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

The Bonahon-Wong-Yang volume conjecture
for the four-puncture sphere

Tushar Pandey
Abstract.

We compute the Bonahon-Wong-Yang quantum invariant for self-diffeomorphisms of the four-puncture sphere explicitly, based on the representation theory of the Checkov-Fock algebra. As an application of the computation, we verify the volume conjecture proposed by Bonahon-Wong-Yang for the four puncture sphere bundles with some technical conditions. Together with the one-puncture torus bundles [BWY3], this is the first family of examples for which the conjecture is verified providing evidence for the conjecture.

1. Introduction

The original Volume Conjecture of Kashaev-Murakami-Murakami [K97, MM01] proposes a relationship between the evaluation of N-th colored Jones polynomial of a link in SS3\SS^{3} at a root of unity q=exp(2πiN)q=\exp\left(\frac{2\pi i}{N}\right) with the hyperbolic volume of the complement of the link. The current paper is devoted to another volume conjecture, developed by Bonahon-Wong-Yang [BWY1] as a toy model for the original Volume Conjecture. This conjecture is based on recent developments of the representation theory of the Kauffman bracket skein algebra 𝒦q(S)\mathcal{K}^{q}(S) of a surface [BW16, FKBL19, GJS19], a mathematical object arising from the theory underlying the Jones polynomial of a link in SS3\SS^{3}.

Let SS be a surface with negative Euler characteristic χ(S)<0\chi(S)<0, and consider an orientation-preserving pseudo-Anosov diffeomorphism φ:SS\varphi\colon S\rightarrow S. Let [r]𝒳SL2()(S)[r]\in\mathcal{X}_{\mathrm{SL}_{2}(\mathbb{C})}(S) be an irreducible φ\varphi-invariant character, namely represented by an irreducible group homomorphism r:π1(S)SL2()r\colon\pi_{1}(S)\rightarrow\mathrm{SL}_{2}(\mathbb{C}) such that [rφ]=[r][r\circ\varphi^{*}]=[r], where φ:π1(S)π1(S)\varphi^{*}\colon\pi_{1}(S)\rightarrow\pi_{1}(S) is induced by the diffeomorphism φ\varphi. The combination of Thurston’s Hyperbolization Theorem and Thurston’s Hyperbolic Dehn Filling Theorem [BP] , applied to the mapping torus Mφ=S×[0,1]/(x,1)(φ(x),0)M_{\varphi}=S\times[0,1]/(x,1)\sim(\varphi(x),0), shows that there exist many such invariant characters [r][r].

For nn odd select, at each puncture vv of SS, a weight pvp_{v}\in\mathbb{C} such that Tn(pv)=Tracer(αv)T_{n}(p_{v})=-\operatorname{Trace}r(\alpha_{v}) where αv\alpha_{v} is a loop going around the puncture vv. Then, for the quantum parameter q=exp(2πin)q=\exp\left(\frac{2\pi i}{n}\right), results from [BW16, FKBL19, GJS19] associate to this data an irreducible representation ρr:𝒦q(S)End(V)\rho_{r}\colon\mathcal{K}^{q}(S)\rightarrow\mathrm{End}(V) of the Kauffman bracket skein algebra 𝒦q(S)\mathcal{K}^{q}(S). If, in addition, we choose the puncture weights pvp_{v} to be φ\varphi-invariant, namely pv=pφ(v)p_{v}=p_{\varphi(v)}, the uniqueness part of these results shows that ρr:𝒦q(S)End(V)\rho_{r}\colon\mathcal{K}^{q}(S)\rightarrow\mathrm{End}(V) and ρrφ:𝒦q(S)End(V)\rho_{r}\circ\varphi_{*}\colon\mathcal{K}^{q}(S)\rightarrow\mathrm{End}(V) are isomorphic, where φ:𝒦q(S)𝒦q(S)\varphi_{*}\colon\mathcal{K}^{q}(S)\rightarrow\mathcal{K}^{q}(S) is defined by φ([K])=[(φ×Id[0,1])(K)]\varphi_{*}([K])=[(\varphi\times Id_{[0,1]})(K)] for every [K]𝒦q(S)[K]\in\mathcal{K}^{q}(S) represented by a framed link KS×[0,1]K\subset S\times[0,1]. For punctured surfaces, this rr is generic, that is, there exists a Zariski dense subset where the uniqueness holds, while for closed surfaces, the result holds for all irreducible rr. Namely, there exists a linear isomorphism Λφ,rq:VV\Lambda_{\varphi,r}^{q}\colon V\rightarrow V such that

(ρrφ)(X)=Λφ,rqρr(X)(Λφ,rq)1\displaystyle(\rho_{r}\circ\varphi_{*})(X)=\Lambda_{\varphi,r}^{q}\circ\rho_{r}(X)\circ(\Lambda_{\varphi,r}^{q})^{-1}

for every X𝒦q(S)X\in\mathcal{K}^{q}(S). By irreducibility of ρr\rho_{r}, this intertwiner Λφ,rq\Lambda_{\varphi,r}^{q} is uniquely determined up to conjugation and scalar multiplication. In particular, if we normalize it so that detΛφ,rq=1\det\Lambda_{\varphi,r}^{q}=1, the modulus |TraceΛφ,rq||\operatorname{Trace}\Lambda_{\varphi,r}^{q}| is uniquely determined.

For each puncture vv with a loop αv\alpha_{v} going around it, write the eigenvalues of r(αv)SL2()r(\alpha_{v})\in\mathrm{SL}_{2}(\mathbb{C}) as e±θv-e^{\pm\theta_{v}} for some θv\theta_{v}\in\mathbb{C}. In particular, Tracer(αv)=eθveθv\operatorname{Trace}r(\alpha_{v})=-e^{\theta_{v}}-e^{-\theta_{v}}. Then, an elementary property of the Chebychev polynomial TnT_{n} is that the solutions of the equation

Tn(pv)=Tracer(αv)\displaystyle T_{n}(p_{v})=-\operatorname{Trace}r(\alpha_{v})

are all numbers of the form pv=t+t1p_{v}=t+t^{-1} with tn=eθvt^{n}=e^{\theta_{v}}. In particular, we can use the solution pv=eθvn+eθvnp_{v}=e^{\frac{\theta_{v}}{n}}+e^{-\frac{\theta_{v}}{n}}.

Conjecture 1.1.

[BWY1] In the above setup,

limn1nlog|TraceΛφ,rq|=14πvol(Mφ)\displaystyle\lim_{n\rightarrow\infty}\frac{1}{n}\log|\operatorname{Trace}\;\Lambda_{\varphi,r}^{q}|=\frac{1}{4\pi}\operatorname{vol}(M_{\varphi})

where vol(Mφ)\operatorname{vol}(M_{\varphi}) is the volume of the complete hyperbolic metric of the mapping torus Mφ=S×[0,1]/(x,1)(φ(x),0)M_{\varphi}=S\times[0,1]/(x,1)\sim(\varphi(x),0).

We verify the conjecture for diffeomorphisms of the four-puncture sphere under a technical assumption that the volume of the mapping torus is large. More precisely, such a mapping torus has a natural ideal triangulation with 2k02k_{0} ideal tetrahedra, which implies that the volume of its complete hyperbolic metric is bounded by 2k0v32k_{0}v_{3}, where v31.0145v_{3}\approx 1.0145\dots is the volume of the regular ideal tetrahedron. We will require the volume to be at least 2(k01)v32(k_{0}-1)v_{3}.

Theorem 1.2.

Let φ:S0,4S0,4\varphi\colon S_{0,4}\rightarrow S_{0,4} be an orientation-preserving, pseudo-Anosov diffeomorphism of the four-puncture sphere. Let [r]𝒳SL2()(S)[r]\in\mathcal{X}_{\mathrm{SL}_{2}(\mathbb{C})}(S) be a generic φ\varphi-invariant character, with puncture weights p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} that are φ\varphi-invariant. Assume in addition that vol(Mφ)>2(k01)v3\operatorname{vol}(M_{\varphi})>2(k_{0}-1)v_{3}, where 2k02k_{0} is the number of tetrahedra appearing in the canonical ideal triangulation of the mapping torus MφM_{\varphi} and where v3v_{3} is the volume of the regular ideal hyperbolic tetrahedron. Then,

|TraceΛφ,rq|=cen4πvol(Mφ)(1+O(1n))\displaystyle\left|\operatorname{Trace}\;\Lambda_{\varphi,r}^{q}\right|=ce^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})}\left(1+O\left(\tfrac{1}{n}\right)\right)

where c0c\neq 0 is a constant that does not grow exponentially in n.

Remark 1.3.

There is at least one infinite family of mapping torus for four-puncture spheres that satisfy the condition in the theorem 1.2, the diffeomorphism word is given by (LR)k(LR)^{k} for any k>1k>1, where the result holds true. This is because for LRLR, there are four tetrahedra with volume is v3v_{3} for each of them, and we can stack set of these four tetrahedra together vertically (giving us (LR)k(LR)^{k} for stacking k times) to get the desired result.

As a Corollary, we prove that the Conjecture 1.1 holds in this case.

Corollary 1.4.

Under the hypotheses of Theorem 1.2,

limn1nlog|TraceΛφ,rq|=14πvol(Mφ).\displaystyle\lim_{n\rightarrow\infty}\frac{1}{n}\log|\operatorname{Trace}\;\Lambda_{\varphi,r}^{q}|=\frac{1}{4\pi}\operatorname{vol}(M_{\varphi}).

Acknowledgements. This research was partially supported by the NSF grants DMS-1812008 and DMS-2203334 (PI: Tian Yang). The author would like to thank his supervisor Tian Yang for guidance and support. The author would also like to thank Francis Bonahon and Ka Ho Wong for useful discussions.

2. Surface diffeomorphisms and the Chekhov-Fock algebra

2.1. Surface diffeomorphisms for Four-puncture sphere

Here, we recall the surface diffeomorphism for four-puncture spheres. The mapping class group of the four-puncture sphere is isomorphic to the semi-direct product PSL2()(2×2)\text{PSL}_{2}({\mathbb{Z}})\ltimes({\mathbb{Z}}_{2}\times{\mathbb{Z}}_{2}) [FM]. Consider the generators L=(1101)L=\begin{pmatrix}1&1\\ 0&1\end{pmatrix} and R=(1011)R=\begin{pmatrix}1&0\\ 1&1\end{pmatrix} for PSL2()\text{PSL}_{2}({\mathbb{Z}}) and (ψ1,ψ2)2×2(\psi_{1},\psi_{2})\in{\mathbb{Z}}^{2}\times{\mathbb{Z}}^{2}.
Note that the elements LL and RR fix one puncture while ψ1,ψ2\psi_{1},\psi_{2} permutes the punctures. Therefore, we can write any orientation preserving diffeomorphism φ:S0,4S0,4\varphi:S_{0,4}\to S_{0,4} as

(2.1) φ=Ln1Rm1LnkRmkψ1ϵ1ψ2ϵ2\displaystyle\varphi=L^{n_{1}}R^{m_{1}}\dots L^{n_{k}}R^{m_{k}}\psi_{1}^{\epsilon_{1}}\psi_{2}^{\epsilon_{2}}

where ni,mi>0n_{i},m_{i}\in{\mathbb{Z}}_{>0} and ϵ1,ϵ2{0,1}\epsilon_{1},\epsilon_{2}\in\{0,1\}. For the diffeomorphism to be pseudo-Anosov, we require k1k\geq 1 [GF]. As explained in the subsequent sections, these generators will play a fundamental role in the computation of the invariant.

2.2. Ideal triangulation and edge weight system

We represent the four-puncture sphere as S0,4=S^0,4{p1,p2,p3,p4}S_{0,4}=\hat{S}_{0,4}-\{p_{1},p_{2},p_{3},p_{4}\} where p1,p2,p3,p4p_{1},p_{2},p_{3},p_{4} are the four punctures, and S^0,4\hat{S}_{0,4} is a compact surface such that the ideal triangulation of S0,4S_{0,4} is the ideal triangulation of S^0,4\hat{S}_{0,4} with the punctures as the vertex set. The triangulation has 6 edges and 4 faces.
Given an ideal triangulation τ\tau, we assign complex numbers aia_{i}\in\mathbb{C}^{*} to each edge γi\gamma_{i} of τ\tau. This determines a character [r~]𝒳PSL2()[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})} (see [BL07] for details). We get an edge weight system 𝕒=(a1,,a6)()6\mathbb{a}=(a_{1},\dots,a_{6})\in(\mathbb{C}^{*})^{6}. These aia_{i}’s are also called the shear bend parameters.

Lemma 2.1.

[BWY1] Given an ideal triangulation τ\tau , there exists a Zariski-open dense subset U𝒳PSL2()(S0,4)U\subset\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}) such that every character [r~]U[\tilde{r}]\in U is associated as above to an edge weight system 𝕒()6\mathbb{a}\in(\mathbb{C}^{*})^{6} for τ\tau.

2.3. Ideal triangulation sweeps and φ\varphi-invariant characters

Recall that the edge labels are fixed in the triangulation. The two elementary operations on a triangulation are either an edge re-indexing or a diagonal exchange (see [BWY1]). The edge re-indexing is just a permutation of the edges γi\gamma_{i} such that γi=γσ(i)\gamma^{\prime}_{i}=\gamma_{\sigma(i)} for some permutation σ\sigma of the set {1,,6}\{1,\dots,6\}. The diagonal exchange along the edge γi0\gamma_{i_{0}} is such that the ideal triangulation τ\tau with edges γi\gamma_{i} are replaced with an ideal triangulation τ\tau^{\prime} with edges γi\gamma^{\prime}_{i} such that γi=γi\gamma^{\prime}_{i}=\gamma_{i} when ii0i\neq i_{0} and γi0\gamma^{\prime}_{i_{0}} is the other diagonal of the square formed by the two faces of τ\tau that are adjacent to γi0\gamma_{i_{0}}. For reference, see figure 3 (the labels are slightly different and would make sense later). None of the edges in the figure are identified.

We now look at how the edge weight system changes. If the triangulation τ\tau^{\prime} is obtained by re-indexing by a permutation σ\sigma, then the edge weights ai=aσ(i)a^{\prime}_{i}=a_{\sigma(i)} define the same character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}). If τ\tau^{\prime} is obtained by a diagonal exchange at its i0i_{0}-th edge, there exists edge weights aia^{\prime}_{i} for τ\tau^{\prime} define the same character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}) as long as ai01a_{i_{0}}\neq-1. The formula is defined in [BL07] as follows

ai={aiif ii0,i1,i2,i3,i4ai01if i=i0ai(1+ai0)if i=i1 or i3ai(1+ai01)1if i=i2 or i4.\displaystyle a^{\prime}_{i}=\begin{cases}a_{i}&\text{if }i\neq i_{0},i_{1},i_{2},i_{3},i_{4}\\ a_{i_{0}}^{-1}&\text{if }i=i_{0}\\ a_{i}(1+a_{i_{0}})&\text{if }i=i_{1}\text{ or }i_{3}\\ a_{i}(1+a_{i_{0}}^{-1})^{-1}&\text{if }i=i_{2}\text{ or }i_{4}.\end{cases}

Suppose that we can connect the two ideal triangulations τ\tau and φ(τ)\varphi(\tau) by a finite sequence of ideal triangulations τ=τ(0),τ(1),,τ(k0)=φ(τ)\tau=\tau^{(0)},\tau^{(1)},\dots,\tau^{(k_{0})}=\varphi(\tau), where each τ(k)\tau^{(k)} is obtained from τ(k1)\tau^{(k-1)} by an edge re-indexing or a diagonal exchange (such a sequence always exists). This is called an ideal triangulation sweep from τ\tau to τ\tau^{\prime}.

We start with an edge weight system a(0)=(a1(0),,a6(0))()6a^{(0)}=(a^{(0)}_{1},\dots,a^{(0)}_{6})\in(\mathbb{C}^{*})^{6} for τ(0)=τ\tau^{(0)}=\tau. By using the formulas above, we obtain an edge weight system a(k)()6a^{(k)}\in(\mathbb{C}^{*})^{6} for each τ(k)\tau^{(k)} that all define the same character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}). This edge weight system is well defined as long as ai(k1)1a_{i}^{(k-1)}\neq-1 whenever we have a diagonal exchange at the i-th edge. We then have an edge weight system a(0),a(1),,a(k0)a^{(0)},a^{(1)},\dots,a^{(k_{0})} for an ideal triangulation sweep τ=τ(0),,τ(k0)=φ(τ)\tau=\tau^{(0)},\dots,\tau^{(k_{0})}=\varphi(\tau).

By construction, an edge weight system for an ideal triangulation sweep uniquely determines a character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}).

Lemma 2.2.

[BWY1] Let τ=τ(0),τ(1),,τ(k0)=φ(τ)\tau=\tau^{(0)},\tau^{(1)},\dots,\tau^{(k_{0})}=\varphi(\tau) be an ideal triangulation sweep from τ\tau to φ(τ)\varphi(\tau). If the weight system a(0),a(1),,a(k0)a^{(0)},a^{(1)},\dots,a^{(k_{0})} for this sweep is such that a(k0)=a(0)a^{(k_{0})}=a^{(0)}, then the associated character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}) is fixed by the action of φ\varphi on 𝒳PSL2()(S0,4)\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}).

Conversely, there is a Zariski open dense subset U𝒳PSL2()(S0,4)U\subset\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}) such that every φ\varphi-invariant character [r~]U[\tilde{r}]\in U is associated in this way to an edge weight system a(0),,a(k0)a^{(0)},\dots,a^{(k_{0})} with a(k0)=a(0)a^{(k_{0})}=a^{(0)}. ∎

2.4. Chekhov-Fock Algebra for the four-puncture sphere

We will work with the Chekhov-Fock (CF) algebra of the four punctured spheres to compute the intertwiner Λφ,rq\Lambda_{\varphi,r}^{q}. By [BWY1, Theorem 16], we see that this intertwiner is conjugate to the intertwiner coming from the Kauffman bracket skein algebra. This implies the trace of the intertwiner remains unchanged, and thus we can use the CF algebra intertwiner to get the invariant.

Let the edges of the four-punctured sphere be labeled as follows:

Refer to caption
Figure 1. Triangulation for four puncture sphere

The Chekhov-Fock algebra is the algebra of the Laurent polynomials in X1,,X6X_{1},\dots,X_{6} associated with the edges of triangulation as in the figure, 𝒯τq(S0,4)=[X1±1,,X6±1]\mathcal{T}^{q}_{\tau}(S_{0,4})=\mathbb{C}[X_{1}^{\pm 1},\dots,X_{6}^{\pm 1}]. The relations for the algebra are given by ”qq-commutativity”, that is

XiXj=q2σijXjXi,\displaystyle X_{i}X_{j}=q^{2\sigma_{ij}}X_{j}X_{i},

where σij=[Uncaptioned image]\sigma_{ij}=\vbox{\hbox{\includegraphics[width=85.35826pt]{sigma_ij}}}.
Thus, the qq-commuting relations are:

X1X2=q2X2X1X2X3\displaystyle X_{1}X_{2}=q^{2}X_{2}X_{1}\;\;\;\;\;\;X_{2}X_{3} =q2X3X2X3X5=q2X5X3\displaystyle=q^{2}X_{3}X_{2}\;\;\;\;\;\;\;\;X_{3}X_{5}=q^{-2}X_{5}X_{3}
X1X3=q2X3X1X2X4\displaystyle X_{1}X_{3}=q^{-2}X_{3}X_{1}\;\;\;\;\;\;X_{2}X_{4} =q2X4X2X4X5=q2X5X4\displaystyle=q^{-2}X_{4}X_{2}\;\;\;\;\;\;X_{4}X_{5}=q^{2}X_{5}X_{4}
X1X5=q2X5X1X2X6\displaystyle X_{1}X_{5}=q^{2}X_{5}X_{1}\;\;\;\;\;\;X_{2}X_{6} =q2X6X2X4X6=q2X6X4\displaystyle=q^{2}X_{6}X_{2}\;\;\;\;\;\;\;\;X_{4}X_{6}=q^{-2}X_{6}X_{4}
X1X6=q2X6X1X3X4\displaystyle X_{1}X_{6}=q^{-2}X_{6}X_{1}\;\;\;\;\;\;X_{3}X_{4} =q2X4X3X5X6=q2X6X5.\displaystyle=q^{2}X_{4}X_{3}\;\;\;\;\;\;\;\;X_{5}X_{6}=q^{2}X_{6}X_{5}.

See [BL07] for more details on the algebra and its representations.

The classification of irreducible representations of 𝒯τq(S0,4)\mathcal{T}^{q}_{\tau}(S_{0,4}) involves central elements Pj𝒯τq(S0,4)P_{j}\in\mathcal{T}^{q}_{\tau}(S_{0,4}) for j=1,2,3,4j=1,2,3,4 associated to each puncture on the surface.

Theorem 2.3.

[BL07, Theorem 21] If q is a primitive n-th root of unity with n odd, and if ρ:𝒯τq(S0,4)End(V)\rho\colon\mathcal{T}^{q}_{\tau}(S_{0,4})\rightarrow\mathrm{End}(V) is a finite-dimensional irreducible representation, then dim(V)=n\dim(V)=n and there exists a1,a2,,a6,p1,p2,p3,p4a_{1},a_{2},\dots,a_{6},p_{1},p_{2},p_{3},p_{4}\in\mathbb{C}^{*} such that

ρ(Xin)=aiIdV\displaystyle\rho(X_{i}^{n})=a_{i}\operatorname{Id}_{V} ρ(Pj)=pjIdV\displaystyle\rho(P_{j})=p_{j}\operatorname{Id}_{V}

for every edge i=1,,6i=1,\dots,6 of the ideal triangulation τ\tau and punctures j=1,,4j=1,\dots,4. Furthermore, ρ\rho is determined up to isomorphism by this set of invariants a1,,a6,p1,,p4a_{1},\dots,a_{6},p_{1},\dots,p_{4}\in\mathbb{C}^{*}. Such a collection is an irreducible representation if and only if

p1n=a1a2a3\displaystyle p_{1}^{n}=a_{1}a_{2}a_{3} p2n=a2a4a6\displaystyle p_{2}^{n}=a_{2}a_{4}a_{6} p3n=a1a5a6\displaystyle p_{3}^{n}=a_{1}a_{5}a_{6} p4n=a3a4a5.\displaystyle p_{4}^{n}=a_{3}a_{4}a_{5}.

We start with a triangulation τ(0)\tau^{(0)}. To each edge eie_{i}, we will assign xi(0)x_{i}^{(0)}\in\mathbb{C}^{*} in τ(0)\tau^{(0)}. The central elements of the Chekhov-Fock algebra corresponding to the four punctures are

P1=q1X1X2X3,P2=\displaystyle P_{1}=q^{-1}X_{1}X_{2}X_{3},\;\;P_{2}= qX2X4X6,P3=q1X1X5X6,P4=q1X3X4X5.\displaystyle qX_{2}X_{4}X_{6},\;\;P_{3}=q^{-1}X_{1}X_{5}X_{6},\;\;P_{4}=q^{-1}X_{3}X_{4}X_{5}.

