d 𝒩=2\mathcal{N}=2 theories exhibit a “universality” in the sense that their behavior near roots of unity is controlled by semi-classical quantities such as the twisted superpotential [16]. Here, we summarize this 3d analogue of modularity in the form of the following conjecture that will be useful to us not only in the analysis of the superconformal index (q)\mathcal{I}(q) but also in its close cousins that encode other types of BPS spectra. In other words, our analysis below will assume the following behavior of the relevant supersymmetric partition functions.

Conjecture 1.3

Supersymmetric partition functions have the following asymptotic behavior near q=e2πirq=e^{-2\pi ir} or equivalently near τ=r\tau=-r, with r/r\in{\mathbb{Q}}/{\mathbb{Z}}.

Z(q)exp[1r(𝒲~r(0)+r𝒲~r(1)+r2𝒲~r(2)+)],Z(q)\sim\exp[-\frac{1}{\hbar_{r}}\left(\widetilde{\mathcal{W}}_{r}^{(0)}+\hbar_{r}\widetilde{\mathcal{W}}_{r}^{(1)}+\hbar_{r}^{2}\widetilde{\mathcal{W}}_{r}^{(2)}+\cdots\right)], (7)

where er=qe2πir=e2πi(τ+r)e^{-\hbar_{r}}=qe^{2\pi ir}=e^{2\pi i(\tau+r)}, and 𝒲~r(j)\widetilde{\mathcal{W}}_{r}^{(j)}\in{\mathbb{C}}.

Using this conjecture as an intermediate step, we can derive the density of states in Conjecture 1.2 in much the same way as the Cardy formula (1) is derived in two-dimensional conformal field theories. In other words, as we explain in more detail in the next section, it reduces the analysis to the study of the asymptotic behavior near the natural boundary |q|=1|q|=1.

We verified Conjectures 1.2 and 1.3 in a large variety of 3d 𝒩=2\mathcal{N}=2 theories. In Lagrangian theories, such as Φ3\Phi^{3} theory mentioned earlier, the analysis easily follows the strategy outlined above because indices can be explicitly written as sums of basic ingredients made of qq-Pochhammer symbols, and their asymptotics at roots of unity is well known.

We also considered some 3d 𝒩=2\mathcal{N}=2 theories whose Lagrangian description is not known at present. A large class of such theories comes from 3d-3d correspondence. As argued in [17], one should expect to be able to realize theories T[M3]T[M_{3}] as gauge theories with “non-linear” matter, i.e. as Skyrme type models where 3d 𝒩=2\mathcal{N}=2 chiral multiplets take values in complex group manifolds GG_{{\mathbb{C}}}. At present, such a description is not developed. However, in the context of 3d-3d correspondence, there is another strategy available to us that can be very helpful in approaching Conjectures 1.2 and 1.3. It relies on the fact that, for any 3-manifold M3M_{3}, theories T[M3]T[M_{3}] admit a canonical choice of 2d 𝒩=(0,2)\mathcal{N}=(0,2) boundary conditions labeled by Spin(M3)c{}^{c}(M_{3}). This allows us to reduce the analysis of the superconformal index (q)\mathcal{I}(q) to a (sometimes simpler) analysis of a family of BPS partition functions on D2×qS1D^{2}\times_{q}S^{1} with boundary conditions labeled by Spinc-structures. (The reason this strategy does not admit an obvious extension outside 3d-3d correspondence is that general 3d 𝒩=2\mathcal{N}=2 theories do not have a “canonical” set of 2d 𝒩=(0,2)\mathcal{N}=(0,2) boundary conditions.)

The rest of the paper is organized as follows. In section 2, we imitate Cardy’s derivation of (1) in the context of 3d 𝒩=2\mathcal{N}=2 theories, using (7) as an assumption. We then illustrate general ideas and considerations with a concrete example of Φ3\Phi^{3} theory, whose BPS spectrum already appeared in Figure 1. In section 3, we turn to examples of 3d 𝒩=2\mathcal{N}=2 theories, for which Lagrangian description is not known at present. Although we draw such examples from 3d-3d correspondence, one can probably consider other sources. In all instances, we find evidence for Conjectures 1.2 and 1.3 and compute the corresponding values of ceffc_{\text{eff}}.

2 Indices near the unit circle

Supersymmetric partition functions are expected to be holomorphic functions of qq inside the unit circle. Using this as an assumption, we can express the nn-th coefficient of the qq-series using the Cauchy integral formula. That is for Z(q)=n=0anqnZ(q)=\sum_{n=0}^{\infty}a_{n}q^{n},

an=dq2πiqqnZ(q),a_{n}=\oint\frac{\differential q}{2\pi iq}q^{-n}Z(q), (8)

where the contour encloses q=0q=0. Near the unit circle, Z(q)Z(q) diverges or goes to zero. As we take the contour close to the unit circle, the integral (8) is dominated by saddle points. For example, the integrand that appears in the computation of the 100th100^{\text{th}} coefficient in the W=Φ3W=\Phi^{3} theory has two dominant saddles near q=e±2πi3q=e^{\pm\frac{2\pi i}{3}} (see Figure 2).

Refer to caption
Figure 2: |Φ3(q)q100||\mathcal{I}_{\Phi^{3}}(q)q^{-100}| on the contour q=e2πi(θ+0.01i)q=e^{2\pi i(\theta+0.01i)} with θ(12,12)\theta\in(-\frac{1}{2},\frac{1}{2})

.

Supersymmetric partition functions may have singularities (7) at the roots of unity. This aspect is well studied in the context of 3d-3d correspondence where it plays an important role (see e.g. [18, 19]), and is expected to hold more generally. In fact, for a given 3d 𝒩=2\mathcal{N}=2 theory, different supersymmetric partition functions often exhibit “universality,” i.e. have the same asymptotic behavior [16]. Therefore, we expect the saddle points of the integral to be located near the roots of unity. Suppose we have a dominant saddle near q=e2πirq=e^{-2\pi ir}. Using the asymptotic behavior of supersymmetric partition functions from Conjecture 1.3, we have

anr12+iϵr+12+iϵdτexp[1r(𝒲~r(0)+r(𝒲~r(1)2πirn)+r2(𝒲~r(2)n))]a_{n}\sim\int_{-r-\frac{1}{2}+i\epsilon}^{-r+\frac{1}{2}+i\epsilon}\differential\tau\hskip 2.84544pt\exp[-\frac{1}{\hbar_{r}}\left(\widetilde{\mathcal{W}}_{r}^{(0)}+\hbar_{r}\left(\widetilde{\mathcal{W}}_{r}^{(1)}-2\pi irn\right)+\hbar_{r}^{2}\left(\widetilde{\mathcal{W}}_{r}^{(2)}-n\right)\right)] (9)

Now the integral has saddle points at τ=r±12πW~r(0)nW~r(2)\tau=-r\pm\frac{1}{2\pi}\sqrt{\frac{\widetilde{W}_{r}^{(0)}}{n-\widetilde{W}_{r}^{(2)}}}, and the value of integrand at these saddle point is

exp(2πirnW~r(1)2iW~r(0)nW~r(0)W~r(2)).\exp(2\pi irn-\widetilde{W}_{r}^{(1)}\mp 2i\sqrt{\widetilde{W}_{r}^{(0)}n-\widetilde{W}_{r}^{(0)}\widetilde{W}_{r}^{(2)}}).