We also define H=q2X1X2X3X4X5X6H=q^{-2}X_{1}X_{2}X_{3}X_{4}X_{5}X_{6}, such that H2=P1P2P3P4H^{2}=P_{1}P_{2}P_{3}P_{4}. We have four central elements and six generators, therefore we can choose two independent generators, say X1,X2X_{1},X_{2}. We rewrite the other generators in terms of these two and the central elements.

(2.2) X3=qP1X21X11X4=P4P2H1X1X5=HP11P21X2X6=qHP41X21X11.\displaystyle X_{3}=qP_{1}X_{2}^{-1}X_{1}^{-1}\;\;\;\;X_{4}=P_{4}P_{2}H^{-1}X_{1}\;\;\;\;X_{5}=HP_{1}^{-1}P_{2}^{-1}X_{2}\;\;\;\;X_{6}=qHP_{4}^{-1}X_{2}^{-1}X_{1}^{-1}.
Remark 2.4.

Here we are actually using the fractional algebra 𝒯^τq(S0,4)\hat{\mathcal{T}}_{\tau}^{q}(S_{0,4}). This algebra consists of formal rational fractions in XiX_{i} for i=1,,6i=1,\dots,6, modulo the same relations as above.

2.5. Standard representation of the algebra 𝒯^τq(S0,4)\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})

The results follow from [BL07]. The results are originally obtained for the Chekov-Fock algebra but can be extended to the fractional algebra by using algebra homomorphisms.

Lemma 2.5.

A standard representation of the algebra 𝒯^τq(S0,4)\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) associated with non-zero numbers x,y,p1,p2,p3,p4x,y,p_{1},p_{2},p_{3},p_{4}\in\mathbb{C}^{*} with 𝕡=(p1,p2,p3,p4)\mathbb{p}=(p_{1},p_{2},p_{3},p_{4}) is given by a representation ρx,y,𝕡:𝒯^τq(S0,4)End(n)\rho_{x,y,\mathbb{p}}\colon\hat{\mathcal{T}}_{\tau}^{q}(S_{0,4})\to\mathrm{End}(\mathbb{C}^{n}) where, if {w1,,wn}\{w_{1},\dots,w_{n}\} is the standard basis for n\mathbb{C}^{n},

ρx,y,𝕡(X1)(wi)\displaystyle\rho_{x,y,\mathbb{p}}(X_{1})(w_{i}) =xq2iwi\displaystyle=xq^{2i}w_{i}
ρx,y,𝕡(X2)(wi)\displaystyle\rho_{x,y,\mathbb{p}}(X_{2})(w_{i}) =yqiwi+1\displaystyle=yq^{-i}w_{i+1}
ρx,y,𝕡(X3)(wi)\displaystyle\rho_{x,y,\mathbb{p}}(X_{3})(w_{i}) =p1(xy)1qiwi1\displaystyle=p_{1}(xy)^{-1}q^{-i}w_{i-1}
ρx,y,𝕡(X4)(wi)\displaystyle\rho_{x,y,\mathbb{p}}(X_{4})(w_{i}) =xq2ip4p2h1wi\displaystyle=xq^{2i}p_{4}p_{2}h^{-1}w_{i}
ρx,y,𝕡(X5)(wi)\displaystyle\rho_{x,y,\mathbb{p}}(X_{5})(w_{i}) =yqihp11p21wi+1\displaystyle=yq^{-i}hp_{1}^{-1}p_{2}^{-1}w_{i+1}
ρx,y,𝕡(X6)(wi)\displaystyle\rho_{x,y,\mathbb{p}}(X_{6})(w_{i}) =qi(xy)1hp41wi1\displaystyle=q^{-i}(xy)^{-1}hp_{4}^{-1}w_{i-1}

for every i=1,,ni=1,\dots,n and counting indices modulo n.

Proof.

The proof follows from the fact that there are two independent generators X1X_{1} and X2X_{2} and like the one-puncture torus case, by linear algebra (see [BWY1]), one can see that the representations for X1X_{1} and X2X_{2} are given by

ρx,y,𝕡(X1)(wi)=xq2iwi\displaystyle\rho_{x,y,\mathbb{p}}(X_{1})(w_{i})=xq^{2i}w_{i} ρx,y,𝕡(X2)(wi)=yqiwi+1.\displaystyle\rho_{x,y,\mathbb{p}}(X_{2})(w_{i})=yq^{-i}w_{i+1}.

This implies

ρx,y,𝕡(X11)(wi)=x1q2iwi\displaystyle\rho_{x,y,\mathbb{p}}(X_{1}^{-1})(w_{i})=x^{-1}q^{-2i}w_{i} ρx,y,𝕡(X21)(wi)=y1qi1wi1.\displaystyle\rho_{x,y,\mathbb{p}}(X_{2}^{-1})(w_{i})=y^{-1}q^{i-1}w_{i-1}.

The central elements are PjP_{j} for j=1,2,3,4j=1,2,3,4 and HH, with

ρx,y,𝕡(Pj)=pjIdn\displaystyle\rho_{x,y,\mathbb{p}}(P_{j})=p_{j}\operatorname{Id}_{\mathbb{C}^{n}} ρx,y,𝕡(H)=hIdn.\displaystyle\rho_{x,y,\mathbb{p}}(H)=h\operatorname{Id}_{\mathbb{C}^{n}}.

The image for the other XjX_{j} for j=3,4,5,6j=3,4,5,6 depends on the central elements and therefore can be computed using the homomorphism property of the representation. For example,

ρx,y,𝕡(X3)(wi)=\displaystyle\rho_{x,y,\mathbb{p}}(X_{3})(w_{i})= ρs(qP1X21X11)(wi)=qp1ρx,y,𝕡(X21)(ρs(X11)(wi))\displaystyle\rho_{s}(qP_{1}X_{2}^{-1}X_{1}^{-1})(w_{i})=qp_{1}\rho_{x,y,\mathbb{p}}(X_{2}^{-1})(\rho_{s}(X_{1}^{-1})(w_{i}))
=\displaystyle= qp1x1q2iρx,y,𝕡(X21)(wi)=y1p1x1q12iqi1(wi1)=(xy)1p1qiwi1.\displaystyle qp_{1}x^{-1}q^{-2i}\rho_{x,y,\mathbb{p}}(X_{2}^{-1})(w_{i})=y^{-1}p_{1}x^{-1}q^{1-2i}q^{i-1}(w_{i-1})=(xy)^{-1}p_{1}q^{-i}w_{i-1}.

The other representations are calculated similarly.

Note that ρx,y,𝕡(X1n)=xnIdn,ρx,y,𝕡(X2n)=ynIdn\rho_{x,y,\mathbb{p}}(X_{1}^{n})=x^{n}\operatorname{Id}_{\mathbb{C}^{n}},\rho_{x,y,\mathbb{p}}(X_{2}^{n})=y^{n}\operatorname{Id}_{\mathbb{C}^{n}} and ρx,y,𝕡(Pj)=pjIdn\rho_{x,y,\mathbb{p}}(P_{j})=p_{j}\operatorname{Id}_{\mathbb{C}^{n}}. We let a=xna=x^{n} and b=ynb=y^{n}. ∎

Proposition 2.6.

[BL07]

  1. (1)

    Every standard representation ρx,y,𝕡:𝒯^τq(S0,4)End(V)\rho_{x,y,\mathbb{p}}\colon\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})\rightarrow\mathrm{End}(V) is irreducible.

  2. (2)

    Every irreducible representation of 𝒯^τq(S0,4)\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) is isomorphic to a standard representation ρx,y,𝕡\rho_{x,y,\mathbb{p}}.

  3. (3)

    Two standard representations ρx,y,𝕡\rho_{x,y,\mathbb{p}} and ρx,y,𝕡\rho_{x^{\prime},y^{\prime},\mathbb{p}^{\prime}} are isomorphic if and only if xn=xn,yn=ynx^{n}={x^{\prime}}^{n},y^{n}=y^{\prime n} and 𝕡=𝕡\mathbb{p}=\mathbb{p}^{\prime}.

The above proposition implies that any irreducible representation of the Chekhov-Fock algebra is classified up to isomorphism by the numbers a,b,p1,p2,p3,p4,ha,b,p_{1},p_{2},p_{3},p_{4},h\in\mathbb{C}^{*}.

2.6. Coordinate Change isomorphisms

We study the coordinate change isomorphism of the Chekhov-Fock algebra for any two ideal triangulations τ\tau and τ\tau^{\prime} as Φττq:𝒯^τq(S0,4)𝒯^τq(S0,4)\Phi_{\tau\tau^{\prime}}^{q}:\hat{\mathcal{T}}^{q}_{\tau^{\prime}}(S_{0,4})\to\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}). This isomorphism satisfies the fundamental relation

Φττ′′q=ΦττqΦττ′′q\displaystyle\Phi_{\tau\tau^{\prime\prime}}^{q}=\Phi_{\tau\tau^{\prime}}^{q}\circ\Phi_{\tau^{\prime}\tau^{\prime\prime}}^{q}

for any three ideal triangulations τ,τ,τ′′\tau,\tau^{\prime},\tau^{\prime\prime}. These isomorphisms are given by explicit formulas [BL07]. We consider the triangulation sweep from τ\tau to φ(τ)\varphi(\tau) and write down the explicit isomorphisms.

If two ideal triangulations differ by an edge re-indexing, that is, if τ\tau^{\prime} is obtained from τ\tau by an edge re-indexing with the i-th edge of τ\tau^{\prime} being the σ(i)\sigma(i)-th edge of τ\tau for some permutation σ\sigma of {1,2,,6}\{1,2,\dots,6\}, then Φττq:𝒯^τq(S0,4)𝒯^τq(S0,4)\Phi_{\tau\tau^{\prime}}^{q}:\hat{\mathcal{T}}^{q}_{\tau^{\prime}}(S_{0,4})\to\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) is the unique isomorphism given by Φτ,τq(Xi)=Xσ(i)\Phi_{\tau,\tau^{\prime}}^{q}(X_{i}^{\prime})=X_{\sigma(i)}.

Refer to caption
Figure 2. A diagonal exchange

If τ\tau^{\prime} is obtained from τ\tau using a diagonal exchange as in figure 2, with the edge re-indexing, then Φττq:𝒯^τq(S0,4)𝒯^τq(S0,4)\Phi_{\tau\tau^{\prime}}^{q}:\hat{\mathcal{T}}^{q}_{\tau^{\prime}}(S_{0,4})\to\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) is the unique algebra homomorphism given by

(2.3) Φττq(Xi)={Xi if ii0,i1,i2,i3,i4Xi01 if i=i0Xi(1+q1Xi0) if i=i1 or i3Xi(1+q1Xi01)1 if i=i2 or i4.\displaystyle\Phi_{\tau\tau^{\prime}}^{q}(X_{i}^{\prime})=\begin{cases}X_{i}&\text{ if }i\neq i_{0},i_{1},i_{2},i_{3},i_{4}\\ X_{i_{0}}^{-1}&\text{ if }i=i_{0}\\ X_{i}(1+q^{-1}X_{i_{0}})&\text{ if }i=i_{1}\text{ or }i_{3}\\ X_{i}(1+q^{-1}X_{i_{0}}^{-1})^{-1}&\text{ if }i=i_{2}\text{ or }i_{4}.\end{cases}

We now consider an algebra homomorphism ρΦττq:𝒯^τq(S0,4)End(V)\rho\circ\Phi^{q}_{\tau\tau^{\prime}}\colon\hat{\mathcal{T}}^{q}_{\tau^{\prime}}(S_{0,4})\to\mathrm{End}(V). For more details on the algebra homomorphism, see [BWY1, Section 3.4].

3. The Quantum Invariant

3.1. Isomorphisms of the Chekhov-Fock algebra

We define the invariant using an ideal triangulation sweep from an ideal triangulation τ\tau to an ideal triangulation φ(τ)\varphi(\tau) of the four-puncture sphere bundle. We have seen that any pseudo-Anosov and orientation-preserving diffeomorphism (2.1) can be written as

φ=Ln1Rm1LnkRmkψ1ϵ1ψ2ϵ2\displaystyle\varphi=L^{n_{1}}R^{m_{1}}\dots L^{n_{k}}R^{m_{k}}\psi_{1}^{\epsilon_{1}}\psi_{2}^{\epsilon_{2}}

where ni,mi>0n_{i},m_{i}\in{\mathbb{Z}}_{>0} and ϵ1,ϵ2{0,1}\epsilon_{1},\epsilon_{2}\in\{0,1\}. We rewrite the diffeomorphism φ\varphi as

(3.1) φ=φ1φk0ψ1ϵ1ψ2ϵ2\displaystyle\varphi=\varphi_{1}\circ\dots\circ\varphi_{k_{0}}\circ\psi_{1}^{\epsilon_{1}}\psi_{2}^{\epsilon_{2}}

where each φk\varphi_{k} corresponds to LL or RR for k={1,,k0}k=\{1,\dots,k_{0}\}. We start with a triangulation τ0\tau_{0} of the surface with independent edges labelled by (X1,X2)(X_{1},X_{2}) and shear-bend parameters (a0,b0)()2(a_{0},b_{0})\in(\mathbb{C}^{*})^{2}.

For each k=1,,k0k=1,\dots,k_{0}, we define the triangulation τk=φ1φk(τ0)\tau_{k}=\varphi_{1}\circ\dots\circ\varphi_{k}(\tau_{0}).

Refer to caption
Figure 3. We do diagonal exchange along X2,X5X_{2},X_{5} for L and X3,X6X_{3},X_{6} for R

We start with the triangulation (figure 3). For each generator L or R, of the four-punctured sphere, we use two tetrahedra for diagonal exchange. The edges in the diagonal exchanges are X3,X6X_{3},X_{6} for R, and X2,X5X_{2},X_{5} for L.

Refer to caption
Figure 4. (Left) Diagonal switch corresponding to the isomorphism L. (Right) Diagonal switch corresponding to the isomorphism R

We now define isomorphisms 𝒯^τq𝒯^τq\hat{\mathcal{T}}^{q}_{\tau}\to\hat{\mathcal{T}}^{q}_{\tau} to study the algebra homomorphisms of the representations under this isomorphism.

Proposition 3.1.

The left isomorphism :𝒯^τq(S0,4)𝒯^τq(S0,4)\mathcal{L}\colon\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})\to\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) is defined by

(X1)=X21\displaystyle\mathcal{L}(X_{1})=X_{2}^{-1} (X4)\displaystyle\mathcal{L}(X_{4}) =X51\displaystyle=X_{5}^{-1}
(X2)=(1+qX2)(1+qX5)X4\displaystyle\mathcal{L}(X_{2})=(1+qX_{2})(1+qX_{5})X_{4} (X5)\displaystyle\mathcal{L}(X_{5}) =(1+qX2)(1+qX5)X1\displaystyle=(1+qX_{2})(1+qX_{5})X_{1}
(X3)=(1+qX21)1(1+qX51)1X3\displaystyle\mathcal{L}(X_{3})=(1+qX_{2}^{-1})^{-1}(1+qX_{5}^{-1})^{-1}X_{3} (X6)\displaystyle\mathcal{L}(X_{6}) =(1+qX21)1(1+qX51)1X6.\displaystyle=(1+qX_{2}^{-1})^{-1}(1+qX_{5}^{-1})^{-1}X_{6}.

The right isomorphism :𝒯^τq(S0,4)𝒯^τq(S0,4)\mathcal{R}\colon\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})\to\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})

(X1)=X31\displaystyle\mathcal{R}(X_{1})=X_{3}^{-1} (X4)\displaystyle\mathcal{R}(X_{4}) =X61\displaystyle=X_{6}^{-1}
(X2)=(1+qX3)(1+qX6)X2\displaystyle\mathcal{R}(X_{2})=(1+qX_{3})(1+qX_{6})X_{2} (X5)\displaystyle\mathcal{R}(X_{5}) =(1+qX3)(1+qX6)X5\displaystyle=(1+qX_{3})(1+qX_{6})X_{5}
(X3)=(1+qX31)1(1+qX61)1X4\displaystyle\mathcal{R}(X_{3})=(1+qX_{3}^{-1})^{-1}(1+qX_{6}^{-1})^{-1}X_{4} (X6)\displaystyle\mathcal{R}(X_{6}) =(1+qX31)(1+qX61)1X1.\displaystyle=(1+qX_{3}^{-1})(1+qX_{6}^{-1})^{-1}X_{1}.
Proof.

Let’s look at left isomorphism first. The right isomorphism follows the same idea.

In figure 4, X2=X21,X5=X51X^{\prime}_{2}=X_{2}^{-1},X^{\prime}_{5}=X_{5}^{-1} because the diagonal exchange is done along those two edges. From eq 2.3, X4X^{\prime}_{4} first becomes (1+qX5)X4(1+qX_{5})X_{4} through the diagonal exchange along X5X_{5} and after the diagonal exchange along X2X_{2}, it becomes (1+qX2)(1+qX5)X4(1+qX_{2})(1+qX_{5})X_{4}. Similarly, the other edge changes are captured.

Now, we look at the re-indexing. We want to compare the changed (with diagonal switch) triangulation to have the same ”qq-commuting” relations, therefore we re-index the edges as in figure 3. Therefore, (X5)\mathcal{L}(X_{5}) becomes X1X^{\prime}_{1}, (X1)\mathcal{L}(X_{1}) becomes X2X^{\prime}_{2}, (X2)\mathcal{L}(X_{2}) becomes X4X^{\prime}_{4} and so on.

Note that the order of the diagonal switch is important, or else we will get an additional factor of q. ∎

We want to determine the representations ρx1,y1,𝕡(k)\rho_{x_{1},y_{1},\mathbb{p}^{(k)}}\circ\mathcal{L} and ρx1,y1,𝕡(k)\rho_{x_{1},y_{1},\mathbb{p}^{(k)}}\circ\mathcal{R} of 𝒯^τ(k)q\hat{\mathcal{T}}^{q}_{\tau^{(k)}} for a standard representation ρx1,y1,𝕡(k):𝒯^τ(k)q(S0,4)End(V)\rho_{x_{1},y_{1},\mathbb{p}^{(k)}}\colon\hat{\mathcal{T}}_{\tau^{(k)}}^{q}(S_{0,4})\to\mathrm{End}(V).

Proposition 3.2.

Let xk1,yk1x_{k-1},y_{k-1}\in\mathbb{C}^{*} and 𝕡(k1)()4\mathbb{p}^{(k-1)}\in(\mathbb{C}^{*})^{4} be given.
If yk1n1y_{k-1}^{n}\neq-1 and hnyk1n(p1(k1))n(p2(k1))nh^{n}y_{k-1}^{n}\neq-(p^{(k-1)}_{1})^{n}(p^{(k-1)}_{2})^{n}, choose uk,u^k,vk,v^ku_{k},\hat{u}_{k},v_{k},\hat{v}_{k}\in\mathbb{C}^{*} such that

uk=qyk1\displaystyle u_{k}=qy_{k-1} u^k=qh(p1k1)1(p2(k1))1yk1\displaystyle\hat{u}_{k}=qh(p^{k-1}_{1})^{-1}(p^{(k-1)}_{2})^{-1}y_{k-1} vkn=1+ukn\displaystyle v_{k}^{n}=1+u_{k}^{n} v^kn=1+u^kn.\displaystyle\hat{v}_{k}^{n}=1+\hat{u}_{k}^{n}.

Then the representation ρxk1,yk1,𝕡(k1)\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\circ\mathcal{L} is isomorphic to a standard representation ρxk,yk,𝕡(k)\rho_{x_{k},y_{k},\mathbb{p}^{(k)}} with

xk=yk11yk=vkv^kp2(k1)p4(k1)h1xk1\displaystyle x_{k}=y_{k-1}^{-1}\;\;\;\;\;\;\;\;\;\;\;\;y_{k}=v_{k}\hat{v}_{k}p^{(k-1)}_{2}p^{(k-1)}_{4}h^{-1}x_{k-1}
p1(k)=p4(k1)p2(k)=p2(k1)p3(k)=p3(k1)p4(k)=p1(k1).\displaystyle p^{(k)}_{1}=p^{(k-1)}_{4}\;\;\;\;p^{(k)}_{2}=p^{(k-1)}_{2}\;\;\;\;p^{(k)}_{3}=p^{(k-1)}_{3}\;\;\;\;p^{(k)}_{4}=p^{(k-1)}_{1}.

Similarly, if (p1(k1))nxk1nyk1n(p^{(k-1)}_{1})^{n}\neq-x_{k-1}^{n}y_{k-1}^{n} and (p4(k1))nxk1nyk1nhn(p^{(k-1)}_{4})^{n}x_{k-1}^{n}y_{k-1}^{n}\neq-h^{n}, choose uk,u^k,vk,v^ku_{k},\hat{u}_{k},v_{k},\hat{v}_{k}\in\mathbb{C}^{*} such that

uk=qp1(k1)xk11yk11\displaystyle u_{k}=qp^{(k-1)}_{1}x_{k-1}^{-1}y_{k-1}^{-1} u^k=qh(p4(k1))1xk11yk11\displaystyle\hat{u}_{k}=qh(p^{(k-1)}_{4})^{-1}x_{k-1}^{-1}y_{k-1}^{-1} vkn=1+ukn\displaystyle v_{k}^{n}=1+u_{k}^{n} v^kn=1+u^kn.\displaystyle\hat{v}_{k}^{n}=1+\hat{u}_{k}^{n}.

Then the representation ρxk1,yk1,𝕡(k1)\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\circ\mathcal{R} is isomorphic to a standard representation ρxk,yk,𝕡(k)\rho_{x_{k},y_{k},\mathbb{p}^{(k)}} with

xk=(p1(k1))1xk1yk1yk=vkv^kyk1\displaystyle x_{k}=(p^{(k-1)}_{1})^{-1}x_{k-1}y_{k-1}\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;y_{k}=v_{k}\hat{v}_{k}y_{k-1}
p1(k)=p2(k1)p2(k)=p1(k1)p3(k)=p3(k1)p4(k)=p4(k1).\displaystyle p^{(k)}_{1}=p^{(k-1)}_{2}\;\;\;\;p^{(k)}_{2}=p^{(k-1)}_{1}\;\;\;\;p^{(k)}_{3}=p^{(k-1)}_{3}\;\;\;\;p^{(k)}_{4}=p^{(k-1)}_{4}.
Proof.

The computation can be found in [BL07, Lemma 27] corresponding to the isomorphisms defined in the previous Proposition 3.2. ∎

Remark 3.3.

The puncture weight p3p_{3} does not change in either \mathcal{L} or \mathcal{R}.

The shear bend parameters (a,b)()2(a,b)\in(\mathbb{C}^{*})^{2} are just n-th powers of (x,y)()2(x,y)\in(\mathbb{C}^{*})^{2}. Therefore, under the two isomorphisms, for the choices uk,vk,u^k,v^ku_{k},v_{k},\hat{u}_{k},\hat{v}_{k}\in\mathbb{C}^{*}, the shear bend parameters change in the following manner

:\displaystyle\mathcal{L}\colon ak=bk11\displaystyle a_{k}=b_{k-1}^{-1} bk=vknv^kn(p2(k1))n(p4(k1))nhnak1;\displaystyle b_{k}=v_{k}^{n}\hat{v}_{k}^{n}(p^{(k-1)}_{2})^{n}(p^{(k-1)}_{4})^{n}h^{-n}a_{k-1};
(3.2) :\displaystyle\mathcal{R}\colon ak=(p1(k1))nak1bk1\displaystyle a_{k}=(p^{(k-1)}_{1})^{-n}a_{k-1}b_{k-1} bk=vknv^knbk1.\displaystyle b_{k}=v_{k}^{n}\hat{v}_{k}^{n}b_{k-1}.

3.2. Intertwiners

Let τ=τ(0),τ(1),,τ(k0)=φ(τ)\tau=\tau^{(0)},\tau^{(1)},\dots,\tau^{(k_{0})}=\varphi(\tau) be an ideal triangulation sweep connecting τ\tau and φ(τ)\varphi(\tau) and let a=a(0),a(1),,a(k0)=aa=a^{(0)},a^{(1)},\dots,a^{(k_{0})}=a be a periodic edge weight system for this ideal triangulation sweep, defining a φ\varphi-invariant character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}). The classification of irreducible representations involves, in addition to an edge weight system a()6a\in(\mathbb{C}^{*})^{6} for τ\tau, puncture weights pjp_{j}\in\mathbb{C}^{*} for j=1,2,3,4j=1,2,3,4 (see [BWY1, Section 3.4]). We choose the puncture weights such that they are φ\varphi-invariant.

Let ρ:𝒯^τq(S0,4)End(V)\rho:\hat{\mathcal{T}}_{\tau}^{q}(S_{0,4})\to\mathrm{End}(V) be an irreducible representation of the (fractional) Chekhov-Fock algebra of τ\tau that is classified by the edge weight system a()6a\in(\mathbb{C}^{*})^{6} and φ\varphi-invariant puncture weights. The representation ρΦτφ(τ)q:𝒯^φ(τ)q(S0,4)End(V)\rho\circ\Phi_{\tau\varphi(\tau)}^{q}\colon\hat{\mathcal{T}}^{q}_{\varphi(\tau)}(S_{0,4})\to\mathrm{End}(V) is classified by the same edge weight system and puncture weights as ρ\rho. The diffeomorphism φ\varphi induces a natural isomorphism ψφ(τ)τq:𝒯^τq(S0,4)𝒯^φ(τ)q(S0,4)\psi^{q}_{\varphi(\tau)\tau}\colon\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})\to\hat{\mathcal{T}}^{q}_{\varphi(\tau)}(S_{0,4}) sending the generators of 𝒯^τq(S0,4)\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) corresponding to the i-th edge of τ\tau to the generator of 𝒯^φ(τ)q(S0,4)\hat{\mathcal{T}}^{q}_{\varphi(\tau)}(S_{0,4}) corresponding to the i-th edge of φ(τ)\varphi(\tau). The relabelling is done in a consistent way such that the generators always satisfy the relation XiXj=q2σijXjXiX_{i}X_{j}=q^{2\sigma_{ij}}X_{j}X_{i} in both 𝒯^τq(S0,4)\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}) and 𝒯^φ(τ)q(S0,4)\hat{\mathcal{T}}^{q}_{\varphi(\tau)}(S_{0,4}).

We consider the representation ρΦτφ(τ)qΨφ(τ)τq:𝒯^τq(S0,4)End(V)\rho\circ\Phi_{\tau\varphi(\tau)}^{q}\circ\Psi^{q}_{\varphi(\tau)\tau}\colon\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4})\to\mathrm{End}(V). Its invariants are the same edge weight system a()6a\in(\mathbb{C}^{*})^{6} as ρΦτφ(τ)q\rho\circ\Phi_{\tau\varphi(\tau)}^{q}. This implies that the representations ρ\rho and ρΦτφ(τ)qΨφ(τ)τq:𝒯^τqEnd(V)\rho\circ\Phi_{\tau\varphi(\tau)}^{q}\circ\Psi^{q}_{\varphi(\tau)\tau}\colon\hat{\mathcal{T}}^{q}_{\tau}\to\mathrm{End}(V) are isomorphic, by an isomorphism Λφ,r~q:VV\Lambda_{\varphi,\tilde{r}}^{q}\colon V\to V such that

(3.3) (ρΦτφ(τ)qΨφ(τ)τq)(X)=Λφ,r~qρ(X)Λφ,r~q1End(V)\displaystyle\left(\rho\circ\Phi_{\tau\varphi(\tau)}^{q}\circ\Psi^{q}_{\varphi(\tau)\tau}\right)(X)=\Lambda_{\varphi,\tilde{r}}^{q}\circ\rho(X)\circ{\Lambda_{\varphi,\tilde{r}}^{q}}^{-1}\in\mathrm{End}(V)

for every X𝒯^τq(S0,4)X\in\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}). We normalize it such that the determinant detΛφ,r~q=1\det\Lambda_{\varphi,\tilde{r}}^{q}=1.

Proposition 3.4.

[BWY1, Proposition 15] Let Λφ,r~q:VV\Lambda_{\varphi,\tilde{r}}^{q}\colon V\to V be the normalized intertwiner as above. Then, up to conjugation, Λφ,r~q\Lambda_{\varphi,\tilde{r}}^{q} depends only on the diffeomorphism φ\varphi, the ideal triangulation sweep from τ\tau to φ(τ)\varphi(\tau), the periodic edge weight system for this sweep and the φ\varphi-invariant puncture weights.
In particular, |TraceΛφ,r~q||\operatorname{Trace}\Lambda_{\varphi,\tilde{r}}^{q}| is uniquely determined by this data.

We compute the intertwiners for each φk\varphi_{k} in the diffeomorphism φ\varphi 3.1.
Consider the discrete quantum dilogarithm function

QDLq(u,v|j)=vjk=1j(1+uq2k)\displaystyle QDL^{q}(u,v|j)=v^{-j}\prod_{k=1}^{j}(1+uq^{-2k})

and

Dq(u)=j=1nQDLq(u,v|j).\displaystyle D^{q}(u)=\prod_{j=1}^{n}QDL^{q}(u,v|j).

The left and right intertwiners are the isomorphisms that satisfy

(ρxk1,yk1,𝕡(k1))(W)=(ΛL)uk,vk,u^k,v^kρxk,yk,𝕡(k)(ΛL)uk,vk,u^k,v^k1\displaystyle(\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\circ\mathcal{L})(W)=(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}\;\;\rho_{x_{k},y_{k},\mathbb{p}^{(k)}}\;\;(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}^{-1}
(ρxk1,yk1,𝕡(k1))(W)=(ΛR)uk,vk,u^k,v^kρxk,yk,𝕡(k)(ΛR)uk,vk,u^k,v^k1\displaystyle(\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\circ\mathcal{R})(W)=(\Lambda_{R})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}\;\;\rho_{x_{k},y_{k},\mathbb{p}^{(k)}}\;\;(\Lambda_{R})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}^{-1}

for every W𝒯^τq(S0,4)W\in\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}).

Proposition 3.5.

The left and right intertwiners are given by

((ΛL)uk,vk,u^k,v^k)i,j=q(j2i2+4ij+ij)/2QDLq(uk,vk|j)QDL(u^k,v^k|j)nDq(uk)1/nD(u^k)1/n\displaystyle((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}=\frac{q^{(j^{2}-i^{2}+4ij+i-j)/2}QDL^{q}(u_{k},v_{k}|j)QDL(\hat{u}_{k},\hat{v}_{k}|j)}{\sqrt{n}D^{q}(u_{k})^{1/n}D(\hat{u}_{k})^{1/n}}
((ΛR)uk,vk,u^k,v^k)i,j=q(3j2+i24ij+ij)/2QDLq(uk,vk|j)QDL(u^k,v^k|j)nDq(uk)1/nD(u^k)1/n.\displaystyle((\Lambda_{R})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}=\frac{q^{(3j^{2}+i^{2}-4ij+i-j)/2}QDL^{q}(u_{k},v_{k}|j)QDL(\hat{u}_{k},\hat{v}_{k}|j)}{\sqrt{n}D^{q}(u_{k})^{1/n}D(\hat{u}_{k})^{1/n}}.
Proof.

For isomorphism (ΛL)uk,vk,u^k,v^k(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}, suppose yk1n1y_{k-1}^{n}\neq-1. We have uk,vk,u^k,v^ku_{k},v_{k},\hat{u}_{k},\hat{v}_{k}\in\mathbb{C}, such that

uk=qxk1,vkn=1+ukn,u^k=qp3p4hyk1,v^kn=1+u^kn.\displaystyle u_{k}=qx_{k}^{-1},\;\;\;v_{k}^{n}=1+u_{k}^{n},\;\;\;\hat{u}_{k}=q\frac{p_{3}p_{4}}{h}y_{k-1},\;\;\;\hat{v}_{k}^{n}=1+\hat{u}_{k}^{n}.

We do the computation for the \mathcal{L} case first. Evaluation when W=X1W=X_{1}

ρxk1,yk1,𝕡(k1)(X1)(ΛL)uk,vk,u^k,v^k(wj)=ρxk1,yk1,𝕡(k1)(X21)i((ΛL)uk,vk,u^k,v^k)i,jwi\displaystyle\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\circ\mathcal{L}(X_{1})\;\;(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}(w_{j})=\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}(X_{2}^{-1})\sum_{i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}w_{i}
=1yk1i((ΛL)uk,vk,u^k,v^k)i,jqi1wi1=1yk1i((ΛL)uk,vk,u^k,v^k)i+1,jqiwi.\displaystyle=\frac{1}{y_{k-1}}\sum_{i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}q^{i-1}w_{i-1}=\frac{1}{y_{k-1}}\sum_{i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i+1,j}q^{i}w_{i}.

Now other side of the equation:

(ΛL)uk,vk,u^k,v^kρxk,yk,𝕡(k)(X1)(wj)=(ΛL)uk,u^kxkq2jwj=xkq2ji((ΛL)uk,vk,u^k,v^k)i,jwi.\displaystyle(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}\;\;\rho_{x_{k},y_{k},\mathbb{p}^{(k)}}(X_{1})(w_{j})=(\Lambda_{L})_{u_{k},\hat{u}_{k}}x_{k}q^{2j}w_{j}=x_{k}q^{2j}\sum_{i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}w_{i}.

From the above two equations we get:

(3.4) ((ΛL)uk,vk,u^k,v^k)i+1,j=xkyk1q2ji((ΛL)uk,vk,u^k,v^k)i,j\displaystyle((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i+1,j}=x_{k}y_{k-1}q^{2j-i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}
(3.5) ((ΛL)uk,vk,u^k,v^k)i1,j=xk1yk11qi2j1((ΛL)uk,vk,u^k,v^k)i,j.\displaystyle((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i-1,j}=x_{k}^{-1}y_{k-1}^{-1}q^{i-2j-1}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}.

Evaluation with W=X2W=X_{2}:

ρxk1,yk1,𝕡(k1)(X2)(ΛL)uk,vk,u^k,v^k(wj)\displaystyle\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\circ\mathcal{L}(X_{2})\cdot(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}(w_{j})
=ρxk1,yk1,𝕡(k1)((1+qX2)(1+qX5)X4)i((ΛL)uk,vk,u^k,v^k)i,jwi\displaystyle=\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\left((1+qX_{2})\left(1+qX_{5}\right)X_{4}\right)\sum_{i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}w_{i}
=p2p4xk1hρxk1,yk1,𝕡(k1)((1+qX2)(1+qH1P3P4X2))iq2i((ΛL)uk,vk,u^k,v^k)i,jwi\displaystyle=\frac{p_{2}p_{4}x_{k-1}}{h}\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}\left((1+qX_{2})\left(1+qH^{-1}P_{3}P_{4}X_{2}\right)\right)\sum_{i}q^{2i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}w_{i}
=p2p4xk1h(1+p3p4q12jhxk)ρxk1,yk1,𝕡(k1)(1+qX2)iq2i((ΛL)uk,vk,u^k,v^k)i,jwi\displaystyle=\frac{p_{2}p_{4}x_{k-1}}{h}\left(1+\frac{p_{3}p_{4}q^{-1-2j}}{hx_{k}}\right)\rho_{x_{k-1},y_{k-1},\mathbb{p}^{(k-1)}}(1+qX_{2})\sum_{i}q^{2i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}w_{i}
=ip2p4xk1hq2i(1+p3p4q12jhxk)(1+q12jxk)((ΛL)uk,vk,u^k,v^k)i,jwi\displaystyle=\sum_{i}\frac{p_{2}p_{4}x_{k-1}}{h}q^{2i}\left(1+\frac{p_{3}p_{4}q^{-1-2j}}{hx_{k}}\right)\left(1+\frac{q^{-1-2j}}{x_{k}}\right)((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}w_{i}

where we used (3.4) and (3.5) multiple times.

(ΛL)uk,vk,u^k,v^kρxk,yk,𝕡(k)(X2)(wj)\displaystyle(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}\cdot\rho_{x_{k},y_{k},\mathbb{p}^{(k)}}(X_{2})(w_{j}) =(ΛL)uk,vk,u^k,v^kykqjwj+1\displaystyle=(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}y_{k}q^{-j}w_{j+1}
=ykqji((ΛL)uk,vk,u^k,v^k)i,j+1wi.\displaystyle=y_{k}q^{-j}\sum_{i}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j+1}w_{i}.

Thus,

(3.6) ((ΛL)uk,vk,u^k,v^k)i,j+1=p2p4xk1q2i+j(1+q12jxk)(1+hq12jxkp3p4)hyk((ΛL)uk,vk,u^k,v^k)i,j.\displaystyle((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j+1}=\frac{p_{2}p_{4}x_{k-1}q^{2i+j}\left(1+\frac{q^{-1-2j}}{x_{k}}\right)\left(1+\frac{hq^{-1-2j}}{x_{k}p_{3}p_{4}}\right)}{hy_{k}}((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j}.

Choose

((ΛL)uk,vk,u^k,v^k)0,0=1nDq(uk)1/nD(u^k)1/n.((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{0,0}=\frac{1}{\sqrt{n}D^{q}(u_{k})^{1/n}D(\hat{u}_{k})^{1/n}}\;.

By (3.4) and (3.6), we get:

((ΛL)uk,vk,u^k,v^k)i,0\displaystyle((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,0} =qi(i1)/2\displaystyle=q^{-i(i-1)/2}
((ΛL)uk,vk,u^k,v^k)i,j\displaystyle((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,j} =(p2p4xk1q2ihyk)j(qj(j1)/2l=0j1(1+q12lxk)(1+hq12lxkp3p4))\displaystyle=\left(\frac{p_{2}p_{4}x_{k-1}q^{2i}}{hy_{k}}\right)^{j}\left(q^{j(j-1)/2}\prod_{l=0}^{j-1}\left(1+\frac{q^{-1-2l}}{x_{k}}\right)\left(1+\frac{hq^{-1-2l}}{x_{k}p_{3}p_{4}}\right)\right)
×((ΛL)uk,vk,u^k,v^k)i,0\displaystyle\times((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,0}
=(p2p4xk1q2ihyk)j(qj(j1)/2l=1j(1+p1p2yk1q12lh)(1+yk1q12l))\displaystyle=\left(\frac{p_{2}p_{4}x_{k-1}q^{2i}}{hy_{k}}\right)^{j}\left(q^{j(j-1)/2}\prod_{l=1}^{j}\left(1+\frac{p_{1}p_{2}y_{k-1}q^{1-2l}}{h}\right)\left(1+y_{k-1}q^{1-2l}\right)\right)
×((ΛL)uk,vk,u^k,v^k)i,0\displaystyle\times((\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}})_{i,0}
=qj2i2+4ij+ij2QDLq(uk,vk|j)QDLq(u^k,v^k|j)1nDq(uk)1/nD(u^k)1/n.\displaystyle=q^{\frac{j^{2}-i^{2}+4ij+i-j}{2}}QDL^{q}(u_{k},v_{k}|j)QDL^{q}(\hat{u}_{k},\hat{v}_{k}|j)\frac{1}{\sqrt{n}D^{q}(u_{k})^{1/n}D(\hat{u}_{k})^{1/n}}\;.

A similar computation with W=X1W=X_{1} and X2X_{2} gives the result for \mathcal{R}. ∎

Remark 3.6.

For the generators ψ1\psi_{1} and ψ2\psi_{2}, the isomorphic representations correspond to an edge-re-indexing. It’s easy to check that the intertwiner is trivial for any edge re-indexing. These generators add to the periodic edge weight system up to two tuples (a1(k0+1),,a6(k0+1))(a_{1}^{(k_{0}+1)},\dots,a_{6}^{(k_{0}+1)}) and (a1(k0+2),,a6(k0+2))(a_{1}^{(k_{0}+2)},\dots,a_{6}^{(k_{0}+2)}), that are related to the last one (a1(k0),,a6(k0))(a_{1}^{(k_{0})},\dots,a_{6}^{(k_{0})}) by a permutation. We will ignore the (possible) additional permutation for proof purposes. We add a remark every time this permutation comes up, but it will not affect the proof.

For computation, we will work with the diffeomorphism where the first generator is LL. We have a periodic edge weight system, that is, (a0,b0)=(ak0,bk0)(a_{0},b_{0})=(a_{k_{0}},b_{k_{0}}).

Proposition 3.7.

For the mapping torus, with the identification of (x,1)(ϕ(x),0)(x,1)\sim(\phi(x),0) where ϕ:ΣΣ\phi\colon\Sigma\rightarrow\Sigma is the surface diffeomorphism, we get a twist intertwiner T:nnT\colon\mathbb{C}^{n}\rightarrow\mathbb{C}^{n} such that ρx0,y0,𝕡(0)(W)=Tρxk0,yk0,𝕡(k0)(W)T1\rho_{x_{0},y_{0},\mathbb{p}^{(0)}}(W)=T\circ\rho_{x_{k_{0}},y_{k_{0}},\mathbb{p}^{(k_{0})}}(W)\circ T^{-1} for every W𝒯^τq(S0,4)W\in\hat{\mathcal{T}}^{q}_{\tau}(S_{0,4}). Then,

(3.7) Ti,j=q2jl2jl1δi,j+l1\displaystyle T_{i,j}=q^{-2jl_{2}-jl_{1}}\delta_{i,j+l_{1}}

where the indices are modulo n and x0=q2l1xk0x_{0}=q^{2l_{1}}x_{k_{0}} and y0=q2l2yk0y_{0}=q^{2l_{2}}y_{k_{0}}.

Proof.

Follows from direct calculation. ∎

3.3. Logarithm Choices

For the edge weight system of the triangulation sweep, we need to make choices of logarithms to make things well-defined.

We fix complex numbers Ak,BkA_{k},B_{k}\in\mathbb{C}, such that ak=eAk,bk=eBka_{k}=e^{A_{k}},b_{k}=e^{B_{k}}. Fix θj\theta_{j}\in\mathbb{C}, such that we have Trace(αj)=eθjeθj=pjnpjn(\alpha_{j})=-e^{\theta_{j}}-e^{-\theta_{j}}=-p_{j}^{n}-p_{j}^{-n} where αjπ1(S0,4)\alpha_{j}\in\pi_{1}(S_{0,4}) is the peripheral curve around the puncture and pjp_{j}\in\mathbb{C} is the φ\varphi-invariant puncture weight labelled according to an ideal triangulation τ\tau. We therefore have eθj=pjne^{\theta_{j}}=p_{j}^{n} for j{1,2,3,4}j\in\{1,2,3,4\}. We fix hn=e(θ1+θ2+θ3+θ4)/2.h^{n}=e^{(\theta_{1}+\theta_{2}+\theta_{3}+\theta_{4})/2}.
For each k={1,,k0}k=\{1,\dots,k_{0}\}, we fix Uk,Vk,U^k,V^kU_{k},V_{k},\hat{U}_{k},\hat{V}_{k}\in\mathbb{C} such that

eUk=ak1\displaystyle e^{U_{k}}=a_{k}^{-1} eVk=1+eUk\displaystyle e^{V_{k}}=1+e^{U_{k}}
eU^k={hn(p1(k1))n(p2(k1))nak1 if φk=Lhn(p1(k1))n(p4(k1))nak1 if φk=R\displaystyle e^{\hat{U}_{k}}=\begin{cases}h^{-n}(p^{(k-1)}_{1})^{n}(p^{(k-1)}_{2})^{n}a_{k}^{-1}&\text{ if }\varphi_{k}=L\\ h^{n}(p^{(k-1)}_{1})^{-n}(p^{(k-1)}_{4})^{-n}a_{k}^{-1}&\text{ if }\varphi_{k}=R\end{cases} eV^k=1+eV^k.\displaystyle e^{\hat{V}_{k}}=1+e^{\hat{V}_{k}}.

From 3.1, we define

:Ak=Bk1\displaystyle\mathcal{L}\colon A_{k}=-B_{k-1} Bk=Vk+V~k+(θ2(k1)+θ4(k1)θ1(k1)θ3(k1))/2+Ak1\displaystyle B_{k}=V_{k}+\tilde{V}_{k}+(\theta_{2}^{(k-1)}+\theta_{4}^{(k-1)}-\theta_{1}^{(k-1)}-\theta_{3}^{(k-1)})/2+A_{k-1}
:Ak=θ1(k1)+Ak1+Bk1\displaystyle\mathcal{R}\colon A_{k}=-\theta^{(k-1)}_{1}+A_{k-1}+B_{k-1} Bk=Vk+V~k+Bk1.\displaystyle B_{k}=V_{k}+\tilde{V}_{k}+B_{k-1}.

Since the edge weight system is periodic, we get

eAk0=eA0\displaystyle e^{A_{k_{0}}}=e^{A_{0}} eBk0=eB0.\displaystyle e^{B_{k_{0}}}=e^{B_{0}}.

This implies there exists integers l^1,l^2\hat{l}_{1},\hat{l}_{2}\in{\mathbb{Z}} such that

A0=Ak0+2πil^1\displaystyle A_{0}=A_{k_{0}}+2\pi i\hat{l}_{1} B0=Bk0+2πil^2.\displaystyle B_{0}=B_{k_{0}}+2\pi i\hat{l}_{2}.

Note that we need x0=q2l1xk0x_{0}=q^{2{l}_{1}}x_{k_{0}} and y0=q2l2yk0y_{0}=q^{2{l}_{2}}y_{k_{0}}. This implies

eA0n=eAk0n+2πil^1n=xk0q2l1\displaystyle e^{\frac{A_{0}}{n}}=e^{\frac{A_{k_{0}}}{n}+\frac{2\pi i\hat{l}_{1}}{n}}=x_{k_{0}}q^{2{l}_{1}} eB0n=eBk0n+2πil^2n=yk0q2l2.\displaystyle e^{\frac{B_{0}}{n}}=e^{\frac{B_{k_{0}}}{n}+\frac{2\pi i\hat{l}_{2}}{n}}=y_{k_{0}}q^{2{l}_{2}}.

Thus,

l1=l^1(n1)22\displaystyle l_{1}=\hat{l}_{1}\frac{(n-1)^{2}}{2} l2=l^2(n1)22.\displaystyle l_{2}=\hat{l}_{2}\frac{(n-1)^{2}}{2}.

3.4. Explicit formula for the invariant

Proposition 3.8.