As ana_{n} are real numbers, we expect equally dominant saddle points near τ=r\tau=r, which have the same real part but opposite imaginary part. Considering only the most dominant saddles, at large nn we get,

anRe[exp(2π23ceffn+2πirn)],a_{n}\sim\mathrm{Re}\left[\exp(\sqrt{\frac{2\pi^{2}}{3}c_{\text{eff}}\hskip 4.26773ptn}+2\pi irn)\right], (10)

where ceff=6π2W~r(0)c_{\text{eff}}=-\frac{6}{\pi^{2}}\widetilde{W}_{r}^{(0)}.

2.1 Example: W=Φ3W=\Phi^{3} theory

The supersymmetric index of the W=Φ3W=\Phi^{3} theory is given by Φ3(q)=(q2/3,q)(q1/3,q)\mathcal{I}_{\Phi^{3}}(q)=\frac{(q^{2/3},q)_{\infty}}{(q^{1/3},q)_{\infty}}. An important detail is that this index has a cubic branch cut from 0 to 1-1. To avoid working with branch cuts, we change the variables qq3q\rightarrow q^{3}. Now the index can be written as

Φ3(q)=(q2,q3)(q,q3).\mathcal{I}_{\Phi^{3}}(q)=\frac{(q^{2},q^{3})_{\infty}}{(q,q^{3})_{\infty}}. (11)

The index is well-defined inside the unit circle. We can write it as

Φ3(q)=exp(n=0Bnn!(logq3)n1(Li2n(q2)Li2n(q))).\mathcal{I}_{\Phi^{3}}(q)=\exp(\sum_{n=0}^{\infty}\frac{B_{n}}{n!}\left(\log q^{3}\right)^{n-1}(\mathrm{Li}_{2-n}(q^{2})-\mathrm{Li}_{2-n}(q))). (12)

From the equation (12), we can see that the index Φ3(q)\mathcal{I}_{\Phi^{3}}(q) has singularities at the cubic roots of unity. As qq approaches these points, the asymptotic forms of Φ3(q)\mathcal{I}_{\Phi^{3}}(q) looks like

Φ3(q)\displaystyle\mathcal{I}_{\Phi^{3}}(q) q1exp(13log(0)+O(1)),\displaystyle\sim_{q\rightarrow 1}\exp(-\frac{1}{3}\log(\hbar_{0})+O\left(1\right)), (13)
Φ3(q)\displaystyle\mathcal{I}_{\Phi^{3}}(q) qe2πi3exp(Li2(e2πi3)Li2(e4πi3)3h13+O(1)),\displaystyle\sim_{q\rightarrow e^{-\frac{2\pi i}{3}}}\exp(-\frac{\mathrm{Li}_{2}(e^{-\frac{2\pi i}{3}})-\mathrm{Li}_{2}(e^{-\frac{4\pi i}{3}})}{3h_{\frac{1}{3}}}+O(1)), (14)
Φ3(q)\displaystyle\mathcal{I}_{\Phi^{3}}(q) qe2πi3exp(Li2(e2πi3)Li2(e4πi3)3h13+O(1)).\displaystyle\sim_{q\rightarrow e^{\frac{2\pi i}{3}}}\exp(-\frac{\mathrm{Li}_{2}(e^{\frac{2\pi i}{3}})-\mathrm{Li}_{2}(e^{\frac{4\pi i}{3}})}{3h_{-\frac{1}{3}}}+O(1)). (15)

See appendix A for details. As expected from Figure 2, the singularities at q=e±2πi3q=e^{\pm\frac{2\pi i}{3}} are more dominant compared to that at q=1q=1. Therefore, we get dominant contribution from saddle points near q=e±2πi3q=e^{\pm\frac{2\pi i}{3}}. The effective super-potential W~±13(0)\widetilde{W}_{\pm\frac{1}{3}}^{(0)} is,

W~±13(0)=±Li2(e4πi3)Li2(e2πi3)3.\widetilde{W}_{\pm\frac{1}{3}}^{(0)}=\pm\frac{\mathrm{Li}_{2}(e^{\frac{4\pi i}{3}})-\mathrm{Li}_{2}(e^{\frac{2\pi i}{3}})}{3}. (16)

Hence, for the W=Φ3W=\Phi^{3} theory, the asymptotics of coefficients ana_{n} and the effective central charge ceffc_{\text{eff}} are given by

an\displaystyle a_{n} Re[exp(43(Li2(e2πi3)Li2(e4πi3))n+2πin3)],\displaystyle\sim\mathrm{Re}\left[\exp(\sqrt{\frac{4}{3}\left(\mathrm{Li}_{2}(e^{\frac{2\pi i}{3}})-\mathrm{Li}_{2}(e^{\frac{4\pi i}{3}})\right)\hskip 4.26773ptn}+\frac{2\pi in}{3})\right], (17)
ceff\displaystyle c_{\text{eff}} =2π2(Li2(e2πi3)Li2(e4πi3))0.274227i.\displaystyle=\frac{2}{\pi^{2}}\left(\mathrm{Li}_{2}(e^{\frac{2\pi i}{3}})-\mathrm{Li}_{2}(e^{\frac{4\pi i}{3}})\right)\approx 0.274227i. (18)

3 Strongly coupled examples: theories T[M3]T[M_{3}]

For intrinsically strongly coupled 3d 𝒩=2\mathcal{N}=2 theories, we turn to 3d-3d correspondence. This correspondence associates 3d 𝒩=2\mathcal{N}=2 theories to 3-manifolds via compactification of 6d (0,2)(0,2) fivebrane theory on 3-manifolds. For a given choice of a 3-manifold, the resulting 3d 𝒩=2\mathcal{N}=2 theory is usually denoted T[M3]T[M_{3}].333Note, that in early studies of this correspondence, a Lagrangian description for a particular sector of the full theory T[M3]T[M_{3}] was proposed [20], but it will not suffice for our analysis below since abelian flat connections on M3M_{3} are not accounted by this sector of T[M3]T[M_{3}]. See [17, 21] for a relatively recent account of some of these issues and a more precise characterization of the sector of T[M3]T[M_{3}] described by the “DGG theory.” In this context, a large supply of 3d 𝒩=2\mathcal{N}=2 strongly coupled theories comes from the rich world of 3-manifolds. Suppressed in the notation T[M3]T[M_{3}] is also a choice of the root system, i.e. a choice of 6d (0,2)(0,2) theory. In this paper, we focus mostly on the simplest non-trivial case of G=SU(2)G=SU(2) (or, equivalently, G=SL(2,)G_{{\mathbb{C}}}=SL(2,{\mathbb{C}})) that corresponds to two fivebranes in M-theory realization of 3d-3d correspondence. It would be very interesting to extend the analysis below to higher rank version of 3d-3d correspondence and to root systems of Cartan type DD and EE.