For the intertwiner Λφ,r~q\Lambda_{\varphi,\tilde{r}}^{q} and uk,u^k,vk,v^ku_{k},\hat{u}_{k},v_{k},\hat{v}_{k}\in\mathbb{C} defined above,

Trace(Λφ,r~q)=Ani1,,ik0=1n\displaystyle\operatorname{Trace}(\Lambda_{\varphi,\tilde{r}}^{q})=\text{A}_{n}\sum_{i_{1},\dots,i_{k_{0}}=1}^{n} exp(2πin(ik0l^2i1l^1+k=1k0il+12+ϵk+12(il2+il+124ilil+1)))\displaystyle\exp{\left(\frac{2\pi i}{n}\left(-i_{k_{0}}\hat{l}_{2}-i_{1}\hat{l}_{1}+\sum_{k=1}^{k_{0}}i_{l+1}^{2}+\frac{\epsilon_{k+1}}{2}\left(i_{l}^{2}+i_{l+1}^{2}-4i_{l}i_{l+1}\right)\right)\right)}
(3.8) ×k=1k0QDLq(uk,vk|ik)QDLq(u^k,v^k|ik).\displaystyle\times\prod_{k=1}^{k_{0}}QDL^{q}(u_{k},v_{k}|i_{k})\;QDL^{q}(\hat{u}_{k},\hat{v}_{k}|i_{k}).

where ϵk=1\epsilon_{k}=1 if Λk=(ΛR)uk,vk,u^k,v^k\Lambda_{k}=(\Lambda_{R})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}, and ϵk=1\epsilon_{k}=-1 if Λk=(ΛL)uk,vk,u^k,v^k\Lambda_{k}=(\Lambda_{L})_{u_{k},v_{k},\hat{u}_{k},\hat{v}_{k}}, ϵk0+1=ϵ1,ik0+1=i1\epsilon_{k_{0}+1}=\epsilon_{1},\;i_{k_{0}+1}=i_{1} and

An=ql^1l^2/2l^1/4nk0/2l=1k0Dq(ul)1/nD(u^l)1/n.\displaystyle\text{A}_{n}=\frac{q^{\hat{l}_{1}\hat{l}_{2}/2-\hat{l}_{1}/4}}{n^{k_{0}/2}\prod_{l=1}^{k_{0}}D^{q}(u_{l})^{1/n}D(\hat{u}_{l})^{1/n}}.
Proof.

We know that the intertwiner can be written as

Λφ,r~q=Λ1Λk0Tl1l2.\displaystyle\Lambda_{\varphi,\tilde{r}}^{q}=\Lambda_{1}\circ\dots\circ\Lambda_{k_{0}}\circ T_{l_{1}l_{2}}.

Observe that

2j2+ij+ϵk(j2+i24ij)\displaystyle 2j^{2}+i-j+\epsilon_{k}(j^{2}+i^{2}-4ij)

gives us the power of qq for both \mathcal{L} and \mathcal{R} by letting ϵk=1\epsilon_{k}=-1 for \mathcal{L} and ϵk=1\epsilon_{k}=1 for \mathcal{R}.
Thus, the trace of this intertwiner can be written as

Trace(Λφ,r~q)=\displaystyle\operatorname{Trace}(\Lambda_{\varphi,\tilde{r}}^{q})= i0,,ik0=1n(Λ1)i0,i1(Λ2)i1i2(Λk0)ik01ik0(Tl1l2)ik0i0\displaystyle\sum_{i_{0},\dots,i_{k_{0}}=1}^{n}(\Lambda_{1})_{i_{0},i_{1}}(\Lambda_{2})_{i_{1}i_{2}}\dots(\Lambda_{k_{0}})_{i_{k_{0}-1}i_{k_{0}}}(T_{l_{1}l_{2}})_{i_{k_{0}}i_{0}}
=\displaystyle= i0,,ik0nq2i0l2i0l1+i0ik02k=1k0qik2+ϵk2((ikik1)22ikik1)\displaystyle\sum_{i_{0},\dots,i_{k_{0}}}^{n}q^{-2i_{0}l_{2}-i_{0}l_{1}+\frac{i_{0}-i_{k_{0}}}{2}}\prod_{k=1}^{k_{0}}q^{i_{k}^{2}+\frac{\epsilon_{k}}{2}((i_{k}-i_{k-1})^{2}-2i_{k}i_{k-1})}
×QDLq(uk,vk|ik)QDLq(u^k,v^k|ik)nDq(uk)1/nD(u^k)1/nδik0,i0+l1.\displaystyle\times\frac{QDL^{q}(u_{k},v_{k}|i_{k})QDL^{q}(\hat{u}_{k},\hat{v}_{k}|i_{k})}{\sqrt{n}D^{q}(u_{k})^{1/n}D(\hat{u}_{k})^{1/n}}\delta_{i_{k_{0}},i_{0}+l_{1}}.
Trace(Λφ,r~q)=\displaystyle\operatorname{Trace}(\Lambda_{\varphi,\tilde{r}}^{q})= Ani0,,ik0nq2(ik0l1)l2(ik0l1+12)l1+i12+ϵ12((i1+l1ik0)22i1(ik0l1))\displaystyle A^{\prime}_{n}\sum_{i_{0},\dots,i_{k_{0}}}^{n}q^{-2(i_{k_{0}}-l_{1})l_{2}-(i_{k_{0}}-l_{1}+\frac{1}{2})l_{1}+i_{1}^{2}+\frac{\epsilon_{1}}{2}((i_{1}+l_{1}-i_{k_{0}})^{2}-2i_{1}(i_{k_{0}}-l_{1}))}
×qk=2k0(ik2+ϵk2(ik2+ik124ikik1))k=1k0QDLq(uk,vk|ik)QDLq(u^k,v^k|ik)\displaystyle\times q^{\sum_{k=2}^{k_{0}}(i_{k}^{2}+\frac{\epsilon_{k}}{2}(i_{k}^{2}+i_{k-1}^{2}-4i_{k}i_{k-1}))}\prod_{k=1}^{k_{0}}QDL^{q}(u_{k},v_{k}|i_{k})QDL^{q}(\hat{u}_{k},\hat{v}_{k}|i_{k})
=\displaystyle= Ani1,,ik0exp(2πin(ik0l^2i1l^1+k=1k0il+12+ϵk+12(il2+il+124ilil+1)))\displaystyle A_{n}\sum_{i_{1},\dots,i_{k_{0}}}\exp{\left(\frac{2\pi i}{n}\left(-i_{k_{0}}\hat{l}_{2}-i_{1}\hat{l}_{1}+\sum_{k=1}^{k_{0}}i_{l+1}^{2}+\frac{\epsilon_{k+1}}{2}\left(i_{l}^{2}+i_{l+1}^{2}-4i_{l}i_{l+1}\right)\right)\right)}
×k=1k0QDLq(uk,vk|ik)QDLq(u^k,v^k|ik).\displaystyle\times\prod_{k=1}^{k_{0}}QDL^{q}(u_{k},v_{k}|i_{k})\;QDL^{q}(\hat{u}_{k},\hat{v}_{k}|i_{k}).

4. Quantum Dilogarithms

Now that we have the explicit algebraic formula for the intertwiner, we try to write the discrete dilogarithm function as a restriction of a continuous one. We use [CF] to recall definitions of different dilogarithm functions.

4.1. Fadeev quantum dilogarithm

The small continuous quantum dilogarithm for >0\hbar>0 is the function:

(4.1) li2(z)=2πiΩe(2zπ)t4tsinh(πt)sin(πt)𝑑t\displaystyle li_{2}^{\hbar}(z)=2\pi i\hbar\int_{\Omega}\frac{e^{(2z-\pi)t}}{4tsinh(\pi t)sin(\pi\hbar t)}dt

where the domain Ω\Omega is the real line with a small interval around 0 replaced by a semi-circle above the real line in the positive imaginary direction. The integral converges for π2<𝔢z<π+π2-\frac{\pi\hbar}{2}<\mathfrak{Re}\;z<\pi+\frac{\pi\hbar}{2}. The big continuous quantum dilogarithm is defined as

Li2(z)=e12πili2(z)\displaystyle Li_{2}^{\hbar}(z)=e^{\frac{1}{2\pi i\hbar}li_{2}^{\hbar}(z)}

satisfying

(4.2) Li2(z+π)\displaystyle Li_{2}^{\hbar}(z+\pi\hbar) =(1e2iz+πi)1Li2(z)\displaystyle=(1-e^{2iz+\pi i\hbar})^{-1}Li_{2}^{\hbar}(z)
(4.3) Li2(z+π)\displaystyle Li_{2}^{\hbar}(z+\pi) =(1+e2iz)1Li2(z).\displaystyle=(1+e^{\frac{2iz}{\hbar}})^{-1}Li_{2}^{\hbar}(z).

The classical dilogarithm function is defined by

li2(z)=0zlog(1t)t𝑑t\displaystyle li_{2}(z)=-\int_{0}^{z}\frac{\log(1-t)}{t}dt

for the principal branch of the logarithm. We can write the small continuous quantum dilogarithm function in terms of the classical dilogarithm function as follows:

Proposition 4.1.

For every z with 0<𝔢z<π0<\mathfrak{Re}\;z<\pi

(4.4) li2(z)=li2(e2iz)+O(2)\displaystyle li_{2}^{\hbar}(z)=li_{2}(e^{2iz})+O(\hbar^{2})

as 0\hbar\rightarrow 0, and this uniformly on compact subsets of the strip {z;0<ez<π}\{z\in\mathbb{C};0<\mathfrak{R}e\;z<\pi\}. ∎

We use equation (4.2) multiple times to get the following lemma

Lemma 4.2.

If q=e2πinq=e^{\frac{2\pi i}{n}}, and u=qe1nUu=qe^{\frac{1}{n}U}, then for every integer j0j\geq 0

(4.5) k=1j(1+uq2k)=Li22n(π2πn+U2ni2πjn)Li22n(π2πn+U2ni).\displaystyle\prod_{k=1}^{j}(1+uq^{-2k})=\frac{Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\frac{2\pi j}{n})}{Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni})}.

Recall that QDLq(u,v|j)=vjk=1j(1+uq2k)QDL^{q}(u,v|j)=v^{-j}\prod_{k=1}^{j}(1+uq^{-2k}). The function li22n(z)li_{2}^{\frac{2}{n}}(z) is only defined in the interval πn<ez<π+πn-\frac{\pi}{n}<\mathfrak{R}e\;z<\pi+\frac{\pi}{n}. Thus we need

πn<π2πn+U2ni2πjn<π+πn.-\frac{\pi}{n}<\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\frac{2\pi j}{n}<\pi+\frac{\pi}{n}\;.

This implies

π2<2πjn<π2.-\frac{\pi}{2}<\frac{2\pi j}{n}<\frac{\pi}{2}\;.

For the interval (π2,3π2)\left(\frac{\pi}{2},\frac{3\pi}{2}\right), we use (4.3). We can always choose the interval (π2,3π2)\left(-\frac{\pi}{2},\frac{3\pi}{2}\right) for 2πjn\frac{2\pi j}{n} because QDLq(u,v|j)QDL^{q}(u,v|j) and QDLq(u^,v^|j)QDL^{q}(\hat{u},\hat{v}|j) are n periodic, which means jj and j+nj+n will correspond to the same QDLqQDL^{q}. In other words, whenever 2π>2πjn>3π22\pi>\frac{2\pi j}{n}>\frac{3\pi}{2}, we replace jj with jnj-n, so that 0>2π(jn)n>π20>\frac{2\pi(j-n)}{n}>-\frac{\pi}{2}

We define

I1=(π2,π2)\displaystyle I^{1}=\left(-\frac{\pi}{2},\frac{\pi}{2}\right) I2=(π2,3π2)\displaystyle I^{2}=\left(\frac{\pi}{2},\frac{3\pi}{2}\right)

and an analytic function in the domain (π/2,3π/2)(-\pi/2,3\pi/2),

QDLq(u,v|α)QD\displaystyle QDL^{q}(u,v|\alpha)QD Lq(u^,v^|α)=\displaystyle L^{q}(\hat{u},\hat{v}|\alpha)=
{eα(V+V~)2πLi22n(π2πn+U2niα)Li22n(π2πn+U^2niα)Li22n(π2πn+U2ni)Li22n(π2πn+U^2ni) if αI1(1ineU2)(1ineU^2)eα(V+V~)2πLi22n(π2πn+U2niα)Li22n(π2πn+U^2niα)Li22n(π2πn+U2ni)Li22n(π2πn+U^2ni) if αI2.\displaystyle\begin{cases}e^{-\frac{\alpha(V+\tilde{V})}{2\pi}}\frac{Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\alpha)Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{\hat{U}}{2ni}-\alpha)}{Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni})Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{\hat{U}}{2ni})}&\text{ if }\alpha\in I^{1}\\ (1-i^{n}e^{\frac{U}{2}})(1-i^{n}e^{\frac{\hat{U}}{2}})e^{-\frac{\alpha(V+\tilde{V})}{2\pi}}\frac{Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\alpha)Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{\hat{U}}{2ni}-\alpha)}{Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni})Li_{2}^{\frac{2}{n}}(\frac{\pi}{2}-\frac{\pi}{n}+\frac{\hat{U}}{2ni})}&\text{ if }\alpha\in I^{2}.\end{cases}

5. Analytic formula for the invariant

This section provides an analytic formula for the invariant so that techniques like the Poisson summation formula and saddle point method can be used later.
The term in the denominator Dq(u)1nD^{q}(u)^{\frac{1}{n}} is a ratio of hyperbolic sin or cos functions (based on n mod 4) and has a finite limit as nn\rightarrow\infty [BWY2].

Proposition 5.1.

Let U,U^U,\hat{U}\in\mathbb{C} with eU1eU^e^{U}\neq 1\neq e^{\hat{U}} and for every odd n, q=e2πin,u=e1nU,u^=e1nU^q=e^{\frac{2\pi i}{n}},u=e^{\frac{1}{n}U},\hat{u}=e^{\frac{1}{n}\hat{U}}, then

limnn=1mod 4|Dq(u)Dq(u^)|1n\displaystyle\lim_{\begin{subarray}{c}n\rightarrow\infty\\ n=1\;\text{mod}\;4\end{subarray}}|D^{q}(u)D^{q}(\hat{u})|^{\frac{1}{n}} =2mA+mA~4π|coshAπi4coshA~πi4coshA+πi4coshA~+πi4|14\displaystyle=2^{-\frac{\mathfrak{I}m\;A+\mathfrak{I}m\;\tilde{A}}{4\pi}}\left|\frac{\cosh\frac{A-\pi i}{4}\cosh\frac{\tilde{A}-\pi i}{4}}{\cosh\frac{A+\pi i}{4}\cosh\frac{\tilde{A}+\pi i}{4}}\right|^{\frac{1}{4}}
limnn=3 mod 4|Dq(u)Dq(u^)|1n\displaystyle\lim_{\begin{subarray}{c}n\rightarrow\infty\\ n=3\text{ mod }4\end{subarray}}|D^{q}(u)D^{q}(\hat{u})|^{\frac{1}{n}} =2mA+mA~4π|sinhAπi4sinhA~πi4sinhA+πi4sinhA~+πi4|14.\displaystyle=2^{-\frac{\mathfrak{I}m\;A+\mathfrak{I}m\;\tilde{A}}{4\pi}}\left|\frac{\sinh\frac{A-\pi i}{4}\sinh\frac{\tilde{A}-\pi i}{4}}{\sinh\frac{A+\pi i}{4}\sinh\frac{\tilde{A}+\pi i}{4}}\right|^{\frac{1}{4}}.

This is an important part of the computation as we see the terms are finite and do not grow exponentially with nn.
Define linear maps 𝕂:k0k0,𝕃:k0\mathbb{K}\colon{\mathbb{R}}^{k_{0}}\rightarrow{\mathbb{R}}^{k_{0}},\mathbb{L}\colon{\mathbb{R}}^{k_{0}}\rightarrow{\mathbb{R}} as

Kk(𝜶)=ϵk+ϵk+1+22αkϵk+1αk+1ϵkαk1,\displaystyle K_{k}(\bm{\alpha})=\frac{\epsilon_{k}+\epsilon_{k+1}+2}{2}\alpha_{k}-\epsilon_{k+1}\alpha_{k+1}-\epsilon_{k}\alpha_{k-1}, 𝕃(𝜶)=l^2αk0l^1α1.\displaystyle\mathbb{L}(\bm{\alpha})=-\hat{l}_{2}\alpha_{k_{0}}-\hat{l}_{1}\alpha_{1}.

We can rewrite the invariant 3.8 as

(5.1) Trace Λφ,rq=Ani1,,ik0=1nFn(2πi1n,,2πik0n)\displaystyle\text{Trace }\Lambda_{\varphi,r}^{q}=A_{n}\sum_{i_{1},\dots,i_{k_{0}}=1}^{n}F_{n}\left(\frac{2\pi i_{1}}{n},\dots,\frac{2\pi i_{k_{0}}}{n}\right)

for a map Fn:k0F_{n}\colon{\mathbb{R}}^{k_{0}}\rightarrow\mathbb{C}, where

(5.2) Fn(𝜶)\displaystyle F_{n}(\bm{\alpha}) =(k=1k0QDLq(uk,vk|αk)QDLq(u^k,v^k|αk))exp(ni2π𝕂(𝜶)𝜶+i𝕃(𝜶)),\displaystyle=\left(\prod_{k=1}^{k_{0}}QDL^{q}(u_{k},v_{k}|\alpha_{k})QDL^{q}(\hat{u}_{k},\hat{v}_{k}|\alpha_{k})\right)\exp\left(\frac{ni}{2\pi}\mathbb{K}(\bm{\alpha})\cdot\bm{\alpha}+i\mathbb{L}(\bm{\alpha})\right),

and

An\displaystyle A_{n} =q14(2l^1l^2l^1)nk0/2l=1k0Dq(ul)1/nD(u^l)1/n.\displaystyle=\frac{q^{\frac{1}{4}(2\hat{l}_{1}\hat{l}_{2}-\hat{l}_{1})}}{n^{k_{0}/2}\prod_{l=1}^{k_{0}}D^{q}(u_{l})^{1/n}D(\hat{u}_{l})^{1/n}}\;.

We will study the product of QDLQDL functions QDLq(u,v|α)QDLq(u^,v^|α)QDL^{q}(u,v|\alpha)QDL^{q}(\hat{u},\hat{v}|\alpha) . First, recall

Li22n(π2πn+U2niα)=exp(n4πili22n(π2πn+U2niα))\displaystyle Li_{2}^{\frac{2}{n}}\left(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\alpha\right)=\exp\left(\frac{n}{4\pi i}li_{2}^{\frac{2}{n}}\left(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\alpha\right)\right)

and

li22n(π2πn+U2niα)=li2(e2iα)+2πiUnlog(1+e2iα)+O(1n2).\displaystyle li_{2}^{\frac{2}{n}}\left(\frac{\pi}{2}-\frac{\pi}{n}+\frac{U}{2ni}-\alpha\right)=li_{2}\left(-e^{-2i\alpha}\right)+\frac{2\pi i-U}{n}\log(1+e^{-2i\alpha})+O\left(\frac{1}{n^{2}}\right).

Therefore, we get

QDLq(u,v|α)QDLq(u^,v^|α)=g~n(α)exp(n4πif~(α)+O(1n))\displaystyle QDL^{q}(u,v|\alpha)QDL^{q}(\hat{u},\hat{v}|\alpha)=\tilde{g}_{n}(\alpha)\exp\left(\frac{n}{4\pi i}\tilde{f}(\alpha)+O\left(\frac{1}{n}\right)\right)

where

f~(α)=2li2(e2iα)+π26\displaystyle\tilde{f}(\alpha)=2\operatorname{li_{2}}(-e^{2i\alpha})+\frac{\pi^{2}}{6}

and

g~n(α)={(1+e2iα2)1U+U^4πieαV+V~2π if αI1(1ineU2)(1ineU^2)(1+e2iα2)1U+U^4πieαV+V~2π if αI2.\displaystyle\tilde{g}_{n}(\alpha)=\begin{cases}\left(\frac{1+e^{-2i\alpha}}{2}\right)^{1-\frac{U+\hat{U}}{4\pi i}}e^{-\alpha\frac{V+\tilde{V}}{2\pi}}&\text{ if }\alpha\in I^{1}\\ (1-i^{n}e^{\frac{U}{2}})(1-i^{n}e^{\frac{\hat{U}}{2}})\left(\frac{1+e^{-2i\alpha}}{2}\right)^{1-\frac{U+\hat{U}}{4\pi i}}e^{-\alpha\frac{V+\tilde{V}}{2\pi}}&\text{ if }\alpha\in I^{2}.\end{cases}

Thus, for 𝜶=(α1,,αk0)\bm{\alpha}=(\alpha_{1},\dots,\alpha_{k_{0}}), we can rewrite the function Fn(𝜶)F_{n}(\bm{\alpha}) as

(5.3) Fn(𝜶)=gn(𝜶)exp(n4πif(𝜶)+O(1n))\displaystyle F_{n}(\bm{\alpha})=g_{n}(\bm{\alpha})\exp\left(\frac{n}{4\pi i}f(\bm{\alpha})+O\left(\frac{1}{n}\right)\right)

where

(5.4) gn(𝜶)=ei𝕃(𝜶)k=1k0g~n(αk)\displaystyle g_{n}(\bm{\alpha})=e^{i\mathbb{L}(\bm{\alpha})}\prod_{k=1}^{k_{0}}\tilde{g}_{n}(\alpha_{k})

and

(5.5) f(𝜶)=k=1k02li2(e2iαk)+k0π262𝕂(𝜶)𝜶.\displaystyle f(\bm{\alpha})=\sum_{k=1}^{k_{0}}2\operatorname{li_{2}}(-e^{-2i\alpha_{k}})+\frac{k_{0}\pi^{2}}{6}-2\mathbb{K}(\bm{\alpha})\cdot\bm{\alpha}.

6. Poisson Summation Formula

The main result of this section is Proposition 6.5. We change the sum from finite to infinite using a smooth bump function and then use the Poisson summation formula.

Lemma 6.1.

Let U,U^,V,V~U,\hat{U},V,\tilde{V}\in\mathbb{C} with eU=un,eU^=u^n,eV=1+eU,eV~=1+eU^e^{U}=u^{n},e^{\hat{U}}=\hat{u}^{n},e^{V}=1+e^{U},e^{\tilde{V}}=1+e^{\hat{U}} be given. Then for every δ>0\delta>0 sufficiently small and for n sufficiently large,

|QDLq(u,v|α)QDLq(u^,v^|α)|=κenO1(δlogδ)\displaystyle|QDL^{q}(u,v|\alpha)QDL^{q}(\hat{u},\hat{v}|\alpha)|=\kappa e^{nO_{1}(\delta\log\delta)}

where κ>0\kappa>0 is a constant that does not grow exponentially in nn, for every α(π2,3π2)\alpha\in(-\frac{\pi}{2},\frac{3\pi}{2}) that is at δ\delta distance from the bridge points π2,π2,3π2-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2}.

Proof.