Moreover, for 3d 𝒩=2\mathcal{N}=2 theories that come from 3d-3d correspondence, we have two useful properties that we can leverage in testing Conjectures 1.2 and 1.3:

  • First, we can relate the superconformal index (q)\mathcal{I}(q) to another BPS partition function that was actively studied in recent years, the so-called half-index or partition function on D2×qS1D^{2}\times_{q}S^{1}, where D2D^{2} can be understood as the cigar geometry [16]:

    Z^a(q)=Z(D2×qS1,a).\widehat{Z}_{a}(q)=Z(D^{2}\times_{q}S^{1},\mathcal{B}_{a}). (19)

    This partition function depends on a choice of 2d 𝒩=(0,2)\mathcal{N}=(0,2) boundary condition a\mathcal{B}_{a}, and can be viewed as the elliptic genus of 2d-3d coupled system. Indeed, if 3d 𝒩=2\mathcal{N}=2 theory is gapped (or, gappable), then it is just the usual elliptic genus of the 2d 𝒩=(0,2)\mathcal{N}=(0,2) boundary theory a\mathcal{B}_{a}. The physics of such coupled 2d-3d systems is very rich and has been an area of active research in recent years. In particular, as we already mentioned above, in the context of 3d-3d correspondence, theories T[M3]T[M_{3}] come equipped with a “canonical” set of boundary conditions a\mathcal{B}_{a} labeled by aSpinc(M3)a\in\text{Spin}^{c}(M_{3}). The significance of this insignificant-sounding fact is that (q)\mathcal{I}(q) can be expressed as a sum of more basic building blocks [14]:

    T[M3](q)=a|𝒲a|ZT[M3](D2×qS1,a)ZT[M3¯](D2×qS1,a),\mathcal{I}_{T[M_{3}]}(q)=\sum_{a}|\mathcal{W}_{a}|\,Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{a})\,Z_{T[\overline{M_{3}}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{a}), (20)

    where the line over M3M_{3} indicates that M3¯\overline{M_{3}} and M3M_{3} have opposite orientations, and

    ZT[M3](D2×qS1,a)=Z^a(M3,q)η(q)Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{a})=\frac{\widehat{Z}_{a}(M_{3},q)}{\eta(q)} (21)

    is the normalized version of (19). (For higher-rank version of 3d-3d correspondence, this normalization would involve a factor η(q)rank(G)\eta(q)^{\text{rank}(G)}.)

  • The second feature of 3d 𝒩=2\mathcal{N}=2 theories T[M3]T[M_{3}] is that, by the very nature of 3d-3d correspondence, physical properties and observables in T[M3]T[M_{3}] are topological invariants of M3M_{3}. In particular, using the relation (21) and the topological interpretation of Z^a(q)\widehat{Z}_{a}(q) as a qq-series invariant that provides the non-perturbative definition for complex Chern-Simons theory on M3M_{3}, we will be able to propose an explicit expression for ceffc_{\text{eff}} in terms of topological invariants of M3M_{3}. The key element is the trans-series expansion of Z^a(q)\widehat{Z}_{a}(q):

    Z^(q)\displaystyle\widehat{Z}(q) =πiαπ0(flat[M3,SL(2,)])×nα,0e4π2SαZαPert(),\displaystyle=\sqrt{\frac{\pi i}{\hbar}}\sum_{\mathbb{\bbalpha}\in\pi_{0}(\mathcal{M}_{\text{flat}}[M_{3},SL(2,{\mathbb{C}})])\times{\mathbb{Z}}}n_{\mathbb{\bbalpha},0}\,e^{\frac{4\pi^{2}}{\hbar}S_{\mathbb{\bbalpha}}}Z_{\alpha}^{\text{Pert}}(\hbar),
    =πiαπ0(flat[M3,SL(2,)])mn(α,m),0q~CS(α)+mZαPert(),\displaystyle=\sqrt{\frac{\pi i}{\hbar}}\sum_{\alpha\in\pi_{0}(\mathcal{M}_{\text{flat}}[M_{3},SL(2,{\mathbb{C}})])}\sum_{m\in{\mathbb{Z}}}n_{(\alpha,m),0}\hskip 2.84544pt\widetilde{q}^{-\mathrm{CS}(\alpha)+m}Z_{\alpha}^{\text{Pert}}(\hbar), (22)

    writen here for homology spheres and for G=SL(2,)G_{\mathbb{C}}=SL(2,{\mathbb{C}}). The derivation of this expansion and its generalization to other 3-manifolds can be found e.g. in [14, 22]. The notations used in (22) and their application to the computation of ceffc_{\text{eff}} will be explained shortly.

Since 3d-3d correspondence tells us [23] that the values of twisted superpotential of T[M3]T[M_{3}] at its critical points are the SL(2,)SL(2,{\mathbb{C}}) Chern-Simons values of M3M_{3}, one might expect a priori that the value of ceffc_{\text{eff}} for T[M3]T[M_{3}] is related to the Chern-Simons values of M3M_{3}. However, for a given M3M_{3}, there are, in general, many complex flat connections, but there is only one value of ceffc_{\text{eff}}. Therefore, a natural question is: If this reasoning is on the right track, which particular Chern-Simons value of M3M_{3} determines the value of ceffc_{\text{eff}} for a given T[M3]T[M_{3}]? Or, is it a certain combination of Chern-Simons values? If so, which combination? And, what is the topological significance of complex flat connections that contribute to ceffc_{\text{eff}}?

All these questions can be answered with the help of the resurgent analysis, which among other things, leads to (22). So, let us discuss this aspect in more detail, starting with the notations. As in [22], we labeled (connected components of the space of) complex flat connections on M3M_{3} by

απ0(flat[M3,SL(2,)])\alpha\in\pi_{0}\left(\mathcal{M}_{\text{flat}}\left[M_{3},SL(2,{\mathbb{C}})\right]\right) (23)

and introduce a notation

απ0(flat[M3,SL(2,)])×\mathbb{\bbalpha}\in\pi_{0}\left(\mathcal{M}_{\text{flat}}\left[M_{3},SL(2,{\mathbb{C}})\right]\right)\times{\mathbb{Z}} (24)

for their integral lift. In other words, α\alpha is an equivalence class of α\mathbb{\bbalpha}. We follow a general rule that letters from the Greek alphabet denote generic flat connections, whereas letters from the Latin alphabet refer to abelian flat connections. The Chern-Simons value of any SL(2,)SL(2,{\mathbb{C}}) flat connection is only defined modulo 11, that is CS(α)/\mathrm{CS}(\alpha)\in{\mathbb{C}}/{\mathbb{Z}}. However, for the lift α\mathbb{\bbalpha} of α\alpha, we can define the Chern-Simons value, SαS_{\mathbb{\bbalpha}}, valued in {\mathbb{C}}, such that Sα=CS(α)mod1S_{\mathbb{\bbalpha}}=\mathrm{CS}(\alpha)\mod 1.

The coupling constant in (non-perturbative) complex Charn-Simons theory is a continuous complex parameter qq, the same variable that appears in the superconformal index (3) of the 3d 𝒩=2\mathcal{N}=2 theory T[M3]T[M_{3}]. It is related to the perturbative coupling constant \hbar in the formal power series ZαPert()Z_{\alpha}^{\text{Pert}}(\hbar) via q=eq=e^{-\hbar}. In comparison to analytic continuation from Chern-Simons theory with compact gauge group, it is also sometimes useful to keep in mind the relation to “level” kk, namely q=e=e2πikq=e^{-\hbar}=e^{\frac{2\pi i}{k}}. And another standard notation from complex Chern-Simons theory that we will need is q~=e4π2/\widetilde{q}=e^{-4\pi^{2}/\hbar}. For convenience, we summarize these variables here

q=e,q~=e4π2/q=e^{-\hbar}\qquad,\qquad\widetilde{q}=e^{-4\pi^{2}/\hbar} (25)

The last and, arguably, the most delicate ingredient in the trans-series expansion (22) is the set of trans-series coefficients nα,0n_{\mathbb{\bbalpha},0}. At present, no general systematic way for computing nα,0n_{\mathbb{\bbalpha},0} is known, unlike Chern-Simons values that can be computed e.g. from surgery presentations of M3M_{3} and in multiple other ways. Therefore, in our general analysis below, we will not make any assumptions about the values of these coefficients, and when it comes to explicit calculations in concrete examples, we determine their values either from modular properties or numerically.

Now we are ready to combine (20), (21), and (22) to explore the implications. Suppose that M3M_{3} is a homology sphere, and suppose that the most dominant contribution to (22) near q=1q=1 (that is, 0\hbar\approx 0) comes from the term e4π2Sαe^{\frac{4\pi^{2}}{\hbar}S_{\mathbb{\bbalpha}}}. Then, near q=e1q=e^{-\hbar}\approx 1, the asymptotics of ZT[M3](D2×qS1,0)Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}) is given by

ZT[M3](D2×qS1,0)=Z^(q)η(q)exp[π26(1+24Sβ)].Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0})=\frac{\widehat{Z}(q)}{\eta(q)}\sim\exp[\frac{\pi^{2}}{6\hbar}\left(1+24S_{\mathbb{\bbbeta}}\right)]. (26)

Therefore, the growth of the coefficients in ZT[M3](D2×qS1,0)Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}), and ceff,12T[M3]c_{\text{eff},\frac{1}{2}}^{T[M_{3}]} for the half index are given by

an\displaystyle a_{n}\sim exp[2π23[1+24Sβ]n],\displaystyle\exp[\sqrt{\frac{2\pi^{2}}{3}\left[1+24S_{\mathbb{\bbbeta}}\right]n}], ceff,12T[M3]\displaystyle c_{\text{eff},\frac{1}{2}}^{T[M_{3}]} =1+24Sβ.\displaystyle=1+24S_{\mathbb{\bbbeta}}. (27)

Just as we expressed the asymptotic behavior of ZT[M3](D2×qS1,0)Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}), we can write the asymptotic behavior of ZT[M3¯](D2×qS1,0)Z_{T[\overline{M_{3}}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}) for M3¯\overline{M_{3}}, (M3M_{3} with orientation reversal),

ZT[M3¯](D2×qS1,0)exp[π26(1+24Sβ)].Z_{T[\overline{M_{3}}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0})\sim\exp[\frac{\pi^{2}}{6\hbar}\left(1+24S_{\mathbb{\bbbeta}^{\prime}}\right)]. (28)

Then, the asymptotic behavior of the index is given by

T[M3]exp[π26(2+24Sβ+24Sβ)].\mathcal{I}_{T[M_{3}]}\sim\exp[\frac{\pi^{2}}{6\hbar}\left(2+24S_{\mathbb{\bbbeta}^{\prime}}+24S_{\mathbb{\bbbeta}}\right)]. (29)

and ceffc_{\text{eff}} for T[M3]T[M_{3}] comes out to be

ceff= 2+24Sβ+24Sβ.c_{\text{eff}}\;=\;2+24S_{\mathbb{\bbbeta}}+24S_{\mathbb{\bbbeta}^{\prime}}. (30)

Given that M3M_{3} and M3¯\overline{M_{3}} differ by the orientation reversal, one might expect SβS_{\mathbb{\bbbeta}} and SβS_{\mathbb{\bbbeta}^{\prime}} to cancel each other, resulting in ceff=2c_{\text{eff}}=2. However, as we will see shortly, this does not happen, even in simple examples. Moreover, it is important to stress that ceffc_{\text{eff}} is determined not only by Chern-Simons values, but also by trans-series coefficients or, equivalently, by values of β\mathbb{\bbbeta} and β\mathbb{\bbbeta}^{\prime} that appear in (30). It is this latter data that is hard to determine a priori, and which certainly deserves to be studied more systematically.

3.1 Examples: M3=Σ(2,3,6±1)M_{3}=\Sigma(2,3,6\pm 1)

Let us consider an example of T[M3]T[M_{3}] for M3=Σ(2,3,5)M_{3}=\Sigma(2,3,5). The three-manifold Σ(2,3,5)\Sigma(2,3,5) has three SL(2,)SL(2,{\mathbb{C}}) flat connections, one abelian connection, and two non-abelian connections. We denote the abelian connection by α0\alpha_{0} and the two non-abelian connections by α1\alpha_{1} and α2\alpha_{2}. The Chern-Simons value of the abelian connection is zero, while the Chern-Simons values of non-abelian connections are given by

CS(α1)\displaystyle\mathrm{CS}(\alpha_{1}) =1120mod1\displaystyle=-\frac{1}{120}\mod 1 CS(α1)\displaystyle\mathrm{CS}(\alpha_{1}) =49120mod1.\displaystyle=-\frac{49}{120}\mod 1. (31)

For the three-manifold Σ(2,3,5)¯\overline{\Sigma(2,3,5)} with opposite orientation, we have the corresponding counterparts of these three connections, which we denote by α0,α1\alpha_{0}^{\prime},\alpha_{1}^{\prime}, and α2\alpha_{2}^{\prime}. Since Σ(2,3,5)¯\overline{\Sigma(2,3,5)} has opposite orientation compared to that of Σ(2,3,5)\Sigma(2,3,5), the Chern-Simons values of α0,α1\alpha_{0}^{\prime},\alpha_{1}^{\prime}, and α2\alpha_{2}^{\prime} are also opposite:

CS(α0)\displaystyle\mathrm{CS}(\alpha_{0}^{\prime}) =0mod1\displaystyle=0\mod 1 CS(α1)\displaystyle\mathrm{CS}(\alpha_{1}^{\prime}) =1120mod1\displaystyle=\frac{1}{120}\mod 1 CS(α1)\displaystyle\mathrm{CS}(\alpha_{1}) =49120mod1.\displaystyle=\frac{49}{120}\mod 1. (32)

The Z^(q)\widehat{Z}(q) invariant for M3=Σ(2,3,5)¯M_{3}=\overline{\Sigma(2,3,5)} is related to the order-five mock theta function χ0(q)\chi_{0}(q) in the following way,

Z^(Σ(2,3,5)¯;q)=q32χ0(q)=q32n=0qn(qn+1)n.\widehat{Z}(\overline{\Sigma(2,3,5)};q)=q^{\frac{3}{2}}\chi_{0}(q)=q^{\frac{3}{2}}\sum_{n=0}^{\infty}\frac{q^{n}}{(q^{n+1})_{n}}. (33)

The modular properties of this mock theta function have been studied in [24, 25]. The modular transform of χ0(q)\chi_{0}(q) is given by,

q1120(χ0(q)2)=\displaystyle q^{-\frac{1}{120}}(\chi_{0}(q)-2)= π555q~1120(χ0(q~)2)π5+55q~49120+1χ1(q~)\displaystyle-\sqrt{\frac{\pi}{\hbar}}\sqrt{\frac{5-\sqrt{5}}{5}}\widetilde{q}^{-\frac{1}{120}}(\chi_{0}(\widetilde{q})-2)-\sqrt{\frac{\pi}{\hbar}}\sqrt{\frac{5+\sqrt{5}}{5}}\widetilde{q}^{-\frac{49}{120}+1}\chi_{1}(\widetilde{q})
1352πJ(15,5),\displaystyle\hskip 28.45274pt-\sqrt{\frac{135\hbar}{2\pi}}J\left(\frac{1}{5},5\hbar\right), (34)

where q~=e4π2/\widetilde{q}=e^{-4\pi^{2}/\hbar}, χ1(q)\chi_{1}(q) is another order-five mock theta function given by