We use the fact that mli2(e2iδ)=2Λ(δ)\mathfrak{I}m\;\text{li}_{2}(e^{2i\delta})=2\Lambda(\delta) and Λ(δ)=O(δlogδ)\Lambda(\delta)=O(\delta\log\delta). Here, Λ(θ)\Lambda(\theta) is the Lobachevsky function defined by

Λ(θ)=0θlog|2sint|dt\displaystyle\Lambda(\theta)=-\int_{0}^{\theta}\log|2\sin t|dt

for θ[0,π]\theta\in[0,\pi]. Observe that Λ(0)=Λ(π/2)=0\Lambda(0)=\Lambda(\pi/2)=0 and Λ(π+θ)=Λ(θ)\Lambda(\pi+\theta)=\Lambda(\theta).
If α=π2±δ\alpha=\frac{\pi}{2}\pm\delta,

|QDLq(u,v|α)QDLq(u^,v^|α)|=|(1e±2iδ2)1U+U^4πieV+V~2π(±δπ2)|exp(n2πO1(δlogδ)).\displaystyle|QDL^{q}(u,v|\alpha)QDL^{q}(\hat{u},\hat{v}|\alpha)|=\left|\left(\frac{1-e^{\pm 2i\delta}}{2}\right)^{1-\frac{U+\hat{U}}{4\pi i}}e^{\frac{V+\tilde{V}}{2\pi}(\pm\delta-\frac{\pi}{2})}\right|\exp\left(\frac{n}{2\pi}O_{1}(\delta\log\delta)\right).

The term in front of the exponential is a constant that does not grow exponentially in nn. Similarly, when α=π2+δ\alpha=-\frac{\pi}{2}+\delta we get the same result. For the other case, that is, α=3π2δ\alpha=\frac{3\pi}{2}-\delta, we use equation (4.3) and it only changes the constant term. ∎

Thus, near the bridge points π2,π2,3π2-\tfrac{\pi}{2},\tfrac{\pi}{2},\tfrac{3\pi}{2}, the estimate is small.
We extend the domain of the function f(𝜶)f(\bm{\alpha}) from {\mathbb{R}} to \mathbb{C}. The function f(𝜶)f(\bm{\alpha}) is well-defined and holomorphic in the domain {αk|e(αk)(π2,3π2)}\{\alpha_{k}\in\mathbb{C}\big{|}\mathfrak{R}e\;(\alpha_{k})\in(-\tfrac{\pi}{2},\tfrac{3\pi}{2})\}. This follows straight from the domain of the function li22n(z)\text{li}_{2}^{\frac{2}{n}}(z) and our definition of gng_{n} (5.4).
Let αk=xk+iyk\alpha_{k}=x_{k}+iy_{k} for all k=1,,k0k=1,\dots,k_{0}. Define

h(𝕪)=mf(𝜶)=k=1k02mli2(e2iαk)2m𝕂(𝜶)𝜶\displaystyle h(\mathbb{y})=\mathfrak{I}m\;f(\bm{\alpha})=\sum_{k=1}^{k_{0}}2\;\mathfrak{I}m\;\text{li}_{2}(-e^{-2i\alpha_{k}})-2\;\mathfrak{I}m\mathbb{K}(\bm{\alpha})\cdot\bm{\alpha}

as a function of the variables y1,,yky_{1},\dots,y_{k}.

Lemma 6.2.

If 0<xk<π20<x_{k}<\frac{\pi}{2} or π<xk<3π2\pi<x_{k}<\frac{3\pi}{2}, then the function h(𝕪)h(\mathbb{y}) is strictly convex in the variables yky_{k}. If π2<xk<π\frac{\pi}{2}<x_{k}<\pi or π2<xk<0-\frac{\pi}{2}<x_{k}<0, then the function is strictly concave in yky_{k}.

Proof.
h(𝕪)yk=4mlog(1+e2ixk+2yk)2(ϵk+ϵk+1+2)xk+4ϵk+1xk+4ϵkxk1\displaystyle\frac{\partial h(\mathbb{y})}{\partial y_{k}}=-4\;\mathfrak{I}m\;\log\left(1+e^{-2ix_{k}+2y_{k}}\right)-2(\epsilon_{k}+\epsilon_{k+1}+2)x_{k}+4\epsilon_{k+1}x_{k}+4\epsilon_{k}x_{k-1}

We get that

2h(𝕪)yk2=4sin(2xk)cos(2xk)+cosh(2yk)\displaystyle\frac{\partial^{2}h(\mathbb{y})}{\partial y_{k}^{2}}=\frac{4\sin(2x_{k})}{\cos(2x_{k})+\cosh(2y_{k})}

The denominator is positive as cosh(2yk)1,cos(2xk)1,xkπ2, and cos(2xk)1\cosh(2y_{k})\geq 1,\cos(2x_{k})\geq-1,x_{k}\neq\frac{\pi}{2},\text{ and }\cos(2x_{k})\neq-1. The second derivative is strictly positive when 0<xk<π20<x_{k}<\frac{\pi}{2} or π<xk<3π2\pi<x_{k}<\frac{3\pi}{2}.
The second derivative is strictly negative when π2<xk<0-\frac{\pi}{2}<x_{k}<0 or π2<xk<π\frac{\pi}{2}<x_{k}<\pi. ∎

Corollary 6.3.

The function h(𝕩)=mf(𝛂)h(\mathbb{x})=\mathfrak{I}m\;f(\bm{\alpha}) as a function of 𝕩\mathbb{x} is strictly convex in the region π2<xk<0-\frac{\pi}{2}<x_{k}<0 and π2<xk<π\frac{\pi}{2}<x_{k}<\pi. h(𝕩)h^{\prime}(\mathbb{x}) is strictly concave in the region 0<xk<π20<x_{k}<\frac{\pi}{2} and π<xk<3π2\pi<x_{k}<\frac{3\pi}{2}.

Proof.

This follows from the holomorphicity of the function f(𝜶)f(\bm{\alpha}). ∎

We now see that the function FnF_{n} has an exponential growth rate smaller than the volume of the mapping torus in certain regions (next Proposition). This will make our estimates easier. Note that here we use the assumption that the volume is greater than 2(k01)v32(k_{0}-1)v_{3} where v3v_{3} is the volume of an ideal hyperbolic regular tetrahedron.

Proposition 6.4.

For ϵ>0\epsilon>0, we can choose a sufficiently small δ>0\delta>0, such that if any αk\alpha_{k} is not in (δ,π2δ)(π+δ,3π2δ)\left(\delta,\frac{\pi}{2}-\delta\right)\cup\left(\pi+\delta,\frac{3\pi}{2}-\delta\right), then

|Fn(𝜶)|<O(en4π(2(k01)v3+ϵ)),\displaystyle\left|F_{n}(\bm{\alpha})\right|<O\left(e^{\frac{n}{4\pi}\left(2(k_{0}-1)v_{3}+\epsilon\right)}\right),

where v3v_{3} is the volume of an ideal regular hyperbolic tetrahedron.

Proof.

Suppose αk\alpha_{k} for k=1,,k0k=1,\dots,k_{0} is not in the desired domain. We split the proof into three cases.

Case I: αk\alpha_{k} is near the bridge points. When αk\alpha_{k} is δ\delta-close to {π2,π2,3π2}\{-\frac{\pi}{2},\frac{\pi}{2},\frac{3\pi}{2}\}, then, by lemma 6.1, we have the desired result.

Case II: When αk{δ,π+δ}\alpha_{k}\in\{\delta,\pi+\delta\} for any k{1,,k0}k\in\{1,\dots,k_{0}\}, then we have

|QDLq(uk,vk|αk)QDLq(u^k,v^k|αk)|=|g~n(αk)|exp(n2πmli2(e2iδ))\displaystyle|QDL^{q}(u_{k},v_{k}|\alpha_{k})QDL^{q}(\hat{u}_{k},\hat{v}_{k}|\alpha_{k})|=|\tilde{g}_{n}(\alpha_{k})|\exp\left(\frac{n}{2\pi}\mathfrak{I}m\;li_{2}(-e^{-2i\delta})\right)
=exp(n2πO1(δlogδ)+O2(1))\displaystyle=\exp\left(\frac{n}{2\pi}O_{1}(\delta\log\delta)+O_{2}(1)\right)

where O2(1)O_{2}(1) comes from the fact that |g~n(αk)||\tilde{g}_{n}(\alpha_{k})| is bounded by some constant and O(δlogδ)O(\delta\log\delta) comes from the fact that mli2(e2i(π/2δ))=2Λ(π/2δ)=O(δlogδ)\mathfrak{I}m\;\operatorname{li_{2}}(e^{2i(\pi/2-\delta)})=2\Lambda(\pi/2-\delta)=O(\delta\log\delta) .

Case III: When αk(π2,0)(π2,π)\alpha_{k}\in(-\frac{\pi}{2},0)\cup(\frac{\pi}{2},\pi), we refer to the lemma 6.2. In the interval, the function mf(𝜶)\mathfrak{I}m\;f(\bm{\alpha}) is convex in the variables xkx_{k}. This implies the maximum is attained at the boundary. However, at the boundary, we have xk{π2+δ,δ,π2+δ,πδ}x_{k}\in\left\{-\frac{\pi}{2}+\delta,-\delta,\frac{\pi}{2}+\delta,\pi-\delta\right\}, and by Lemma 6.1, we see that the value is small. ∎

We add a bump function to define the invariant on a larger space with compact support, in order to apply the Poisson Summation Formula and change from sum to integrals. We look at the behavior of the function FnF_{n} in the region (π/2,3π/2)k0(-\pi/2,3\pi/2)^{k_{0}}. Let J=(j1,,jk0)J=(j_{1},\dots,j_{k_{0}}) be a multi-index where each jkj_{k} can be 0 or 1. For δ>0\delta>0, let

Dδ,J={𝕩k0|jkπ+δ<xk<jkπ+π2δ,k=1,,k0}.\displaystyle D_{\delta,J}^{{\mathbb{R}}}=\left\{\mathbb{x}\in{\mathbb{R}}^{k_{0}}\bigg{|}j_{k}\pi+\delta<x_{k}<j_{k}\pi+\frac{\pi}{2}-\delta,\forall k=1,\dots,k_{0}\right\}.

Consider 𝒟δ\mathcal{D}^{{\mathbb{R}}}_{\delta} as union of all Dδ,JD_{\delta,J}^{{\mathbb{R}}}. When δ=0\delta=0, we call it 𝒟\mathcal{D}^{{\mathbb{R}}}. Similarly, we define

Dδ,J={𝕫k0|jkπ+δ<e(zk)<jkπ+π2δ,k=1,,k0}.\displaystyle D_{\delta,J}^{\mathbb{C}}=\left\{\mathbb{z}\in\mathbb{C}^{k_{0}}\bigg{|}j_{k}\pi+\delta<\mathfrak{R}e\;(z_{k})<j_{k}\pi+\frac{\pi}{2}-\delta,\forall k=1,\dots,k_{0}\right\}.

Now, consider a smooth bump function

ψ:k0\displaystyle\psi\colon{\mathbb{R}}^{k_{0}}\rightarrow{\mathbb{R}}

defined by

{ψ(𝕩)=1 if 𝕩𝒟δ/2¯0<ψ(𝕩)<1 if 𝕩𝒟\𝒟δ/2¯ψ(𝕩)=0 if 𝕩𝒟.\displaystyle\begin{cases}\psi(\mathbb{x})=1&\text{ if }\mathbb{x}\in\overline{\mathcal{D}^{{\mathbb{R}}}_{\delta/2}}\\ 0<\psi(\mathbb{x})<1&\text{ if }\mathbb{x}\in\mathcal{D}^{{\mathbb{R}}}\backslash\overline{\mathcal{D}_{\delta/2}^{{\mathbb{R}}}}\\ \psi(\mathbb{x})=0&\text{ if }\mathbb{x}\notin\mathcal{D}^{{\mathbb{R}}}.\end{cases}

We define a function

Gn(𝜶)=ψ(𝜶)Fn(𝜶).\displaystyle G_{n}(\bm{\alpha})=\psi(\bm{\alpha})F_{n}(\bm{\alpha}).

We can then write our invariant as

TraceΛφ,rq=An(i1,,ik0)k0Gn(2πi1n,,2πik0n)+O(en4π(2(k01)v3+ϵ).\displaystyle\operatorname{Trace}\;\Lambda_{\varphi,r}^{q}=A_{n}\sum_{(i_{1},\dots,i_{k_{0}})\in{\mathbb{Z}}^{k_{0}}}G_{n}\left(\frac{2\pi i_{1}}{n},\dots,\frac{2\pi i_{k_{0}}}{n}\right)+O\left(e^{\frac{n}{4\pi}(2(k_{0}-1)v_{3}+\epsilon}\right).

Because the function GnG_{n} is smooth and with compact support, we have the following classical formula, see [O16]

Proposition 6.5.

(Poisson Summation Formula) For the function Gn:k0G_{n}\colon{\mathbb{R}}^{k_{0}}\rightarrow\mathbb{C} defined above,

(i1,,ik0)k0Gn(2πi1n,,2πik0n)=𝕞k0G^n(𝕞)\displaystyle\sum_{(i_{1},\dots,i_{k_{0}})\in{\mathbb{Z}}^{k_{0}}}G_{n}\left(\frac{2\pi i_{1}}{n},\dots,\frac{2\pi i_{k_{0}}}{n}\right)=\sum_{\mathbb{m}\in{\mathbb{Z}}^{k_{0}}}\hat{G}_{n}(\mathbb{m})

for the Fourier coefficients

G^n(𝕞)=(n2π)k0𝒟Gn(𝜶)e1n𝕞𝜶𝑑𝜶.\displaystyle\hat{G}_{n}(\mathbb{m})=\left(\frac{n}{2\pi}\right)^{k_{0}}\int_{\mathcal{D}^{{\mathbb{R}}}}G_{n}(\bm{\alpha})e^{\sqrt{-1}n\mathbb{m}\cdot\bm{\alpha}}d\bm{\alpha}.

7. Geometry of critical points

In this section, we will see how the critical points relate to the geometry of the mapping torus MφM_{\varphi}. The main result of this section is Proposition 7.5.

7.1. The function ff

Recall that the function f(𝜶)f(\bm{\alpha}) in equation (5.5) is defined by

f(𝜶)=k=1k02li2(e2iαk)2𝕂(𝜶)𝜶.\displaystyle f(\bm{\alpha})=\sum_{k=1}^{k_{0}}2\text{li}_{2}(-e^{-2i\alpha_{k}})-2\mathbb{K}(\bm{\alpha})\cdot\bm{\alpha}.

Then, the derivative is given by

(7.1) f(𝜶)αk=4ilog(1+e2iαk)2(ϵk+ϵk+1+2)αk+2ϵkαk1+2ϵk+1αk+1.\displaystyle\frac{\partial f(\bm{\alpha})}{\partial\alpha_{k}}=4i\log(1+e^{-2i\alpha_{k}})-2(\epsilon_{k}+\epsilon_{k+1}+2)\alpha_{k}+2\epsilon_{k}\alpha_{k-1}+2\epsilon_{k+1}\alpha_{k+1}.

The second derivatives are given by

2f(𝜶)αk2\displaystyle\frac{\partial^{2}f(\bm{\alpha})}{\partial\alpha_{k}^{2}} =8e2iαk1+e2iαk2(ϵk+ϵk+1+2),\displaystyle=\frac{8e^{-2i\alpha_{k}}}{1+e^{-2i\alpha_{k}}}-2(\epsilon_{k}+\epsilon_{k+1}+2),
2f(𝜶)αkαj\displaystyle\frac{\partial^{2}f(\bm{\alpha})}{\partial\alpha_{k}\partial\alpha_{j}} ={2ϵk if j=k12ϵk+1 if j=k+1.\displaystyle=\begin{cases}2\epsilon_{k}&\text{ if }j=k-1\\ 2\epsilon_{k+1}&\text{ if }j=k+1.\end{cases}

From lemma 6.2, we know that for αk=xk+iyk\alpha_{k}=x_{k}+iy_{k},

mf(𝜶)yk=4mlog(1+e2ixk+2yk)(ϵk+ϵk+1+2)xk2ϵk+1xk+12ϵkxk1.\displaystyle\frac{\partial\;\mathfrak{I}m\;f(\bm{\alpha})}{\partial y_{k}}=-4\;\mathfrak{I}m\;\log\left(1+e^{-2ix_{k}+2y_{k}}\right)-(\epsilon_{k}+\epsilon_{k+1}+2)x_{k}-2\epsilon_{k+1}x_{k+1}-2\epsilon_{k}x_{k-1}.

We get that

(7.2) 2mf(𝜶)yk2=4sin(2xk)cos(2xk)+cosh(2yk).\displaystyle\frac{\partial^{2}\;\mathfrak{I}m\;f(\bm{\alpha})}{\partial y_{k}^{2}}=\frac{4\sin(2x_{k})}{\cos(2x_{k})+\cosh(2y_{k})}.

We call the unique critical point in the region {(α1,,αk0)|(x1,,xk0)(0,π2)k0}\{(\alpha_{1},\dots,\alpha_{k_{0}})\big{|}(x_{1},\dots,x_{k_{0}})\in\left(0,\frac{\pi}{2}\right)^{k_{0}}\} as

(7.3) 𝜶𝟘=(α10,,αk00).\displaystyle\bm{\alpha}^{\mathbb{0}}=(\alpha_{1}^{0},\dots,\alpha_{k_{0}}^{0}).

This critical point exists because of the existence of geometric triangulation of the four-punctured sphere bundle [GF, Appendix A].

7.2. Edge weight system and shape parameters

Recall that a pseudo-Anosov diffeomorphism φ\varphi gives rise to an ideal triangulation sweep for φ\varphi, such that

τk=φ1φ2φk(τ0).\displaystyle\tau_{k}=\varphi_{1}\circ\varphi_{2}\circ\dots\circ\varphi_{k}(\tau_{0}).

A periodic edge weight system for this ideal triangulation sweep is a family (a0,,f0),,(a_{0},\dots,f_{0}),\dots, (ak0,,fk0)=(a0,,f0)(a_{k_{0}},\dots,f_{k_{0}})=(a_{0},\dots,f_{0}) where (ak,bk)(a_{k},b_{k}) is related to (ak1,bk1)(a_{k-1},b_{k-1}) by (3.1) and the other edges are related by the central elements pjp_{j} for j=1,2,3,4j=1,2,3,4. Such a periodic edge weight system determines a character [r~]𝒳PSL2()(S0,4)[\tilde{r}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}) that is fixed under the action of φ\varphi.

Proposition 7.1.

Let 𝛂c\bm{\alpha}^{c} be a smooth critical point of the function f(𝛂)f(\bm{\alpha}). Then, we have a unique periodic edge weight system (a0,,f0),(a1,,f1),,(ak0,,fk0)=(a0,,f0)(a_{0},\dots,f_{0}),(a_{1},\dots,f_{1}),\dots,(a_{k_{0}},\dots,f_{k_{0}})=(a_{0},\dots,f_{0}), for the ideal triangulation sweep τ0,τ1,,τk0=φ(τ0)\tau_{0},\tau_{1},\dots,\tau_{k_{0}}=\varphi(\tau_{0}) such that

ak=e2iαk\displaystyle a_{k}=e^{2i\alpha_{k}}

for every k=1,2,,k0k=1,2,\dots,k_{0}. We also get that pjn=1p_{j}^{n}=1 for j = 1,2,3,4.

Proof.

By 7.1, the critical point equation of ff is

4ilog(1+e2iαk)=4ϵkαk1+2(ϵk+ϵk+1+2)αk4ϵk+1αk+1\displaystyle 4i\log(1+e^{-2i\alpha_{k}})=-4\epsilon_{k}\alpha_{k-1}+2(\epsilon_{k}+\epsilon_{k+1}+2)\alpha_{k}-4\epsilon_{k+1}\alpha_{k+1}

which implies that

(1+e2iαk)2=(e2iαk1)ϵk(e2iαk)ϵk+ϵk+121(e2iαk+1)ϵk+1\displaystyle(1+e^{-2i\alpha_{k}})^{2}=(e^{2i\alpha_{k-1}})^{\epsilon_{k}}(e^{2i\alpha_{k}})^{-\frac{\epsilon_{k}+\epsilon_{k+1}}{2}-1}(e^{2i\alpha_{k+1}})^{\epsilon_{k+1}}

for k=1,,k0k=1,\dots,k_{0} and where the indices are counted modulo k0k_{0}.
We define ak=e2iαka_{k}=e^{2i\alpha_{k}} using induction. Assume that the first diagonal exchange in the diffeomorphism is \mathcal{L}. The other case for \mathcal{R} is similar. We start with

a0=d0=e2iα0\displaystyle a_{0}=d_{0}=e^{2i\alpha_{0}} b0=e0=e2iα1\displaystyle b_{0}=e_{0}=e^{-2i\alpha_{1}} c0=f0=e2i(α1α0).\displaystyle c_{0}=f_{0}=e^{2i(\alpha_{1}-\alpha_{0})}.

Then, a1=b01=e2iα1a_{1}=b_{0}^{-1}=e^{2i\alpha_{1}}. We also have pjn=1p_{j}^{n}=1 for all j=1,2,3,4j=1,2,3,4. This shows that (ak,,fk)(a_{k},\dots,f_{k}) is uniquely determined by αk\alpha_{k} for k=1,,k0k=1,\dots,k_{0} as the consecutive edge weights are related by (3.1).
For k2k\geq 2, assume that ak=e2iαka_{k}=e^{2i\alpha_{k}} and ak1=e2iαk1a_{k-1}=e^{2i\alpha_{k-1}}. We’ll prove that ak+1=e2iαk+1a_{k+1}=e^{2i\alpha_{k+1}}.
There are four cases. (φk1,φk)={(L,L),(L,R),(R,L),(R,R)}(\varphi_{k-1},\varphi_{k})=\{(L,L),(L,R),(R,L),(R,R)\}.
CASE (LL). ϵk=1=ϵk+1\epsilon_{k}=-1=\epsilon_{k+1}
From the edge weight system, we get

ak+1=bk1=(1+ak1)2ak11=(1+e2iαk)2e2iαk1=e2iαk+1.\displaystyle a_{k+1}=b_{k}^{-1}=(1+a_{k}^{-1})^{-2}a_{k-1}^{-1}=(1+e^{-2i\alpha_{k}})^{-2}e^{-2i\alpha_{k-1}}=e^{2i\alpha_{k+1}}.

where the last equality comes from the critical point equation

(1+e2iαk)2=e2iαk12iαk+1.\displaystyle(1+e^{-2i\alpha_{k}})^{2}=e^{-2i\alpha_{k-1}-2i\alpha_{k+1}}.

CASE (LR). ϵk=1,ϵk+1=1\epsilon_{k}=-1,\epsilon_{k+1}=1
From the edge weight system, we get

ak+1=akbk=ak(1+ak1)2ak1=e2iαk+2iαk1(1+e2iαk)2=e2iαk+1.\displaystyle a_{k+1}=a_{k}b_{k}=a_{k}(1+a_{k}^{-1})^{2}a_{k-1}=e^{2i\alpha_{k}+2i\alpha_{k-1}}(1+e^{-2i\alpha_{k}})^{2}=e^{2i\alpha_{k+1}}.

where the last equality comes from the critical point equation

(1+e2iαk)2=e2iαk12iαk+2iαk+1.\displaystyle(1+e^{-2i\alpha_{k}})^{2}=e^{-2i\alpha_{k-1}-2i\alpha_{k}+2i\alpha_{k+1}}.