χ1(q)=n=0qn(qn+1)n+1,\chi_{1}(q)=\sum_{n=0}^{\infty}\frac{q^{n}}{(q^{n+1})_{n+1}}, (35)

and the function J(r,γ)J(r,\gamma) is defined as

J(r,γ)=0e32γx2cosh[(3r2)γx]+cosh[(3r1)γx]cosh[32γx]dx.J(r,\gamma)=\int_{0}^{\infty}e^{-\frac{3}{2}\gamma x^{2}}\frac{\cosh[(3r-2)\gamma x]+\cosh[(3r-1)\gamma x]}{\cosh[\frac{3}{2}\gamma x]}\differential x. (36)

Using the mock-modularity of χ0(q)\chi_{0}(q), we can write down the mock-modular properties of Z^(Σ(2,3,5)¯;q)\widehat{Z}(\overline{\Sigma(2,3,5)};q) as follows,

Z^(Σ(2,3,5)¯;q)=πin=0a~1(n,q)q~1120+n+πin=1a~2(n,q)q~49120+n+Z0Pert(),\widehat{Z}(\overline{\Sigma(2,3,5)};q)=\sqrt{\frac{\pi i}{\hbar}}\sum_{n=0}^{\infty}\widetilde{a}_{1}(n,q)\widetilde{q}^{-\frac{1}{120}+n}+\sqrt{\frac{\pi i}{\hbar}}\sum_{n=1}^{\infty}\widetilde{a}_{2}(n,q)\widetilde{q}^{-\frac{49}{120}+n}+Z_{0}^{\text{Pert}}(\hbar), (37)

where Z0Pert()Z_{0}^{\text{Pert}}(\hbar) is,

Z0Pert()=2q321352πq181120J(15,5),Z_{0}^{\text{Pert}}(\hbar)=2q^{\frac{3}{2}}-\sqrt{\frac{135\hbar}{2\pi}}q^{\frac{181}{120}}J\left(\frac{1}{5},5\hbar\right), (38)

and the coefficients a~1(n,q)\widetilde{a}_{1}(n,q), and a~2(n,q)\widetilde{a}_{2}(n,q) are,

a~1(n,q)\displaystyle\widetilde{a}_{1}(n,q) =iq181120i555a1(n)\displaystyle=iq^{\frac{181}{120}}\sqrt{i\frac{5-\sqrt{5}}{5}}a_{1}(n) a~2(n,q)\displaystyle\widetilde{a}_{2}(n,q) =iq181120i5+55a2(n)\displaystyle=iq^{\frac{181}{120}}\sqrt{i\frac{5+\sqrt{5}}{5}}a_{2}(n) (39)

where a1(n)a_{1}(n) and a2(n)a_{2}(n) are nn-th coefficients of the qq-series (χ0(q)2)(\chi_{0}(q)-2) and χ1(q)\chi_{1}(q), respectively. Notice that the first sum in equation (37) starts from n=0n=0, while the second sum starts from n=1n=1. From this, we can deduce that the most dominant singularity of Z^(Σ(2,3,5)¯;q)\widehat{Z}(\overline{\Sigma(2,3,5)};q) on the unit circle is at q=1q=1 and is due to the term a~1(0,q)q~1120\widetilde{a}_{1}(0,q)\widetilde{q}^{-\frac{1}{120}}. The asymptotic behavior of the half-index ZT[Σ(2,3,5)¯](D2×qS1,0)Z_{T[\overline{\Sigma(2,3,5)}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}) near q=1q=1 is,

ZT[Σ(2,3,5)¯](D2×qS1,0)q~1241120.Z_{T[\overline{\Sigma(2,3,5)}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0})\sim\widetilde{q}^{-\frac{1}{24}-\frac{1}{120}}. (40)

Therefore, ceffc_{\text{eff}} for the half-index ZT[M3](D2×qS1,0)Z_{T[M_{3}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}), with M3=Σ(2,3,5)¯M_{3}=\overline{\Sigma(2,3,5)}, is

ceff,12T[Σ(2,3,5)¯]=65.c_{\text{eff},\frac{1}{2}}^{T[\overline{\Sigma(2,3,5)}]}=\frac{6}{5}. (41)

The Z^(q)\widehat{Z}(q) invariant for M3=Σ(2,3,5)M_{3}=\Sigma(2,3,5) is given by

Z^(Σ(2,3,5),q)=q1n=0(1)nqn(3n1)2(qn+1)n=q1(2q1120Ψ30(1)+(11)+(19)+(29)(q)),\widehat{Z}(\Sigma(2,3,5),q)=q^{-1}\sum_{n=0}^{\infty}\frac{(-1)^{n}q^{\frac{n(3n-1)}{2}}}{(q^{n+1})_{n}}=q^{-1}\left(2-q^{-\frac{1}{120}}\Psi_{30}^{(1)+(11)+(19)+(29)}(q)\right), (42)

where Ψ30(1)+(11)+(19)+(29)(q)\Psi_{30}^{(1)+(11)+(19)+(29)}(q) is a false theta function. Here we have used a shorthand notation

Ψp(a1)+(a2)+(q)=Ψp(a1)(q)+Ψp(a2)(q)+.\Psi_{p}^{(a_{1})+(a_{2})+\cdots}(q)=\Psi_{p}^{(a_{1})}(q)+\Psi_{p}^{(a_{2})}(q)+\cdots. (43)

The false theta functions Ψp(a)\Psi_{p}^{(a)} are given by

Ψp(a)(q)=n2p+asign(n)qn24p.\Psi_{p}^{(a)}(q)=\sum_{n\in 2p{\mathbb{Z}}+a}\mathrm{sign}(n)q^{\frac{n^{2}}{4p}}. (44)

The modular transform of the false theta function Ψp(a)(q)\Psi_{p}^{(a)}(q) is given by

Ψp(a)(q)=2πb=1p1MabΨp(b)(q~)+n=0cnn(14p)n,\Psi_{p}^{(a)}(q)=-\sqrt{-\frac{2\pi}{\hbar}}\sum_{b=1}^{p-1}M_{ab}\Psi_{p}^{(b)}(\widetilde{q})+\sum_{n=0}^{\infty}c_{n}\hbar^{n}\left(\frac{-1}{4p}\right)^{n}, (45)

where n=0cnn(14p)n\sum_{n=0}^{\infty}c_{n}\hbar^{n}\left(\frac{-1}{4p}\right)^{n} is given by the perturbative expansion of partition function of Chern-Simons around the abelian flat connection, and the matrix MabM_{ab} is given by

Mab=2psin(πabp).M_{ab}=\sqrt{\frac{2}{p}}\sin{\frac{\pi ab}{p}}. (46)

Using the false-modularity of false theta functions, we get the modular transform of the Z^(Σ(2,3,5),q)\widehat{Z}(\Sigma(2,3,5),q),

Z^(Σ(2,3,5),q)=πin=0b~1(n,q)q~1120+n+πin=0b~2(n,q)q~49120+n+Z0Pert(),\widehat{Z}(\Sigma(2,3,5),q)=\sqrt{\frac{\pi i}{\hbar}}\sum_{n=0}^{\infty}\widetilde{b}_{1}(n,q)\widetilde{q}^{\frac{1}{120}+n}+\sqrt{\frac{\pi i}{\hbar}}\sum_{n=0}^{\infty}\widetilde{b}_{2}(n,q)\widetilde{q}^{\frac{49}{120}+n}+Z_{0}^{\text{Pert}}(\hbar), (47)

where the coefficients b~1(n,q)\widetilde{b}_{1}(n,q) and b~2(n,q)\widetilde{b}_{2}(n,q) are,

b~1(n,q)\displaystyle\widetilde{b}_{1}(n,q) =iq121120555ib1(n),\displaystyle=-iq^{-\frac{121}{120}}\sqrt{\frac{5-\sqrt{5}}{5i}}b_{1}(n), b~1(n,q)\displaystyle\widetilde{b}_{1}(n,q) =iq1211205+55ib2(n),\displaystyle=-iq^{-\frac{121}{120}}\sqrt{\frac{5+\sqrt{5}}{5i}}b_{2}(n), (48)

where b1(n)b_{1}(n), and b2(n)b_{2}(n) are the coefficients of the qq-series q1120Ψ30(1)+(11)+(19)+(29)(q)q^{-\frac{1}{120}}\Psi_{30}^{(1)+(11)+(19)+(29)}(q), and q49120Ψ30(7)+(13)+(17)+(23)(q)q^{-\frac{49}{120}}\Psi_{30}^{(7)+(13)+(17)+(23)}(q), respectively. The first few terms of these qq-series are

q1120Ψ30(1)+(11)+(19)+(29)(q)\displaystyle q^{-\frac{1}{120}}\Psi_{30}^{(1)+(11)+(19)+(29)}(q) =1+q+q3+q7q8+,\displaystyle=1+q+q^{3}+q^{7}-q^{8}+\cdots, (49)
q49120Ψ30(7)+(13)+(17)+(23)(q)\displaystyle q^{-\frac{49}{120}}\Psi_{30}^{(7)+(13)+(17)+(23)}(q) =1+q+q2+q4q11+.\displaystyle=1+q+q^{2}+q^{4}-q^{11}+\cdots. (50)

Therefore, the dominant term governing the asymptotic behavior of Z^(Σ(2,3,5),q)\widehat{Z}(\Sigma(2,3,5),q) near q=1q=1 is Z0Pert()Z_{0}^{\text{Pert}}(\hbar). Thus, the asymptotic behavior of the half-index ZT[Σ(2,3,5)](D2×qS1,0)Z_{T[\Sigma(2,3,5)]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}) near q=1q=1 is,

ZT[Σ(2,3,5)](D2×qS1,0)q~124,Z_{T[\Sigma(2,3,5)]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0})\sim\widetilde{q}^{-\frac{1}{24}}, (51)

and the asymptotic behavior of the index T[Σ(2,3,5)]\mathcal{I}_{T[\Sigma(2,3,5)]} near q=1q=1 is,

T[Σ(2,3,5)](q)q~11120.\mathcal{I}_{T[\Sigma(2,3,5)]}(q)\sim\widetilde{q}^{-\frac{11}{120}}. (52)

Therefore the ceffc_{\text{eff}} of the index of T[Σ(2,3,5)]T[\Sigma(2,3,5)] is

ceffT[Σ(2,3,5)]=115.c_{\text{eff}}^{T[\Sigma(2,3,5)]}=\frac{11}{5}. (53)

The Z^(q)\widehat{Z}(q) invariant for Σ(2,3,7)¯\overline{\Sigma(2,3,7)} is related to the order-seven mock theta function 0(q)\mathcal{F}_{0}(q) as follows,

Z^(Σ(2,3,7)¯;q)=q120(q)=q12n=0qn2(qn+1)n.\widehat{Z}(\overline{\Sigma(2,3,7)};q)=q^{-\frac{1}{2}}\mathcal{F}_{0}(q)=q^{-\frac{1}{2}}\sum_{n=0}^{\infty}\frac{q^{n^{2}}}{(q^{n+1})_{n}}. (54)

Just as we used the mock-modularity of order-five mock theta function χ0(q)\chi_{0}(q) to determine ceffc_{\text{eff}} for half-index of T[Σ(2,3,5)¯]T[\overline{\Sigma(2,3,5)}]. We can use the mock-modularity of 0(q)\mathcal{F}_{0}(q) to get the asymptotic behavior of the half-index of T[Σ(2,3,7)¯]T[\overline{\Sigma(2,3,7)}] near q=1q=1 and the ceffc_{\text{eff}} for half-index of T[Σ(2,3,7)¯]T[\overline{\Sigma(2,3,7)}]. They are given by

ZT[Σ(2,3,7)¯](D2×qS1,0)\displaystyle Z_{T[\overline{\Sigma(2,3,7)}]}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}) q~121,\displaystyle\sim\widetilde{q}^{-\frac{1}{21}}, ceff,12T[Σ(2,3,7)¯]\displaystyle c_{\text{eff},\frac{1}{2}}^{T[\overline{\Sigma(2,3,7)}]} =87.\displaystyle=\frac{8}{7}. (55)

Just as for Σ(2,3,5)\Sigma(2,3,5), the Z^\widehat{Z}-invariant of Σ(2,3,7)\Sigma(2,3,7) is a linear combination of the false theta functions Ψ168(a)\Psi_{168}^{(a)}. The asymptotic behavior of Z^(Σ(2,3,7),q)\widehat{Z}(\Sigma(2,3,7),q) near q=1q=1 is governed by Z0Pert()Z_{0}^{\text{Pert}}(\hbar). Thus the asymptotic behavior of the index of T[Σ(2,3,7)]T[\Sigma(2,3,7)] near q=1q=1 and the ceffc_{\text{eff}} of the index of T[Σ(2,3,7)]T[\Sigma(2,3,7)] are given by

T[Σ(2,3,7)](q)\displaystyle\mathcal{I}_{T[\Sigma(2,3,7)]}(q) q~556,\displaystyle\sim\widetilde{q}^{-\frac{5}{56}}, ceffT[Σ(2,3,7)]\displaystyle c_{\text{eff}}^{T[\Sigma(2,3,7)]} =157.\displaystyle=\frac{15}{7}. (56)

3.2 Numerical estimates for T[Σ(s,t,st1)]T[\Sigma(s,t,st-1)]

Small surgeries on torus knots give us a special class of Brieskorn homology spheres:

S1r3(T(s,t))\displaystyle S^{3}_{\frac{1}{r}}\left(T(s,t)\right) =Σ(s,t,rst1)¯\displaystyle=\overline{\Sigma(s,t,rst-1)} S1r3(T(s,t))\displaystyle S^{3}_{\frac{1}{r}}\left(T(s,-t)\right) =Σ(s,t,rst+1)¯\displaystyle=\overline{\Sigma(s,t,rst+1)}
S1r3(T(s,t))\displaystyle S^{3}_{-\frac{1}{r}}\left(T(s,-t)\right) =Σ(s,t,rst1)\displaystyle=\Sigma(s,t,rst-1) S1r3(T(s,t))\displaystyle S^{3}_{-\frac{1}{r}}\left(T(s,t)\right) =Σ(s,t,rst+1).\displaystyle=\Sigma(s,t,rst+1).