CASE (RL). ϵk=1,ϵk+1=1\epsilon_{k}=1,\epsilon_{k+1}=-1

ak+1=bk1=(1+e2iαk)2e2iαk+2iαk1=e2iαk+1\displaystyle a_{k+1}=b_{k}^{-1}=(1+e^{-2i\alpha_{k}})^{-2}e^{-2i\alpha_{k}+2i\alpha_{k-1}}=e^{2i\alpha_{k+1}}

CASE (RR). ϵk=1=ϵk+1\epsilon_{k}=1=\epsilon_{k+1}

ak+1=akbk=(1+e2iαk)2e4iαk2iαk1=e2iαk+1\displaystyle a_{k+1}=a_{k}b_{k}=(1+e^{-2i\alpha_{k}})^{2}e^{4i\alpha_{k}-2i\alpha_{k-1}}=e^{2i\alpha_{k+1}}

For the PSL2()\mathrm{PSL}_{2}({\mathbb{Z}}) part of the diffeomorphism, we see that we have the result. If the 2×2{\mathbb{Z}}_{2}\times{\mathbb{Z}}_{2} part of φ\varphi is non-trivial, then that corresponds to flipping edges. Since everything until this point can be expressed in terms of {αk}k=1,,k0\{\alpha_{k}\}_{k=1,\dots,k_{0}}, the permutation of the edges will still be expressed in terms of {αk}k=1,,k0\{\alpha_{k}\}_{k=1,\dots,k_{0}}.
Here, we abuse the notation a little. For k=1,,k0k=1,\dots,k_{0}, we get our edge weights. The diffeomorphism φ\varphi involves (up to) two permutation elements as well, which add to the edge weight system. However, since the opposite edges are the same (in terms of αk\alpha_{k}), the edge weight system does not change in terms of αk\alpha_{k}. We ignore these two elements when we talk about the periodicity of the edge weight system.

It’s important to note that one needs to consider the permutation elements in order for the character [r~hyp][\tilde{r}_{\text{hyp}}] to be φ\varphi-invariant, even though its action on the edge weight system is trivial.

This concludes that every critical point 𝜶𝒄\bm{\alpha^{c}} provides an edge weight system (a0,,f0),,(a_{0},\dots,f_{0}),\dots, (ak0,,fk0)(a_{k_{0}},\dots,f_{k_{0}}) such that ak=e2iαka_{k}=e^{2i\alpha_{k}} for every k. The index k in αk\alpha_{k} is counted modulo k0k_{0}, which means we have (ak0,bk0)=(e2iαk0,e2iαk0+1)=(e2iα0,e2iα1)=(a0,b0)(a_{k_{0}},b_{k_{0}})=(e^{2i\alpha_{k_{0}}},e^{-2i\alpha_{k_{0}+1}})=(e^{2i\alpha_{0}},e^{-2i\alpha_{1}})=(a_{0},b_{0}), where the second equality comes from the fact that αk0\alpha_{k_{0}} and α0\alpha_{0} are the same as the indices for α\alpha are counted modulo k0k_{0}. Since, at each step, we only (possibly) permute pjp_{j}, we see that pj(k)=1p_{j}^{(k)}=1 for j=1,2,3,4j=1,2,3,4 and k=1,,k0k=1,\dots,k_{0}. Thus, the edge weight system is periodic and unique.

We get a unique periodic edge weight system, that corresponds to a hyperbolic character [r~hyp]𝒳PSL2()(S0,4)[\tilde{r}_{\text{hyp}}]\in\mathcal{X}_{\mathrm{PSL}_{2}(\mathbb{C})}(S_{0,4}) that is φ\varphi-invariant. This particular character (and edge weight system) will give us the hyperbolic volume of the mapping torus at the complete structure.

7.3. Critial values and volume

The following lemma is due to Ka Ho Wong [W].

Lemma 7.2.

Consider a function

f(𝜶)=k=1k02li2(e2αk)+Q(𝜶)\displaystyle f(\bm{\alpha})=\sum_{k=1}^{k_{0}}2li_{2}(-e^{-2\alpha_{k}})+Q(\bm{\alpha})

where Q(𝛂)Q(\bm{\alpha}) is a quadratic polynomial with real coefficients in 𝛂\bm{\alpha}. If 𝛂c\bm{\alpha}^{c} is a critical point of ff, then

mf(𝜶c)=k=1k02D(e2iαk)\displaystyle\mathfrak{I}m\;f(\bm{\alpha}^{c})=\sum_{k=1}^{k_{0}}2D(-e^{-2i\alpha_{k}})

where D(z)=mli2(z)+arg(1z)ln|z|D(z)=\mathfrak{I}m\;li_{2}(z)+\arg(1-z)\ln|z| is the Bloch-Wigner dilogarithm function.

Proof.

Let αk=xk+yk\alpha_{k}=x_{k}+y_{k}. Then

mli2(e2iαk)yk=mli2(e2iαk)yk=2arg(1+e2iαk).\displaystyle\frac{\partial\;\mathfrak{I}m\;\text{li}_{2}(-e^{-2i\alpha_{k}})}{\partial y_{k}}=\mathfrak{I}m\;\frac{\partial\;\text{li}_{2}(-e^{-2i\alpha_{k}})}{\partial y_{k}}=-2\arg(1+e^{-2i\alpha_{k}}).

The imaginary part of the quadratic term is linear in yky_{k} and has no constant term because the coefficients are real. Therefore,

mQ(𝜶)=k=1k0ykQ(𝜶)yk.\displaystyle\mathfrak{I}m\;Q(\bm{\alpha})=\sum_{k=1}^{k_{0}}y_{k}\frac{\partial\;Q(\bm{\alpha})}{\partial y_{k}}\;.

At the critical point (because f is holomorphic),

mQ(𝜶c)yk=m 2li2(e2iαkc)yk\displaystyle\frac{\partial\;\mathfrak{I}m\;Q(\bm{\alpha}^{c})}{\partial y_{k}}=-\frac{\partial\;\mathfrak{I}m\;2\;\text{li}_{2}(-e^{-2i\alpha_{k}^{c}})}{\partial y_{k}}

for every k. Therefore, at the critical point 𝜶c\bm{\alpha}^{c}

mQ(𝜶c)=k=1k0yk 2mli2(e2iαkc)yk=k=1k04ykarg(1+e2iαkc)\displaystyle\mathfrak{I}m\;Q(\bm{\alpha}^{c})=-\sum_{k=1}^{k_{0}}y_{k}\frac{\partial\;2\;\mathfrak{I}m\;\text{li}_{2}(-e^{-2i\alpha_{k}^{c}})}{\partial y_{k}}=\sum_{k=1}^{k_{0}}4y_{k}\arg(1+e^{-2i\alpha_{k}^{c}})
=k=1k02D(e2iαkc)k=1k02mli2(e2iαkc).\displaystyle=\sum_{k=1}^{k_{0}}2D(-e^{-2i\alpha_{k}^{c}})-\sum_{k=1}^{k_{0}}2\;\mathfrak{I}m\;\text{li}_{2}(-e^{-2i\alpha_{k}^{c}}).

By definition of the volume of a character, we get the following lemma.

Lemma 7.3.

Let r:π1(Mφ)PSL2()r\colon\pi_{1}(M_{\varphi})\rightarrow\text{PSL}_{2}(\mathbb{C}) be associated to a periodic edge weight system. Then the volume of rr is equal to

vol(r)=2k=1k0D(ak1).\displaystyle vol(r)=2\sum_{k=1}^{k_{0}}D(-a_{k}^{-1}).
Remark 7.4.

Note that we get a factor of 2 in front of the Bloch-Wigner dilogarithm because the exchange involves two tetrahedra.

Proposition 7.5.

Let (𝛂c)(\bm{\alpha}^{c}) be a critical point of the function ff and r:π1(Mφ)PSL2()r\colon\pi_{1}(M_{\varphi})\rightarrow\text{PSL}_{2}(\mathbb{C}) be its associated homomorphism. Then,

mf(𝜶c)=vol(r).\displaystyle\mathfrak{I}m\;f(\bm{\alpha}^{c})=vol(r).
Proof.

This follows from Lemma 7.2 and Lemma 7.3. ∎

8. Saddle point method

We wish to apply the saddle point method, for which we need to deform the integral domain appropriately. Here we make a strong technical assumption that vol(Mφ)>2(k01)v3\operatorname{vol}(M_{\varphi})>2(k_{0}-1)v_{3} where v3v_{3} is the volume of an ideal regular hyperbolic tetrahedron. We expect that this assumption can be lifted with some other techniques.

Proposition 8.1.

[WY20] Let f(z1,,zk0)f(z_{1},\dots,z_{k_{0}}) and g(z1,,zk0)g(z_{1},\dots,z_{k_{0}}) be two holomorphic functions defined on a region Dk0D\subset\mathbb{C}^{k_{0}}. Let SS be an embedded real nn-dimensional closed disk in DD and let (c1,,ck0)S(c_{1},\dots,c_{k_{0}})\in S be a point such that

  1. (1)

    (c1,,ck0)(c_{1},\dots,c_{k_{0}}) is a critical point of ff in DD,

  2. (2)

    e(f)(z1,,zk0)<e(f)(c1,,ck0)\mathfrak{R}e\;(f)(z_{1},\dots,z_{k_{0}})<\mathfrak{R}e\;(f)(c_{1},\dots,c_{k_{0}}) for all (z1,,zk0)S\{(c1,,ck0)}(z_{1},\dots,z_{k_{0}})\in S\backslash\{(c_{1},\dots,c_{k_{0}})\},

  3. (3)

    the domain {z1,,zk0)D|ef(z1,,zk0)<ef(c1,,ck0)}\{z_{1},\dots,z_{k_{0}})\in D|\mathfrak{R}ef(z_{1},\dots,z_{k_{0}})<\mathfrak{R}ef(c_{1},\dots,c_{k_{0}})\} deformation retracts to S\{(c1,,ck0)}S\backslash\{(c_{1},\dots,c_{k_{0}})\},

  4. (4)

    the Hessian matrix Hess(f)(c1,,ck0)\operatorname{Hess}(f)(c_{1},\dots,c_{k_{0}}) is non singular,

  5. (5)

    g(c1,,ck0)g(c_{1},\dots,c_{k_{0}}) is non zero.

Then, for any sequence of holomorphic functions

fn(z1,,zk0)=f(z1,,zk0)+vn(z1,,zk0)n2\displaystyle f_{n}(z_{1},\dots,z_{k_{0}})=f(z_{1},\dots,z_{k_{0}})+\frac{v_{n}(z_{1},\dots,z_{k_{0}})}{n^{2}}

where |vn(z1,,zk0)||v_{n}(z_{1},\dots,z_{k_{0}})| is bounded from above by a constant independent of n in D.
Then,

Sg(z1,,zk0)\displaystyle\int_{S}g(z_{1},\dots,z_{k_{0}}) enfn(z1,,zk0)dz1dzk0\displaystyle e^{nf_{n}(z_{1},\dots,z_{k_{0}})}dz_{1}\dots dz_{k_{0}}
=\displaystyle= (2πn)k02g(c1,,ck0)(1)k0detHess(f)(c1,,ck0)enf(c1,,ck0)(1+O(1n)).\displaystyle\left(\frac{2\pi}{n}\right)^{\frac{k_{0}}{2}}\frac{g(c_{1},\dots,c_{k_{0}})}{\sqrt{(-1)^{k_{0}}\det\operatorname{Hess}(f)(c_{1},\dots,c_{k_{0}})}}e^{nf(c_{1},\dots,c_{k_{0}})}\left(1+O\left(\frac{1}{n}\right)\right).

Remark 8.2.

There is a typo in [WY20]. The exponent of (1)(-1) is missing there.

We will estimate the following:

𝒟ψ(𝜶)gn(𝜶)en4πif(𝜶)+O(1n)e1n𝕞𝜶𝑑𝜶.\displaystyle\int_{\mathcal{D}^{{\mathbb{R}}}}\psi(\bm{\alpha})g_{n}(\bm{\alpha})e^{\frac{n}{4\pi i}f(\bm{\alpha})+O\left(\frac{1}{n}\right)}e^{\sqrt{-1}n\mathbb{m}\cdot\bm{\alpha}}d\bm{\alpha}.

The coefficient of n in the exponent is f(𝜶)4πi+O(1n2)\frac{f(\bm{\alpha})}{4\pi i}+O\left(\frac{1}{n^{2}}\right). Therefore, we have the sequence of holomorphic functions we want.

Let αj=xj+iyj\alpha_{j}=x_{j}+iy_{j} for all j{1,,k0}j\in\{1,\dots,k_{0}\}. For δ>0\delta>0 small enough, choose the region 𝒟δ\mathcal{D}^{\mathbb{C}}_{\delta}, and let {𝜶J𝒟δ,J|J{0,1}k0}\{\bm{\alpha}^{J}\in\mathcal{D}^{\mathbb{C}}_{\delta,J}|J\in\{0,1\}^{k_{0}}\} be the set of critical points where for each fixed (j1,,jk0)(j_{1},\dots,j_{k_{0}}). The critical point is unique in their respective regions. We can assume that there is a critical point in the interval (0,π)k0(0,\pi)^{k_{0}}. For each of these intervals, if any of xj>π2x_{j}>\frac{\pi}{2}, then we get a concave up function in xjx_{j}, and we will get a value lesser than the value at the boundary, which is lesser than vol(Mφ)\operatorname{vol}({M_{\varphi}}). Therefore, the critical point will be in the interval (0,π2)k0(0,π)k0\left(0,\tfrac{\pi}{2}\right)^{k_{0}}\subset\left(0,\pi\right)^{k_{0}}.

Let 𝜶𝟘=((x10+iy10),,(xk00+iyk00))\bm{\alpha}^{\mathbb{0}}=((x_{1}^{0}+iy_{1}^{0}),\dots,(x_{k_{0}}^{0}+iy_{k_{0}}^{0})) be the unique critical point where 𝟘=(0,,0)\mathbb{0}=(0,\dots,0). Let S𝟘=StopSsideS^{\mathbb{0}}=S^{\text{top}}\cup S^{\text{side}} where

Stop\displaystyle S^{\text{top}} ={(x1+iy10,,xk0+iyk00)|(x1,,xk0)(δ,π2δ)k0}\displaystyle=\left\{(x_{1}+iy_{1}^{0},\dots,x_{k_{0}}+iy_{k_{0}}^{0})\big{|}(x_{1},\dots,x_{k_{0}})\in\left(\delta,\frac{\pi}{2}-\delta\right)^{k_{0}}\right\}
Sside\displaystyle S^{\text{side}} ={(x1+ity10,,xk0+ityk00)|t[0,1],(x1,,xk0)𝒟δ,𝟘}.\displaystyle=\left\{(x_{1}+ity_{1}^{0},\dots,x_{k_{0}}+ity_{k_{0}}^{0})\big{|}t\in[0,1],(x_{1},\dots,x_{k_{0}})\in\partial\mathcal{D}^{{\mathbb{R}}}_{\delta,\mathbb{0}}\right\}.

In figure 5, we can see the cube when there are only two variables x1x_{1} and x2x_{2}.

Refer to caption
Figure 5. The domain S0S^{0} simplified for the case where k0=2k_{0}=2. In general, it’s a multi-dimensional cube structure. StopS^{\text{top}} contains the critical point. Here, SsideS^{\text{side}} consists of all the four sides.

Similarly, we define SJ=S~J,topS~J,sideS^{J}=\tilde{S}^{J,\text{top}}\cup\tilde{S}^{J,\text{side}} for J𝟘J\neq\mathbb{0}, where

S~J,top\displaystyle\tilde{S}^{J,\text{top}} ={(x1+iy10,,xk0+iyk00)|(x1,,xk0)Dδ,J}\displaystyle=\left\{(x_{1}+iy_{1}^{0},\dots,x_{k_{0}}+iy_{k_{0}}^{0})\big{|}(x_{1},\dots,x_{k_{0}})\in D^{{\mathbb{R}}}_{\delta,J}\right\}
S~J,side\displaystyle\tilde{S}^{J,\text{side}} ={(x1+ity10,,xk0+ityk00)|t[0,1],(x1,,xk0,m1,,mk0)𝒟δ,J}.\displaystyle=\left\{(x_{1}+ity_{1}^{0},\dots,x_{k_{0}}+ity_{k_{0}}^{0})\big{|}t\in[0,1],(x_{1},\dots,x_{k_{0}},m_{1},\dots,m_{k_{0}})\in\partial\mathcal{D}^{{\mathbb{R}}}_{\delta,J}\right\}.

8.1. Estimate of the leading Fourier coefficient on Dδ,0D_{\delta,\textbf{0}}^{\mathbb{C}}

We show that the (0,,0)(0,\dots,0)-th Fourier coefficient has an exponential growth rate equal to the volume of the manifold on the region Dδ,𝟘D_{\delta,\mathbb{0}}^{\mathbb{C}}. We do this by deforming the integral domain Dδ,𝟘D_{\delta,\mathbb{0}}^{{\mathbb{R}}} to S𝟘S^{\mathbb{0}} and estimating mf(𝜶)\mathfrak{I}m\;f(\bm{\alpha}).

Proposition 8.3.

mf(𝜶)\mathfrak{I}m\;f(\bm{\alpha}) achieves the only absolute maximum at 𝛂𝟘\bm{\alpha}^{\mathbb{0}} on S𝟘S^{\mathbb{0}}.

Proof.

On SsideS^{\text{side}}, for any fixed 𝕩𝒟δ,𝟘\mathbb{x}\in\mathcal{D}^{{\mathbb{R}}}_{\delta,\mathbb{0}}, h𝕩(t)=mf(𝜶t)h_{\mathbb{x}}(t)=\mathfrak{I}m\;f(\bm{\alpha}_{t}) is concave up in t by Lemma 6.2. Therefore, it suffices to check the value of h𝕩(t)h_{\mathbb{x}}(t) at the boundary, t=0t=0 and t=1t=1.
At t=0, we have h𝕩(0)=mf(𝕩)h_{\mathbb{x}}(0)=\mathfrak{I}m\;f(\mathbb{x}). Here at least one of the xj=δx_{j}=\delta or π2δ\frac{\pi}{2}-\delta. By arguments from lemma 6.4, we see that mf(𝕩)<δ+2(k01)v3<vol(Mφ)\mathfrak{I}m\;f(\mathbb{x})<\delta+2(k_{0}-1)v_{3}<\operatorname{vol}(M_{\varphi}), where v3v_{3} is the volume of a regular ideal tetrahedron.

When t=1t=1, h𝕩(1)=mh(x1+iy10,,xk0+iyk00)h_{\mathbb{x}}(1)=\mathfrak{I}m\;h_{\infty}(x_{1}+iy_{1}^{0},\dots,x_{k_{0}}+iy_{k_{0}}^{0}). Now, on StopS^{\text{top}}, the function f(𝜶)f(\bm{\alpha}) is concave down in the e(𝜶)\mathfrak{R}e(\bm{\alpha}) by lemma 6.2. In that region, we get a unique maximum and the value at the boundary is strictly less than the value at the top.

mf(𝜶𝟘)=vol(Mφ)>mf(𝜶)\displaystyle\mathfrak{I}m\;f(\bm{\alpha}^{\mathbb{0}})=\operatorname{vol}(M_{\varphi})>\mathfrak{I}m\;f(\bm{\alpha})

where 𝜶𝜶0\bm{\alpha}\neq\bm{\alpha}^{0}. This includes h𝕩(1)h_{\mathbb{x}}(1). For all 𝕩𝒟δ,𝟘\mathbb{x}\in\mathcal{D}^{{\mathbb{R}}}_{\delta,\mathbb{0}}, we have

mf(x1+iy10,,xk0+iyk00)vol(Mφ).\displaystyle\mathfrak{I}m\;f(x_{1}+iy_{1}^{0},\dots,x_{k_{0}}+iy_{k_{0}}^{0})\leq\operatorname{vol}(M_{\varphi}).

Thus, mf(𝜶)\mathfrak{I}mf(\bm{\alpha}) on SsideS^{\text{side}} is smaller than the volume, and it equals the volume at a unique point 𝜶𝟘\bm{\alpha}^{\mathbb{0}} in StopS^{\text{top}}. ∎

Remark 8.4.

Note that we wish to study the function f(𝜶)4π𝕞𝜶f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha} where 𝕞\mathbb{m} is the Fourier coefficient. The above proposition shows us that the (0,,0)(0,\dots,0)-th Fourier coefficient gives us the volume.

We now show that the hessian is non-singular at the critical point.

Lemma 8.5.

The Hessian of the function f(𝛂)f(\bm{\alpha}) is non-singular at 𝛂c\bm{\alpha}^{c}, where 𝛂c\bm{\alpha}^{c} is a critical point of ff.

Proof.

We first show that the Hessian of the imaginary part of f,Hess(mf(𝜶))f,\operatorname{Hess}(\mathfrak{I}m\;f(\bm{\alpha})) is positive definite at the critical point.
Using equation (7.2), we see that the diagonal terms in the Hessian are 4sin(2xk)cos(2xk)+cosh(2yk)\frac{4\sin(2x_{k})}{\cos(2x_{k})+\cosh(2y_{k})}. The off-diagonal terms are 0, that is, 2mf(𝜶)ykyj=0\tfrac{\partial^{2}\mathfrak{I}m\;f(\bm{\alpha})}{\partial y_{k}\partial y_{j}}=0. The critical points lie in the interval xk(0,π2)(π,3π2)x_{k}\in(0,\tfrac{\pi}{2})\cup(\pi,3\tfrac{\pi}{2}). In this interval, all the diagonal terms are positive. Therefore, the Hessian is positive definite at the critical point.
Using [AGN, Lemma 7.2], if Hess(mf(𝜶))\operatorname{Hess}(\mathfrak{I}mf(\bm{\alpha})) is positive definite at the critical point and Hess(ef(𝜶))\operatorname{Hess}(\mathfrak{R}ef(\bm{\alpha})) is symmetric (by definition of hessian), then Hess(f(𝜶))\operatorname{Hess}(f(\bm{\alpha})) is non-singular at the critical point. ∎

Proposition 8.6.

We have the following estimate

|Dδ,𝟘ψ(x1,,xk0)gn(α1,,αk0)en4πif(α1,,αk0)+O(1n)𝑑α1𝑑αk0|\displaystyle\left|\int_{D_{\delta,\mathbb{0}}^{\mathbb{C}}}\psi(x_{1},\dots,x_{k_{0}})g_{n}(\alpha_{1},\dots,\alpha_{k_{0}})e^{\frac{n}{4\pi i}f(\alpha_{1},\dots,\alpha_{k_{0}})+O\left(\frac{1}{n}\right)}d\alpha_{1}\dots d\alpha_{k_{0}}\right|
=c(1)k0detHess(f)(α10,,αk00)en4πvol(Mφ)(1+O(1n))\displaystyle=\frac{c^{\prime}}{\sqrt{(-1)^{k_{0}}\det\operatorname{Hess}(f)(\alpha_{1}^{0},\dots,\alpha_{k_{0}}^{0})}}e^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})}\left(1+O\left(\frac{1}{n}\right)\right)

where c>0c^{\prime}>0 is a constant.

Proof.