Using the regularised surgery formula from [26] and the two-variable series, FK(x,q)F_{K}(x,q), for torus knots from [27], we can write down closed form formulae for Z^(Σ(s,t,rst±1),q)\widehat{Z}(\Sigma(s,t,rst\pm 1),q) and Z^(Σ(s,t,rst±1)¯,q)\widehat{Z}(\overline{\Sigma(s,t,rst\pm 1)},q) as follows:

Z^(Σ(s,t,rst±1),q)\displaystyle\widehat{Z}(\Sigma(s,t,rst\pm 1),q) =qr+r14j=0ε2j+1q±j(j+1)st±(t21)(s21)4st(qr(j+1212r)2qr(j+12+12r)2),\displaystyle=q^{-\frac{r+r^{-1}}{4}}\sum_{j=0}^{\infty}\varepsilon_{2j+1}q^{\pm\frac{j(j+1)}{st}\pm\frac{(t^{2}-1)(s^{2}-1)}{4st}}\left(q^{r(j+\frac{1}{2}-\frac{1}{2r})^{2}}-q^{r(j+\frac{1}{2}+\frac{1}{2r})^{2}}\right), (57)
Z^(Σ(s,t,rst±1)¯,q)\displaystyle\widehat{Z}(\overline{\Sigma(s,t,rst\pm 1)},q) =qr+r14j=0ε2j+1qj(j+1)st(t21)(s21)4st(qr(j+1212r)2qr(j+12+12r)2)\displaystyle=q^{\frac{r+r^{-1}}{4}}\sum_{j=0}^{\infty}\varepsilon_{2j+1}q^{\mp\frac{j(j+1)}{st}\mp\frac{(t^{2}-1)(s^{2}-1)}{4st}}\left(q^{-r(j+\frac{1}{2}-\frac{1}{2r})^{2}}-q^{-r(j+\frac{1}{2}+\frac{1}{2r})^{2}}\right) (58)
×(1|k|j(1)kqk((2r+1)k+1)2(qr,q2r+1)(qr+1,q2r+1)(q2r+1,q2r+1)).\displaystyle\hskip 85.35826pt\times\left(1-\frac{\sum_{|k|\leq j}(-1)^{k}q^{\frac{k((2r+1)k+1)}{2}}}{(q^{r},q^{2r+1})_{\infty}(q^{r+1},q^{2r+1})_{\infty}(q^{2r+1},q^{2r+1})_{\infty}}\right). (59)

The qq-series, Z^(Σ(s,t,rst±1),q)\widehat{Z}(\Sigma(s,t,rst\pm 1),q) has coefficients valued in {±1,0}\{\pm 1,0\}. It is a linear combination of false theta functions Ψp(a)(q)\Psi_{p}^{(a)}(q). Similar to Σ(2,3,5)\Sigma(2,3,5), the asymptotics of Z^(Σ(s,t,rst±1),q)\widehat{Z}(\Sigma(s,t,rst\pm 1),q) near q=1q=1 are dominated by Z0PertZ_{0}^{\text{Pert}}. Therefore, the asymptotic behavior of ZT[Σ(s,t,rst±1)(D2×qS1,0)Z_{T[\Sigma(s,t,rst\pm 1)}(D^{2}\times_{q}S^{1},\mathcal{B}_{0}) near q=1q=1 is,

ZT[Σ(s,t,rst±1)(D2×qS1,0)q~124.Z_{T[\Sigma(s,t,rst\pm 1)}(D^{2}\times_{q}S^{1},\mathcal{B}_{0})\sim\widetilde{q}^{-\frac{1}{24}}. (60)

On the other hand, the coefficients of the qq-series, Z^(Σ(s,t,rst±1)¯,q)\widehat{Z}(\overline{\Sigma(s,t,rst\pm 1)},q) grow rapidly. From the trans-series representation of Z^(q)\widehat{Z}(q), we expect the asymptotic behavior of Z^(Σ(s,t,rst±1)¯,q)\widehat{Z}(\overline{\Sigma(s,t,rst\pm 1)},q) near q=1q=1 to be of the form, Z^(Σ(s,t,rst±1)¯,q)q~Sα\widehat{Z}(\overline{\Sigma(s,t,rst\pm 1)},q)\sim\widetilde{q}^{-S_{\mathbb{\bbalpha}}}, for some non-abelian flat connection α\alpha. The Chern-Simons values of flat connections on Σ(s,t,rst±1)¯\overline{\Sigma(s,t,rst\pm 1)} are of the form

CS(αn)=m24st(rst±1)mod1,m.\mathrm{CS}(\alpha_{n})=\frac{m^{2}}{4st(rst\pm 1)}\mod 1,\hskip 56.9055ptm\in{\mathbb{Z}}. (61)

Thus, the expected form of trans-series representation of Z^(q)\widehat{Z}(q) tells us that asymptotically, near q=1q=1,

Z^(Σ(s,t,rst±1)¯,q)q~m24st(rst±1)+,\widehat{Z}(\overline{\Sigma(s,t,rst\pm 1)},q)\sim\widetilde{q}^{-\frac{m^{2}}{4st(rst\pm 1)}+\ell}, (62)

for some integers m,m,\ell. This, in turn, tells us that the asymptotic coefficients of the qq-series are

an(r,s,t)exp[bn]exp[16π2(m24st(rst±1))n]a_{n}(r,s,t)\sim\exp[\sqrt{bn}]\sim\exp[\sqrt{16\pi^{2}\left(\frac{m^{2}}{4st(rst\pm 1)}-\ell\right)n}] (63)
Refer to caption
(a) T(3,t=3x+1)T(3,t=3x+1)
Refer to caption
(b) T(3,t=3x+2)T(3,t=3x+2)
Refer to caption
(c) T(2,t=2x+3)T(2,t=2x+3)
Figure 3: Numerical estimates for bb for torus knots of the form T(2,t=2x+1)T(2,t=2x+1), T(3,t=3x+1)T(3,t=3x+1), and T(3,t=3x+2)T(3,t=3x+2).

One needs the precise modular properties or, equivalently, the trans-series coefficients in order to determine the integers mm and \ell. We use the closed form formula (58) and study the growth of coefficients an(s,t)a_{n}(s,t) to get numerical estimates for the integers mm and \ell for +1+1 surgery on various torus knots. In all the examples of torus knots we looked at, the integer \ell is zero. We have plotted the values of bb for torus knots of the form T(2,t=2x+3)T(2,t=2x+3), T(3,t=3x+1)T(3,t=3x+1), and T(3,t=3x+2)T(3,t=3x+2) in Figure 3. The numerical estimates for mm are listed in the table below.

tt\downarrow ss\rightarrow 2 3 4
4 3.90
5 2.72 5.01 6.69
6
7 4.35 7.10 9.54
8 8.16
9 5.88 12.38
10 10.27
11 7.36 11.33 15.21
12
13 8.22 13.45 18.02

In other words, in this family of examples, we have

ceff=2+24m24st(rst±1)c_{\text{eff}}=2+\frac{24m^{2}}{4st(rst\pm 1)} (64)

Note, in our earlier analysis, we determined the exact value m=1m=1 for Σ(2,3,6±1)\Sigma(2,3,6\pm 1).

Acknowledgements

We would like to thank Miranda C. N. Cheng, Boris Feigin, Kathrin Bringmann, John Cardy, Angus Gruen, Antun Milas, Piotr Kucharski, Sunghyuk Park, Du Pei, Silviu Pufu, and Nicolai Reshetikhin for helpful discussions. This work is supported by a Simons Collaboration Grant on New Structures in Low-Dimensional Topology, by the NSF grant DMS-2245099, and by the U.S. Department of Energy, Office of Science, Office of High Energy Physics, under Award No. DE-SC0011632.

Appendix A qq-Pochhammer symbols near the unit circle

In this appendix, we write down a formula for the logarithm of the q-Pochhammer symbol log[(x,q)]\log[(x,q)_{\infty}] in the domain |x|<1|x|<1 and |q|<1|q|<1, and use it to derive asymptotics of (qa,qb)(q^{a},q^{b}) near roots of unity.

log[(x,q)]\displaystyle\log[(x,q)_{\infty}] =k=0log[1xqk]=k=0=1xqk==1xk=0qk,\displaystyle=\sum_{k=0}^{\infty}\log[1-xq^{k}]=-\sum_{k=0}^{\infty}\sum_{\ell=1}^{\infty}\frac{x^{\ell}}{\ell}q^{\ell k}=-\sum_{\ell=1}^{\infty}\frac{x^{\ell}}{\ell}\sum_{k=0}^{\infty}q^{\ell k}, (65)
==1x1q1==1xn=0Bnn!(log[q])n1,\displaystyle=\sum_{\ell=1}^{\infty}\frac{x^{\ell}}{\ell}\frac{1}{q^{\ell}-1}=\sum_{\ell=1}^{\infty}\frac{x^{\ell}}{\ell}\sum_{n=0}^{\infty}\frac{B_{n}}{n!}\left(\log[q^{\ell}]\right)^{n-1}, (66)
=n=0Bnn!(log[q])n1=1xn2,\displaystyle=\sum_{n=0}^{\infty}\frac{B_{n}}{n!}\left(\log[q]\right)^{n-1}\sum_{\ell=1}^{\infty}x^{\ell}\ell^{n-2}, (67)
log[(x,q)]\displaystyle\log[(x,q)_{\infty}] =n=0Bnn!(log[q])n1Li2n(x).\displaystyle=\sum_{n=0}^{\infty}\frac{B_{n}}{n!}\left(\log[q]\right)^{n-1}\mathrm{Li}_{2-n}(x). (68)

Here BnB_{n} are the Bernoulli numbers, given by the following generating series

tet1=n=0Bntnn!.\frac{t}{e^{t}-1}=\sum_{n=0}^{\infty}B_{n}\frac{t^{n}}{n!}. (69)

If rr\in{\mathbb{Z}} is such that ar0modbar\neq 0\mod b, then we can write the asymptotics of (qa,qb)(q^{a},q^{b}) as qe2πirbq\rightarrow e^{\frac{2\pi ir}{b}},

(qa,qb)exp(Li2(qa)log(qb))asqe2πirb.(q^{a},q^{b})\sim\exp(\frac{\mathrm{Li}_{2}(q^{a})}{\log(q^{b})})\qquad\text{as}\quad q\rightarrow e^{\frac{2\pi ir}{b}}. (70)

References

  • [1] A. B. Zamolodchikov. Irreversibility of the Flux of the Renormalization Group in a 2D Field Theory. JETP Lett., 43:730–732, 1986.
  • [2] John L. Cardy. Is There a c Theorem in Four-Dimensions? Phys. Lett. B, 215:749–752, 1988.
  • [3] Edwin Barnes, Kenneth A. Intriligator, Brian Wecht, and Jason Wright. Evidence for the strongest version of the 4d a-theorem, via a-maximization along RG flows. Nucl. Phys. B, 702:131–162, 2004.
  • [4] Zohar Komargodski and Adam Schwimmer. On Renormalization Group Flows in Four Dimensions. JHEP, 12:099, 2011.
  • [5] Daniel L. Jafferis, Igor R. Klebanov, Silviu S. Pufu, and Benjamin R. Safdi. Towards the F-Theorem: N=2 Field Theories on the Three-Sphere. JHEP, 06:102, 2011.
  • [6] Igor R. Klebanov, Silviu S. Pufu, Subir Sachdev, and Benjamin R. Safdi. Entanglement Entropy of 3-d Conformal Gauge Theories with Many Flavors. JHEP, 05:036, 2012.
  • [7] H. Casini and Marina Huerta. On the RG running of the entanglement entropy of a circle. Phys. Rev. D, 85:125016, 2012.
  • [8] Igor R. Klebanov, Tatsuma Nishioka, Silviu S. Pufu, and Benjamin R. Safdi. Is Renormalized Entanglement Entropy Stationary at RG Fixed Points? JHEP, 10:058, 2012.
  • [9] John L. Cardy. Operator Content of Two-Dimensional Conformally Invariant Theories. Nucl. Phys. B, 270:186–204, 1986.
  • [10] Miranda C N Cheng, Sungbong Chun, Francesca Ferrari, Sergei Gukov, and Sarah M. Harrison. 3d modularity. Journal of High Energy Physics, 2019.
  • [11] Tobias Ekholm, Angus Gruen, Sergei Gukov, Piotr Kucharski, Sunghyuk Park, Marko Stošić, and Piotr Sułkowski. Branches, quivers, and ideals for knot complements. J. Geom. Phys., 177:104520, 2022.
  • [12] Miranda C. N. Cheng, Sungbong Chun, Boris Feigin, Francesca Ferrari, Sergei Gukov, Sarah M. Harrison, and Davide Passaro. 3-Manifolds and VOA Characters, 1 2022.
  • [13] David Kutasov and Finn Larsen. Partition sums and entropy bounds in weakly coupled CFT. JHEP, 01:001, 2001.
  • [14] Sergei Gukov, Du Pei, Pavel Putrov, and Cumrun Vafa. BPS spectra and 3-manifold invariants. J. Knot Theor. Ramifications, 29(02):2040003, 2020.
  • [15] Sergei Gukov, Pavel Putrov, and Cumrun Vafa. Fivebranes and 3-manifold homology. JHEP, 07:071, 2017.
  • [16] Abhijit Gadde, Sergei Gukov, and Pavel Putrov. Walls, Lines, and Spectral Dualities in 3d Gauge Theories. JHEP, 05:047, 2014.
  • [17] Sungbong Chun, Sergei Gukov, Sunghyuk Park, and Nikita Sopenko. 3d-3d correspondence for mapping tori. JHEP, 09:152, 2020.
  • [18] Piotr Kucharski. Z^\widehat{Z} invariants at rational τ\tau. JHEP, 09:092, 2019.
  • [19] Hee-Joong Chung. BPS Invariants for 3-Manifolds at Rational Level KK. JHEP, 02:083, 2021.
  • [20] Tudor Dan Dimofte, Davide Gaiotto, and Sergei Gukov. Gauge theories labelled by three-manifolds. Communications in Mathematical Physics, 325:367–419, 2011.
  • [21] Hee-Joong Chung. Index for a Model of 3d-3d Correspondence for Plumbed 3-Manifolds. Nucl. Phys. B, 965:115361, 2021.
  • [22] Sergei Gukov, Marcos Marino, and Pavel Putrov. Resurgence in complex Chern-Simons theory, 5 2016.
  • [23] Tudor Dimofte, Sergei Gukov, and Lotte Hollands. Vortex Counting and Lagrangian 3-manifolds. Lett. Math. Phys., 98:225–287, 2011.
  • [24] Sander P. Zwegers. Mock theta functions. arXiv: Number Theory, 2008.
  • [25] Basil Gordon and Richard J. Mcintosh. Modular Transformations of Ramanujan’s Fifth and Seventh Order Mock Theta Functions. The Ramanujan Journal, 7(1):193–222, 2003.
  • [26] Sunghyuk Park. Inverted state sums, inverted Habiro series, and indefinite theta functions, 6 2021.
  • [27] Sergei Gukov and Ciprian Manolescu. A two-variable series for knot complements. Quantum Topology, 2019.