By analyticity, the integral remains the same if we deform the domain Dδ,𝟘D_{\delta,\mathbb{0}}^{\mathbb{C}} to S𝟘S^{\mathbb{0}}. We apply proposition 8.1 with region Dδ,𝟘D_{\delta,\mathbb{0}}^{\mathbb{C}}, embedded disk S𝟘S^{\mathbb{0}}, the functions f(α1,,αk0)f(\alpha_{1},\dots,\alpha_{k_{0}}),
ψ(x1,,xk0)gn(α1,,αk0)\psi(x_{1},\dots,x_{k_{0}})g_{n}(\alpha_{1},\dots,\alpha_{k_{0}}) and the critical point (α10,,αk00)(\alpha_{1}^{0},\dots,\alpha_{k_{0}}^{0}) given the conditions are satisfied.

(1) and (2) of Proposition 8.1 are satisfied by Proposition 8.3.

For (3), we need to show that the domain

{(α1,,αk0)Dδ,𝟘|mf(α1,,αk0)<mf(α10,,αk00)}\displaystyle\left\{(\alpha_{1},\dots,\alpha_{k_{0}})\in D_{\delta,\mathbb{0}}^{\mathbb{C}}\;|\;\mathfrak{I}m\;f(\alpha_{1},\dots,\alpha_{k_{0}})<\mathfrak{I}m\;f(\alpha_{1}^{0},\dots,\alpha_{k_{0}}^{0})\right\}

deformation retracts to S\(α10,,αk00)S\backslash(\alpha_{1}^{0},\dots,\alpha_{k_{0}}^{0}). Let (θ1,,θk0)Dδ,0(\theta_{1},\dots,\theta_{k_{0}})\in D^{{\mathbb{R}}}_{\delta,0}. Define

Pθ1,,θk0={(α1,,αk0)Dδ,𝟘|e(αk)=θk for k=1,,k0}\displaystyle P_{\theta_{1},\dots,\theta_{k_{0}}}=\{(\alpha_{1},\dots,\alpha_{k_{0}})\in D_{\delta,\mathbb{0}}^{\mathbb{C}}\;|\;\mathfrak{R}e\;(\alpha_{k})=\theta_{k}\text{ for }k=1,\dots,k_{0}\}

and

Dθ1,,θk0={(α1,,αk0)Pθ1,,θk0|mf(α1,,αk0)<mf(α10,,αk00)}.\displaystyle D_{\theta_{1},\dots,\theta_{k_{0}}}=\{(\alpha_{1},\dots,\alpha_{k_{0}})\in P_{\theta_{1},\dots,\theta_{k_{0}}}\;|\;\mathfrak{I}m\;f(\alpha_{1},\dots,\alpha_{k_{0}})<\mathfrak{I}m\;f(\alpha_{1}^{0},\dots,\alpha_{k_{0}}^{0})\}.

By Lemma 6.2, we see that De(α1),,e(αk0)=D_{\mathfrak{R}e\;(\alpha_{1}),\dots,\mathfrak{R}e\;(\alpha_{k_{0}})}=\emptyset as the absolute minimum is achieved in the variable m(αk)\mathfrak{I}m\;(\alpha_{k}) for k=1,,k0k=1,\dots,k_{0} in the region Dδ,𝟘D_{\delta,\mathbb{0}}^{\mathbb{C}}.
Dθ1,,θk0D_{\theta_{1},\dots,\theta_{k_{0}}} is homeomorphic to a disk for (θ1,,θk0)(e(α1),,e(αk0))(\theta_{1},\dots,\theta_{k_{0}})\neq(\mathfrak{R}e(\alpha_{1}),\dots,\mathfrak{R}e(\alpha_{k_{0}})). We deformation retract Dθ1,,θk0D_{\theta_{1},\dots,\theta_{k_{0}}} to {θ1+1m(α10),,θk0+1m(αk0)}\{\theta_{1}+\sqrt{-1}\mathfrak{I}m(\alpha_{1}^{0}),\dots,\theta_{k_{0}}+\sqrt{-1}\mathfrak{I}m(\alpha_{k_{0}})\}.

By Lemma 8.5, the determinant of the Hessian (4) is non-zero at the critical point.

For (5), we get

g(α10,,αk0)=ei𝕃(𝜶)k=1k0g~n(αk0).\displaystyle g(\alpha_{1}^{0},\dots,\alpha_{k_{0}})=e^{i\mathbb{L}(\bm{\alpha}})\prod_{k=1}^{k_{0}}\tilde{g}_{n}(\alpha_{k}^{0}).

Observe that none of the g~n(αk)\tilde{g}_{n}(\alpha_{k}) is zero, because e2iαk1e^{-2i\alpha_{k}}\neq-1 when αk(δ,π2δ)\alpha_{k}\in(\delta,\frac{\pi}{2}-\delta).

For the higher-order part of the function f(𝜶)f(\bm{\alpha}), we have O(1n2)O\left(\frac{1}{n^{2}}\right), which means vnv_{n} is bounded above by a constant independent of n. ∎

8.2. Estimate of other Fourier coefficients in the region Dδ,0D_{\delta,\textbf{0}}^{\mathbb{C}}

Proposition 8.7.

There is an ϵ>0\epsilon>0, such that when mk0m_{k}\neq 0 for some k=1,,k0k=1,\dots,k_{0}, then

|Dδ,𝟘ψ(𝕩)gn(𝜶)en4πi(f(𝜶)4π𝕞𝜶)𝑑𝜶|O(en4π(vol(Mφ)ϵ)).\displaystyle\left|\int_{D_{\delta,\mathbb{0}}^{\mathbb{C}}}\psi(\mathbb{x})g_{n}(\bm{\alpha})e^{\frac{n}{4\pi i}(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})}d\bm{\alpha}\right|\leq O\left(e^{\frac{n}{4\pi}\left(\operatorname{vol}(M_{\varphi})-\epsilon\right)}\right).
Proof.

For simplicity, assume m10m_{1}\neq 0. The same proof works if any other mk0m_{k}\neq 0. For L>0L>0, define

SL,top±={𝜶Dδ,𝟘|m(αk)=0 for k=2,,k0,m(α1)=±L}\displaystyle S_{L,\text{top}}^{\pm}=\left\{\bm{\alpha}\in D_{\delta,\mathbb{0}}^{\mathbb{C}}\;\big{|}\;\mathfrak{I}m(\alpha_{k})=0\text{ for }k=2,\dots,k_{0},\mathfrak{I}m(\alpha_{1})=\pm L\right\}

and

SL,side±={(x1+1l,x2,,xk0)|(x1,,xk0)Dδ,𝟘,l[0,L]}.\displaystyle S_{L,\text{side}}^{\pm}=\left\{(x_{1}+\sqrt{-1}l,x_{2},\dots,x_{k_{0}})\;\big{|}\;(x_{1},\dots,x_{k_{0}})\in\partial D_{\delta,\mathbb{0}}^{{\mathbb{R}}},\;l\in[0,L]\right\}.

Here, SL±=SL,side±SL,top±S_{L}^{\pm}=S_{L,\text{side}}^{\pm}\cup S_{L,\text{top}}^{\pm} is homotopic to Dδ,𝟘D_{\delta,\mathbb{0}}^{{\mathbb{R}}}. We split the proof into two cases.
Case I: m1>0m_{1}>0. We will show that for large enough L, there is an ϵ>0\epsilon>0 such that

m(f(𝜶)4π𝕞𝜶)<vol(Mφ)ϵ\displaystyle\mathfrak{I}m\;\left(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha}\right)<\operatorname{vol}(M_{\varphi})-\epsilon

for 𝕞(0,,0)\mathbb{m}\neq(0,\dots,0) on SL+S_{L}^{+}. Then the result will follow from Proposition 8.1.
We let α1=e(α1)+1l\alpha_{1}=\mathfrak{R}e\;(\alpha_{1})+\sqrt{-1}l for l>0l>0. Then

m(f(𝜶)4π𝕞𝜶)=mf(𝜶)4πm1l.\displaystyle\mathfrak{I}m\;(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})=\mathfrak{I}m\;f(\bm{\alpha})-4\pi m_{1}l.

The derivative with respect to ll gives us

(mf(𝜶)4πm1l)l=4arg(1+e2iα1)2(ϵ1+ϵ2+2)x1+4ϵ2x2+4ϵ1xk04πm1\displaystyle\frac{\partial\;(\mathfrak{I}m\;f(\bm{\alpha})-4\pi m_{1}l)}{\partial l}=-4\arg(1+e^{-2i\alpha_{1}})-2(\epsilon_{1}+\epsilon_{2}+2)x_{1}+4\epsilon_{2}x_{2}+4\epsilon_{1}x_{k_{0}}-4\pi m_{1}
<8x12(ϵ1+ϵ2+2)x1+4ϵ2x2+4ϵ1xk04πm1\displaystyle<8x_{1}-2(\epsilon_{1}+\epsilon_{2}+2)x_{1}+4\epsilon_{2}x_{2}+4\epsilon_{1}x_{k_{0}}-4\pi m_{1}

where the last inequality comes from the fact that 2x1<arg(1+e2iα1)<0-2x_{1}<\arg(1+e^{-2i\alpha_{1}})<0.
We deal with this inequality on a case-by-case basis. We use the fact that m1>0m_{1}>0 and 0<xk<π20<x_{k}<\frac{\pi}{2} for k=1,,k0k=1,\dots,k_{0}. Let δ=min{|xkπ2|;k=1,,k0}>0\delta=\min\{|x_{k}-\frac{\pi}{2}|;k=1,\dots,k_{0}\}>0. Then,
Case a: If ϵ1=1=ϵ2\epsilon_{1}=-1=\epsilon_{2}, then

RHS<8x14x24xk04πm1<8δ.\displaystyle\text{RHS}<8x_{1}-4x_{2}-4x_{k_{0}}-4\pi m_{1}<-8\delta.

Case b: If ϵ1=1,ϵ2=1\epsilon_{1}=-1,\epsilon_{2}=1, then

RHS<4x1+4x24xk04πm1<8δ.\displaystyle\text{RHS}<4x_{1}+4x_{2}-4x_{k_{0}}-4\pi m_{1}<-8\delta.

Case c: If ϵ1=1,ϵ2=1\epsilon_{1}=1,\epsilon_{2}=-1, then

RHS<4x14x2+4xk04πm1<8δ.\displaystyle\text{RHS}<4x_{1}-4x_{2}+4x_{k_{0}}-4\pi m_{1}<-8\delta.

Case d: If ϵ1=1=ϵ2\epsilon_{1}=1=\epsilon_{2}, then

RHS<4x2+4xk04πm1<8δ.\displaystyle\text{RHS}<4x_{2}+4x_{k_{0}}-4\pi m_{1}<-8\delta.

Thus, the function is strictly decreasing in ll. We push the domain along the 1l\sqrt{-1}l direction far enough, the imaginary part of f(𝜶)4πm1α1f(\bm{\alpha})-4\pi m_{1}\alpha_{1} becomes as small as possible. Therefore, there is ϵ>0\epsilon>0 such that

m(f(𝜶)4π𝕞𝜶)<vol(Mφ)ϵ\displaystyle\mathfrak{I}m\;\left(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha}\right)<\operatorname{vol}(M_{\varphi})-\epsilon

on SL,top+S_{L,\text{top}}^{+}.
m(f(𝜶)4π𝕞𝜶)\mathfrak{I}m\;(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha}) is smaller than the volume of MφM_{\varphi} on SL,top\partial S_{L,\text{top}} by lemma 6.2. Thus, it becomes even smaller on the side.

m(f(𝜶)4π𝕞𝜶)<vol(Mφ)ϵ\displaystyle\mathfrak{I}m\;(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})<\operatorname{vol}(M_{\varphi})-\epsilon

on SL,side+S_{L,\text{side}}^{+}.

The case when m1<0m_{1}<0 follows similarly. We work with SLS_{L}^{-}. Let α1=e(α1)1l\alpha_{1}=\mathfrak{R}e\;(\alpha_{1})-\sqrt{-1}l. Then

m(f(𝜶)4π𝕞𝜶)=mf(𝜶)+4πm1l\displaystyle\mathfrak{I}m\;(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})=\mathfrak{I}m\;f(\bm{\alpha})+4\pi m_{1}l

The derivative with respect to ll gives us

(mf(𝜶)+4πm1l)l=\displaystyle\frac{\partial\;(\mathfrak{I}m\;f(\bm{\alpha})+4\pi m_{1}l)}{\partial l}= 4arg(1+e2iα1)+2(ϵ1+ϵ2+2)x14ϵ2x24ϵ1xk0+4πm1\displaystyle 4\arg(1+e^{-2i\alpha_{1}})+2(\epsilon_{1}+\epsilon_{2}+2)x_{1}-4\epsilon_{2}x_{2}-4\epsilon_{1}x_{k_{0}}+4\pi m_{1}
<\displaystyle< 2(ϵ1+ϵ2+2)x14ϵ2x24ϵ1xk0+4πm1\displaystyle 2(\epsilon_{1}+\epsilon_{2}+2)x_{1}-4\epsilon_{2}x_{2}-4\epsilon_{1}x_{k_{0}}+4\pi m_{1}

where the last inequality comes from the fact that 2x1<arg(1+e2iα1)<0-2x_{1}<\arg(1+e^{-2i\alpha_{1}})<0.
We deal with this on a case-by-case basis. We use the fact that m1<0m_{1}<0 and 0<xk<π20<x_{k}<\frac{\pi}{2}. Let δ=min{|xkπ2|;k=1,,k0}>0\delta=\min\{|x_{k}-\frac{\pi}{2}|;k=1,\dots,k_{0}\}>0. Then,
Case I: If ϵ1=1=ϵ2\epsilon_{1}=-1=\epsilon_{2}, then

RHS<4x2+4xk0+4πm1<8δ.\displaystyle\text{RHS}<4x_{2}+4x_{k_{0}}+4\pi m_{1}<-8\delta.

Case II: If ϵ1=1,ϵ2=1\epsilon_{1}=-1,\epsilon_{2}=1, then

RHS<4x14x2+4xk0+4πm1<8δ.\displaystyle\text{RHS}<4x_{1}-4x_{2}+4x_{k_{0}}+4\pi m_{1}<-8\delta.

Case III: If ϵ1=1,ϵ2=1\epsilon_{1}=1,\epsilon_{2}=-1, then

RHS<4x1+4x24xk0+4πm1<8δ.\displaystyle\text{RHS}<4x_{1}+4x_{2}-4x_{k_{0}}+4\pi m_{1}<-8\delta.

Case IV: If ϵ1=1=ϵ2\epsilon_{1}=1=\epsilon_{2}, then

RHS<8x14x24xk0+4πm1<8δ.\displaystyle\text{RHS}<8x_{1}-4x_{2}-4x_{k_{0}}+4\pi m_{1}<-8\delta.

We push the domain along the 1l-\sqrt{-1}l direction far enough, that the imaginary part of f(𝜶)4πm1α1f(\bm{\alpha})-4\pi m_{1}\alpha_{1} becomes as small as possible. Therefore, there is ϵ>0\epsilon>0 such that

m(f(𝜶)4π𝕞𝜶)<vol(Mφ)ϵ\displaystyle\mathfrak{I}m\;\left(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha}\right)<\operatorname{vol}(M_{\varphi})-\epsilon

on SL,topS_{L,\text{top}}^{-}.
m(f(𝜶)4π𝕞𝜶)\mathfrak{I}m\;(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha}) is smaller than the volume of MφM_{\varphi} on SL,top\partial S_{L,\text{top}} by lemma 6.2. Thus, it becomes even smaller on the side.

m(f(𝜶)4π𝕞𝜶)<vol(Mφ)ϵ\displaystyle\mathfrak{I}m\;(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})<\operatorname{vol}(M_{\varphi})-\epsilon

on SL,sideS_{L,\text{side}}^{-}.

Therefore, on both SL±S_{L}^{\pm}, the imaginary part of f(𝜶)4π𝕞𝜶f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha} is less than the volume. ∎

8.3. Estimate of the leading Fourier coefficients in Dδ,JD_{\delta,J}^{\mathbb{C}}

We find the leading Fourier coefficients for each Dδ,JD_{\delta,J}^{\mathbb{C}} and show that it has a growth equal to the volume of the manifold.

Lemma 8.8.

For each J=(j1,,jk0)J=(j_{1},\dots,j_{k_{0}}), the leading Fourier coefficient in the domain Dδ,JD_{\delta,J}^{\mathbb{C}} is given by

𝕞=𝕂(J).\displaystyle\mathbb{m}=-\mathbb{K}(J).
Proof.

Consider the function

mf(𝜶+πJ)=mf(𝜶)m 4π𝕂(J)𝜶.\displaystyle\mathfrak{I}m\;f(\bm{\alpha}+\pi J)=\mathfrak{I}m\;f(\bm{\alpha})-\mathfrak{I}m\;4\pi\mathbb{K}(J)\cdot\bm{\alpha}.

Then,

m(f(𝜶+πJ)+4π𝕂(J)𝜶)=mf(𝜶).\displaystyle\mathfrak{I}m\;(f(\bm{\alpha}+\pi J)+4\pi\mathbb{K}(J)\cdot\bm{\alpha})=\mathfrak{I}m\;f(\bm{\alpha}).

Observe that,

(8.1) m(f(𝜶+πJ)+4π𝕂(J)𝜶)αk=mf(𝜶)αk=0.\displaystyle\frac{\partial\;\mathfrak{I}m\;(f(\bm{\alpha}+\pi J)+4\pi\mathbb{K}(J)\cdot\bm{\alpha})}{\partial\alpha_{k}}=\frac{\partial\;\mathfrak{I}m\;f(\bm{\alpha})}{\partial\alpha_{k}}=0.

Therefore, if 𝜶J=(𝜶𝟘+πJ)\bm{\alpha}^{J}=(\bm{\alpha}^{\mathbb{0}}+\pi J) is a critical point, then m(f(𝜶J)+4π𝕂(J)𝜶J)=mf(𝜶𝟘)=vol(Mφ)\mathfrak{I}m\;(f(\bm{\alpha}^{J})+4\pi\mathbb{K}(J)\cdot\bm{\alpha}^{J})=\mathfrak{I}m\;f(\bm{\alpha}^{\mathbb{0}})=\operatorname{vol}(M_{\varphi}). ∎

Remark 8.9.

Notice that the critical points are just translations by π\pi. In total, there are 2k02^{k_{0}} critical points.

We deform the integral domain Dδ,JD_{\delta,J}^{\mathbb{C}} to S~J\tilde{S}^{J} and estimate m(f(𝜶)+4πK(J)𝜶)\mathfrak{I}m\;(f(\bm{\alpha})+4\pi K(J)\cdot\bm{\alpha}).

Proposition 8.10.

The function

m(f(𝜶)+4πK(J)𝜶)\displaystyle\mathfrak{I}m\;(f(\bm{\alpha})+4\pi K(J)\cdot\bm{\alpha})

achieves the only absolute maximum at 𝛂J\bm{\alpha}^{J} on SJS^{J}.

Proof.

Follows directly from 8.1, Proposition 8.3 along with Lemma 8.8. ∎

Proposition 8.11.

For each J{0,1}k0J\in\{0,1\}^{k_{0}}, we have the following estimate

|Dδ,Jψ(α1,,αk0)gn(α1,,αk0)en4πi(f(α1,,αk0)+4π𝕂(J)𝜶)+O(1n)𝑑α1𝑑αk0|\displaystyle\left|\int_{D_{\delta,J}^{\mathbb{C}}}\psi(\alpha_{1},\dots,\alpha_{k_{0}})g_{n}(\alpha_{1},\dots,\alpha_{k_{0}})e^{\frac{n}{4\pi i}\left(f(\alpha_{1},\dots,\alpha_{k_{0}})+4\pi\mathbb{K}(J)\cdot\bm{\alpha}\right)+O\left(\frac{1}{n}\right)}d\alpha_{1}\dots d\alpha_{k_{0}}\right|
=cJ(1)k0detHess(f)(𝜶J)en4πvol(Mφ)(1+O(1n))\displaystyle=\frac{c_{J}}{\sqrt{(-1)^{k_{0}}\det\operatorname{Hess}(f)(\bm{\alpha}^{J})}}e^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})}\left(1+O\left(\frac{1}{n}\right)\right)

where cJc_{J} is a constant not equal to 0.

Proof.

Since the domain is shifted by π\pi, conditions (1), (2), (3), and (6) are verified.
For (4), notice that the terms in the Hessian that depend on αk\alpha_{k} are of the form e2iαke^{-2i\alpha_{k}}. Adding a π\pi to it does not change the value, therefore, the determinant of the hessian remains unchanged.
For (5), when we have αk+π\alpha_{k}+\pi instead of αk\alpha_{k}, we see a factor of (1ineUk2)(1ineU^k2)0(1-i^{n}e^{\frac{U_{k}}{2}})(1-i^{n}e^{\frac{\hat{U}_{k}}{2}})\neq 0 because eUk2±ie^{\frac{U_{k}}{2}}\neq\pm i as eUk1e^{U_{k}}\neq-1 by assumption. ∎

8.4. Estimate of other Fourier coefficients in the region Dδ,JD_{\delta,J}^{\mathbb{C}}

Proposition 8.12.

There is an ϵ>0\epsilon>0, such that when 𝕞𝕂(J)\mathbb{m}\neq-\mathbb{K}(J), then

|Dδ,Jψ(𝜶)g(𝜶)en4πi(f(𝜶)4π𝕞𝜶)𝑑𝜶|O(en4π(vol(Mφ)ϵ)).\displaystyle\left|\int_{D_{\delta,J}^{\mathbb{C}}}\psi(\bm{\alpha})g(\bm{\alpha})e^{\frac{n}{4\pi i}(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})}d\bm{\alpha}\right|\leq O\left(e^{\frac{n}{4\pi}\left(\operatorname{vol}(M_{\varphi})-\epsilon\right)}\right).
Proof.

For simplicity, assume m1(ϵ1+ϵ2+2)j1/2+2ϵ1jk0+2ϵ2j2m_{1}\neq-(\epsilon_{1}+\epsilon_{2}+2)j_{1}/2+2\epsilon_{1}j_{k_{0}}+2\epsilon_{2}j_{2}. The two cases are:
Case I: m1>(ϵ1+ϵ2+2)j1/2+2ϵ1jk0+2ϵ2j2m_{1}>-(\epsilon_{1}+\epsilon_{2}+2)j_{1}/2+2\epsilon_{1}j_{k_{0}}+2\epsilon_{2}j_{2}
Case II: m1<(ϵ1+ϵ2+2)j1/2+2ϵ1jk0+2ϵ2j2m_{1}<-(\epsilon_{1}+\epsilon_{2}+2)j_{1}/2+2\epsilon_{1}j_{k_{0}}+2\epsilon_{2}j_{2}.
Observe that

m(f(𝜶+πJ)4π𝕞𝜶)\displaystyle\mathfrak{I}m\;(f(\bm{\alpha}+\pi J)-4\pi\mathbb{m}\cdot\bm{\alpha}) =m(f(𝜶+πJ)+4π𝕂(J)𝜶4π(𝕞+𝕂(J))𝜶)\displaystyle=\mathfrak{I}m\;(f(\bm{\alpha}+\pi J)+4\pi\mathbb{K}(J)\cdot\bm{\alpha}-4\pi(\mathbb{m}+\mathbb{K}(J))\cdot\bm{\alpha})
=mf(𝜶)m(4π(𝕞+𝕂(J))𝜶).\displaystyle=\mathfrak{I}m\;f(\bm{\alpha})-\mathfrak{I}m\;(4\pi(\mathbb{m}+\mathbb{K}(J))\cdot\bm{\alpha}).

Thus, by a change of variable 𝜶Dδ,J\bm{\alpha}\in D_{\delta,J}^{\mathbb{C}} to 𝜶+πJDδ,J\bm{\alpha}+\pi J\in D_{\delta,J}^{\mathbb{C}} and then shifting the domain by subtracting πJ\pi J, we have the same estimate as the previous Proposition 8.7. ∎

9. Asymptotics

We first show that the sum of the smaller Fourier coefficients is still small. Then, the leading Fourier coefficients determine the sum and we prove the conjecture.

Proposition 9.1.

The sum of all Fourier coefficients other than the 2k02^{k_{0}} leading ones is small.

(𝜶,𝕞)(𝜶J,𝕂(J))|G^n(𝜶,𝕞)|O(en4π(vol(Mφ)ϵ))\displaystyle\sum_{(\bm{\alpha},\mathbb{m})\neq(\bm{\alpha}^{J},\mathbb{K}(J))}|\hat{G}_{n}(\bm{\alpha},\mathbb{m})|\leq O(e^{\frac{n}{4\pi}(\operatorname{vol}(M_{\varphi})-\epsilon)})
Proof.

This follows straight from the fact that G^n\hat{G}_{n} has compact support, each term of the summand is bounded by O(en4π(vol(Mφ)ϵ))O(e^{\frac{n}{4\pi}(\operatorname{vol}(M_{\varphi})-\epsilon)}) and the Fourier transformation has a convergence tail. For each 𝕞\mathbb{m}, consider the function ψ(𝜶)gn(𝜶)\psi(\bm{\alpha})g_{n}(\bm{\alpha}). For each 𝕞\mathbb{m}, since ψ(𝜶)\psi(\bm{\alpha}) vanishes outside of 𝒟\mathcal{D}^{{\mathbb{R}}}, by integration by parts we have

r2k0j=1k0mj2k0𝒟ψ(𝜶)\displaystyle r^{2k_{0}}\sum_{j=1}^{k_{0}}m_{j}^{2k_{0}}\int_{\mathcal{D}^{{\mathbb{R}}}}\psi(\bm{\alpha}) gn(𝜶)en4π1(f(𝜶)4π𝕞𝜶)d𝜶\displaystyle g_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})}d\bm{\alpha}
=𝒟ψ(𝜶)gn(𝜶)en4π1f(𝜶)((j=1k02k0αj2k0)en4π1(4π𝕞𝜶))𝑑𝜶\displaystyle=\int_{\mathcal{D}^{{\mathbb{R}}}}\psi(\bm{\alpha})g_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}f(\bm{\alpha})}\left(\left(\sum_{j=1}^{k_{0}}\frac{\partial^{2k_{0}}}{\partial\alpha_{j}^{2k_{0}}}\right)e^{\frac{n}{4\pi\sqrt{-1}}(-4\pi\mathbb{m}\cdot\bm{\alpha})}\right)d\bm{\alpha}
=𝒟((j=1k02k0α2k0)(ψ(𝜶)gn(𝜶)en4π1f(𝜶)))en4π14π𝕞𝜶𝑑𝜶\displaystyle=\int_{\mathcal{D}^{{\mathbb{R}}}}\left(\left(\sum_{j=1}^{k_{0}}\frac{\partial^{2k_{0}}}{\partial\alpha^{2k_{0}}}\right)\left(\psi(\bm{\alpha})g_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}f(\bm{\alpha})}\right)\right)e^{\frac{n}{4\pi\sqrt{-1}}-4\pi\mathbb{m}\cdot\bm{\alpha}}d\bm{\alpha}
=𝒟g~n(𝜶)en4π1(f(𝜶)4π𝕞𝜶)𝑑𝜶,\displaystyle=\int_{\mathcal{D}^{{\mathbb{R}}}}\tilde{g}_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})}d\bm{\alpha},

where

g~n(𝜶)=(j=1k02k0αj2k0)(ψ(𝜶)gn(𝜶)en4π1f(𝜶))en4π1f(𝜶)\displaystyle\tilde{g}_{n}(\bm{\alpha})=\frac{\left(\sum_{j=1}^{k_{0}}\frac{\partial^{2k_{0}}}{\partial\alpha_{j}^{2k_{0}}}\right)\left(\psi(\bm{\alpha})g_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}f(\bm{\alpha})}\right)}{e^{\frac{n}{4\pi\sqrt{-1}}f(\bm{\alpha})}}

is a smooth function independent of 𝕞\mathbb{m} and has the form

g~n(𝜶)=g~(𝜶)n2k0+O(n2k01)\displaystyle\tilde{g}_{n}(\bm{\alpha})=\tilde{g}(\bm{\alpha})n^{2k_{0}}+O(n^{2k_{0}-1})

for a smooth function g~(𝜶)\tilde{g}(\bm{\alpha}) independent of nn. Therefore, on a compact subset of 𝒟,g~n(𝜶)/n2k0\mathcal{D}^{\mathbb{C}},\tilde{g}_{n}(\bm{\alpha})/n^{2k_{0}} is bounded from above by some constant C>0C>0 independent of 𝕞\mathbb{m}. Then,

𝕞𝕂(J)|G^n(𝜶,𝕞)|\displaystyle\sum_{\mathbb{m}\neq\mathbb{K}(J)}|\hat{G}_{n}(\bm{\alpha},\mathbb{m})| =(n2π)k0𝕞𝕂(J)|𝒟ψ(𝜶)gn(𝜶)en4π1(f(𝜶)4π𝕞𝜶)𝑑𝜶|\displaystyle=\left(\frac{n}{2\pi}\right)^{k_{0}}\sum_{\mathbb{m}\neq\mathbb{K}(J)}\left|\int_{\mathcal{D}^{{\mathbb{R}}}}\psi(\bm{\alpha})g_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})}d\bm{\alpha}\right|
=(n2π)k0𝕞𝕂(J)1m12k0++mk02k0|𝒟g~n(𝜶)en4π1(f(𝜶)4π𝕞𝜶)𝑑𝜶|\displaystyle=\left(\frac{n}{2\pi}\right)^{k_{0}}\sum_{\mathbb{m}\neq\mathbb{K}(J)}\frac{1}{m_{1}^{2k_{0}}+\dots+m_{k_{0}}^{2k_{0}}}\left|\int_{\mathcal{D}^{{\mathbb{R}}}}\tilde{g}_{n}(\bm{\alpha})e^{\frac{n}{4\pi\sqrt{-1}}(f(\bm{\alpha})-4\pi\mathbb{m}\cdot\bm{\alpha})}d\bm{\alpha}\right|
(n2π)k0𝕞𝕂(J)3ACl=1k0ml2k0O(en4π(vol(Mφ)ϵ))O(en4π(vol(Mφ)ϵ2))\displaystyle\leq\left(\frac{n}{2\pi}\right)^{k_{0}}\sum_{\mathbb{m}\neq\mathbb{K}(J)}\frac{3AC}{\sum_{l=1}^{k_{0}}m_{l}^{2k_{0}}}O\left(e^{\frac{n}{4\pi}(\operatorname{vol}(M_{\varphi})-\epsilon)}\right)\leq O\left(e^{\frac{n}{4\pi}(\operatorname{vol}(M_{\varphi})-\tfrac{\epsilon}{2})}\right)

where AA is a constant depending on the area of the integral. The summation converges because 𝕞𝟘1m12k0++mk02k0\sum_{\mathbb{m}\neq\mathbb{0}}\frac{1}{m_{1}^{2k_{0}}+\dots+m_{k_{0}}^{2k_{0}}} does. ∎

Therefore, the integral with leading Fourier coefficients can be split into multiple parts (2k02^{k_{0}} parts to be precise) and for each of these deformed domains, we have an estimate.

Lemma 9.2.

For a generic character [r~][\tilde{r}] in the same component as the hyperbolic character [r~hyp][\tilde{r}_{\text{hyp}}], the sum

J{0,1}k0gn(𝜶J)0.\displaystyle\sum_{J\in\{0,1\}^{k_{0}}}g_{n}(\bm{\alpha}^{J})\neq 0.
Proof.

Recall the definition of the function gng_{n} 5.4 at 𝜶J\bm{\alpha}^{J}

gn(𝜶J)=ei𝕃(𝜶J)k=1k0(1ineUk2)jk(1ineU^k2)jk(1+e2iαk2)1Uk+U^k4πieαk(Vk+V~k)2π\displaystyle g_{n}(\bm{\alpha}^{J})=e^{i\mathbb{L}(\bm{\alpha}^{J})}\prod_{k=1}^{k_{0}}\left(1-i^{n}e^{\frac{U_{k}}{2}}\right)^{j_{k}}\left(1-i^{n}e^{\frac{\hat{U}_{k}}{2}}\right)^{j_{k}}\left(\frac{1+e^{-2i\alpha_{k}}}{2}\right)^{1-\frac{U_{k}+\hat{U}_{k}}{4\pi i}}e^{-\frac{\alpha_{k}(V_{k}+\tilde{V}_{k})}{2\pi}}
=ei𝕃(𝜶𝟘+πJ)k=1k0(1+e2iαk2)1Uk+U^k4πieαk(Vk+V~k)2πk=1k0(1ineUk2)jk(1ineU^k2)jk.\displaystyle=e^{i\mathbb{L}(\bm{\alpha}^{\mathbb{0}}+\pi J)}\prod_{k=1}^{k_{0}}\left(\frac{1+e^{-2i\alpha_{k}}}{2}\right)^{1-\frac{U_{k}+\hat{U}_{k}}{4\pi i}}e^{-\frac{\alpha_{k}(V_{k}+\tilde{V}_{k})}{2\pi}}\prod_{k=1}^{k_{0}}\left(1-i^{n}e^{\frac{U_{k}}{2}}\right)^{j_{k}}\left(1-i^{n}e^{\frac{\hat{U}_{k}}{2}}\right)^{j_{k}}.

Therefore,

J{0,1}k0gn(𝜶J)=\displaystyle\sum_{J\in\{0,1\}^{k_{0}}}g_{n}(\bm{\alpha}^{J})= ei𝕃(𝜶𝟘)k=1k0(1+e2iαk2)1Uk+U^k4πieαk(Vk+V~k)2π\displaystyle e^{i\mathbb{L}(\bm{\alpha}^{\mathbb{0}})}\prod_{k=1}^{k_{0}}\left(\frac{1+e^{-2i\alpha_{k}}}{2}\right)^{1-\frac{U_{k}+\hat{U}_{k}}{4\pi i}}e^{-\frac{\alpha_{k}(V_{k}+\tilde{V}_{k})}{2\pi}}
×J{0,1}k0(1)𝕃(J)k=1k0(1ineUk2)jk(1ineU^k2)jk\displaystyle\times\sum_{J\in\{0,1\}^{k_{0}}}(-1)^{\mathbb{L}(J)}\prod_{k=1}^{k_{0}}\left(1-i^{n}e^{\frac{U_{k}}{2}}\right)^{j_{k}}\left(1-i^{n}e^{\frac{\hat{U}_{k}}{2}}\right)^{j_{k}}
=\displaystyle= ei𝕃(𝜶𝟘)k=1k0(1+e2iαk2)1Uk+U^k4πieαk(Vk+V~k)2π\displaystyle e^{i\mathbb{L}(\bm{\alpha}^{\mathbb{0}})}\prod_{k=1}^{k_{0}}\left(\frac{1+e^{-2i\alpha_{k}}}{2}\right)^{1-\frac{U_{k}+\hat{U}_{k}}{4\pi i}}e^{-\frac{\alpha_{k}(V_{k}+\tilde{V}_{k})}{2\pi}}
×(1+(1)l^1(1ineU1/2))(1+(1)l^2(1ineUk0/2))k=2k01(2ineUk/2)\displaystyle\times\left(1+(-1)^{\hat{l}_{1}}\left(1-i^{n}e^{U_{1}/2}\right)\right)\left(1+(-1)^{\hat{l}_{2}}\left(1-i^{n}e^{U_{k_{0}}/2}\right)\right)\prod_{k=2}^{k_{0}-1}\left(2-i^{n}e^{U_{k}/2}\right)

Let’s try to see the possible cases when the sum can be zero.

  1. (1)

    e2iαk=1e^{-2i\alpha_{k}}=-1 for any k=1,,k0k=1,\dots,k_{0}. However, e2iαk=ak1=eUk1e^{-2i\alpha_{k}}=a_{k}^{-1}=e^{U_{k}}\neq-1. Therefore, this can not happen.

  2. (2)

    2=ineUk/22=i^{n}e^{U_{k}/2} for any k=2,,k01k=2,\dots,k_{0}-1. This would imply eUk=4=zke^{U_{k}}=-4=-z_{k} where zkz_{k} is the shape parameter. For a geometric triangulation, the imaginary part of zkz_{k} is positive. Thus, this can not happen either.

  3. (3)

    If l^1\hat{l}_{1} is even or l^2\hat{l}_{2} is even, we get case 2. Using the same argument as before, we see that the term (2ineU1/2)0(2-i^{n}e^{U_{1}/2})\neq 0 or (2ineUk0/2)0(2-i^{n}e^{U_{k_{0}}/2})\neq 0.

  4. (4)

    If l^1\hat{l}_{1} or l^2\hat{l}_{2} is odd, then we are left with ineU1/2-i^{n}e^{U_{1}/2} or ineUk0/2-i^{n}e^{U_{k_{0}}/2}, which are non-zero.

Thus, we see that the sum is never equal to 0, as long as we have a geometric triangulation. ∎

Proof of theorem 1.2.

Thus our invariant |TraceΛφ,r~q||\operatorname{Trace}\Lambda^{q}_{\varphi,\tilde{r}}| can be written as

=|1nk0/2k=1k0Dq(uk)1nDq(u^k)1n𝕞k0(n2π)k0𝒟Gn(𝜶)e1n𝕞𝜶𝑑𝜶+O(en4πvol(Mφ)ϵ)|\displaystyle=\left|\frac{1}{n_{k_{0}/2}\prod_{k=1}^{k_{0}}D^{q}(u_{k})^{\frac{1}{n}}D^{q}(\hat{u}_{k})^{\frac{1}{n}}}\sum_{\mathbb{m}\in{\mathbb{Z}}^{k_{0}}}\left(\frac{n}{2\pi}\right)^{k_{0}}\int_{\mathcal{D}^{{\mathbb{R}}}}G_{n}(\bm{\alpha})e^{\sqrt{-1}n\mathbb{m}\cdot\bm{\alpha}}d\bm{\alpha}+O\left(e^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})-\epsilon}\right)\right|
=nk0/2(2π)k0k=1k0|Dq(uk)Dq(u^k)|1n|(8π2)k0/2nk0/2(1)k0detHess(f)(𝜶𝟘)J{0,1}k0gn(𝜶J)en4πvol(Mφ)|\displaystyle=\frac{n^{k_{0}/2}}{(2\pi)^{k_{0}}\prod_{k=1}^{k_{0}}\left|D^{q}(u_{k})D^{q}(\hat{u}_{k})\right|^{\frac{1}{n}}}\left|\frac{(8\pi^{2})^{k_{0}/2}}{n^{k_{0}/2}\sqrt{(-1)^{k_{0}}\det\operatorname{Hess}(f)(\bm{\alpha}^{\mathbb{0}})}}\sum_{J\in\{0,1\}^{k_{0}}}g_{n}(\bm{\alpha}^{J})e^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})}\right|
=1k=1k0|Dq(uk)Dq(u^k)|1n|J{0,1}k0gn(𝜶J)|(1)k0detHess(f)(𝜶𝟘)en4πvol(Mφ)(1+O(1n))\displaystyle=\frac{1}{\prod_{k=1}^{k_{0}}\left|D^{q}(u_{k})D^{q}(\hat{u}_{k})\right|^{\frac{1}{n}}}\frac{\left|\sum_{J\in\{0,1\}^{k_{0}}}g_{n}(\bm{\alpha}^{J})\right|}{\sqrt{(-1)^{k_{0}}\det\operatorname{Hess}(f)(\bm{\alpha}^{\mathbb{0}})}}e^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})}\left(1+O\left(\frac{1}{n}\right)\right)
=cen4πvol(Mφ)(1+O(1n)),\displaystyle=ce^{\frac{n}{4\pi}\operatorname{vol}(M_{\varphi})}\left(1+O\left(\frac{1}{n}\right)\right),

where c 0\neq 0 is a constant defined by

c=|J{0,1}k0gn(𝜶J)|(1)k0detHess(f)(𝜶𝟘)k=1k0|Dq(uk)Dq(u^k)|1n.\displaystyle c=\frac{\left|\sum_{J\in\{0,1\}^{k_{0}}}g_{n}(\bm{\alpha}^{J})\right|}{\sqrt{(-1)^{k_{0}}\det\operatorname{Hess}(f)(\bm{\alpha}^{\mathbb{0}})}\prod_{k=1}^{k_{0}}\left|D^{q}(u_{k})D^{q}(\hat{u}_{k})\right|^{\frac{1}{n}}}.

Corollary 9.3.

When the diffeomorphism word is LRLR, we get the same potential function (with a factor of 2), and therefore the same critical point as [BWY2]. Therefore, the volume is 4v34v_{3} and the traingulation of mapping torus has four tetrahedra where the volume of each of them is v3v_{3}.

Remark 9.4.

The condition that rr is generic for uniqueness of representations from the Kauffman Bracket Skein algebra is also hoped to be extended to all irreducible characters (like in the case of closed surfaces).

References

  • [AGN] F. B. Aribi, F. Guéritaud, E. Piguet-Nakazawa. Geometric triangulations and the Teichmüller TQFT volume conjecture for twist knots. 2021. hal-03103094
  • [BP] R. Benedetti, C. Petronio, Lectures on Hyperbolic Geometry, Springer Berlin, Heidelberg, 1992, https://doi.org/10.1007/978-3-642-58158-8
  • [BL07] F. Bonahon, X. Liu , Representations of the quantum Teichmüller space and invariants of surface diffeomorphisms, Geom. Topol. 11 (2007), 889–937.
  • [BW1] F. Bonahon, H. Wong, Quantum traces for representations of surface groups in SL2(C), Geom. Topol. 15 (2011), no. 3, 1569–1615.
  • [BW16] F. Bonahon, H. Wong, Representations of the Kauffman bracket skein algebra I: invariants and miraculous cancellations, Invent. Math. 204 (2016), no. 1, 195–243.
  • [BW16II] F. Bonahon, H. Wong, Representations of the Kauffman bracket skein algebra II: Punctured surfaces, Algebr. Geom. Topol. 17 (2017), no. 6, 3399–3434.
  • [BWY1] F. Bonahon, H. Wong, T. Yang, Asymptotics of quantum invariants of surface diffeomorphisms I: conjecture and algebraic computations, arXiv: 2112.12852.
  • [BWY2] F. Bonahon, H. Wong, T. Yang, Asymptotics of quantum invariants of surface diffeomorphisms II: The figure eight knot complement, arXiv: 2203.05730.
  • [BWY3] F. Bonahon, H. Wong, T. Yang, Asymptotics of quantum invariants of surface diffeomorphisms III: The one-puncture torus, in preparation.
  • [CF] L. O. Chekhov and V. V. Fok, Quantum Teichmüller spaces, Teoret. Mat. Fiz. 120 (1999), no. 3, 511–528. MR 1737362.
  • [CY] Q. Chen and T. Yang, Volume Conjectures for the Reshetikhin-Turaev and the Turaev-Viro Invariants, Quantum Topol. 9 (2018), no. 3, 419–460.
  • [DG] T. Dimofte and S. Garoufalidis, The quantum content of the gluing equations, Geometry & Topology 17 (2013) 1253-1315.
  • [FM] Farb, Benson, and Dan Margalit. A Primer on Mapping Class Groups (PMS-49). Princeton University Press, 2012. JSTOR, http://www.jstor.org/stable/j.ctt7rkjw.
  • [FKBL19] C. Frohman, J. Kania-Bartoszyńska, T. Lê, Unicity for representations of the Kauffman bracket skein algebra, Invent. Math. 215 (2019), no. 2, 609–650.
  • [GF] F. Guéritaud, On canonical triangulations of once-punctured torus bundles and two-bridge link complements, Geometry & Topology, Geom. Topol. 10(3), 1239-1284, (2006).
  • [GJS19] Iordan Ganev, David Jordan, and Pavel Safronov, The quantum Frobenius for character varieties and multiplicative quiver varieties, preprint, arXiv:1901.11450.
  • [K97] R. Kashaev, The hyperbolic volume of knots from the quantum dilogarithm, Lett. Math. Phys. 39 (1997), no. 3, 269–275.
  • [MM01] H. Murakami and J. Murakami, The colored Jones polynomials and the simplicial volume of a knot, Acta Math. 186 (2001), no. 1, 85–104.
  • [MMOTY] H. Murakami, J. Murakami, M. Okamoto, T. Takata and Y. Yokota, Kashaev’s conjecture and the Chern-Simons invariants of knots and links, Experiment. Math. 11 (2002) 427–435.
  • [NZ] W. Neumann, D. Zagier, Volumes of hyperbolic three-manifolds, Topology 24 (1985) 307–332.
  • [N] W. Neumann, Combinatorics of triangulations and the Chern–Simons invariant for hyperbolic 3-manifolds, from: “Topology ’90”, (B Apanasov, W D Neumann, A W Reid, L Siebenmann, editors), Ohio State Univ. Math. Res. Inst. Publ. 1, de Gruyter, Berlin (1992) 243–271.
  • [O16] T. Ohtsuki, On the asymptotic expansion of the Kashaev invariant of the 525_{2} knot, Quantum Topol. 7 (2016), no. 4, 669–735.
  • [OY] T. Ohtsuki and Y. Yokota, On the asymptotic expansions of the Kashaev invariant of the knots with 6 crossings, Mathematical Proceedings of the Cambridge Philosophical Society, Volume 165, Issue 2, September 2018, pp. 287–339.
  • [O17] T. Ohtuski, On the asymptotic expansion of the Kashaev invariant of the hyperbolic knots with seven crossings, Internat. J. Math. 28 (2017), no. 13, 1750096.
  • [SS] E. Stein and R. Shakarchi, Fourier analysis, An introduction. Princeton Lectures in Analysis, 1. Princeton University Press, Princeton, NJ, 2003. xvi+311 pp. ISBN: 0-691-11384-X.
  • [V08] R. van der Veen, Proof of the volume conjecture for Whitehead chains, Acta Mathematica Vietnamica, Volume 33, Number 3, 2008, pp. 421-431.
  • [W] K. H. Wong, Private communications.
  • [WY20] K. H. Wong and T. Yang, On the Volume Conjecture for hyperbolic Dehn-filled 3-manifolds along the figure-eight knot , Preprint, arXiv:2003.10053.
  • [WY21] K. H. Wong and T. Yang, Relative Reshetikhin-Turaev invariants, hyperbolic cone metrics and discrete Fourier transforms I, Commun. Math. Phys. 400, 1019–1070 (2023). https://doi.org/10.1007/s00220-022-04613-5.