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

Adjoint Reidemeister torsions of two-bridge knots

Seokbeom Yoon [email protected]
Abstract.

We give an explicit formula for the adjoint Reidemeister torsion of two-bridge knots and prove that the adjoint Reidemeister torsion satisfies a certain type of vanishing identities.

Key words and phrases:
Reidemeister torsion, adjoint representation, two-bridge knot, Riley polynomial, character variety

1. Introduction

Let KK be a knot in S3S^{3} and MM be the knot exterior. For an irreducible representation ρ:π1(M)SL2()\rho:\pi_{1}(M)\rightarrow\mathrm{SL}_{2}(\mathbb{C}) the adjoint Reidemeister torsion 𝕋γ(ρ)\mathbb{T}_{\gamma}(\rho) is defined, under some reasonable assumptions, as the sign-refined algebraic torsion of the cochain complex of MM with the coefficient 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{C}) twisted by the adjoint representation of ρ\rho. Here γ\gamma is a simple closed curve in M\partial M which has a role in specifying a basis of the twisted cohomology group H(M;𝔰𝔩2()ρ)H^{\ast}(M;\mathfrak{sl}_{2}(\mathbb{C})_{\rho}), see [Por97, Dub03]. It is known that 𝕋γ(ρ)\mathbb{T}_{\gamma}(\rho) is invariant under conjugation, so the notion of adjoint Reidemeister torsion is also well-defined for the character of ρ\rho.

The adjoint torsion is quite complicated to compute in general and its explicit formula is known only in a few examples, [DHY09, Tra14, Tra18]. All successful computations so far first compute the adjoint twisted Alexander polynomial ΔKAdρ(t)\Delta_{K}^{\mathrm{Ad}\rho}(t) and then obtain the adjoint torsion from the formula of [Yam08]

𝕋λ(ρ)=limt1ΔKAdρ(t)t1,λ:the canonical longitude of K.\mathbb{T}_{\lambda}(\rho)=-\lim_{t\rightarrow 1}\frac{\Delta_{K}^{\mathrm{Ad}\rho}(t)}{t-1},\quad\lambda:\textrm{the canonical longitude of $K$.}

It is computationally advantageous that the adjoint twisted Alexander polynomial is defined from an acyclic chain complex. However, the computation itself would be further complicated, as an indeterminant tt interferes in the Fox differential calculus.

In this paper, we give an explicit formula for the adjoint Reidemeister torsion of two-bridge knots. Our computation uses a well-known observation in [Wei64] that relates the twisted cohomology group and the tangent space of the character variety. We stress that it is simple and effective, compared to the method using the adjoint twisted Alexander polynomial, as we do not require the Fox differential calculus and we can directly obtain a relation of the adjoint torsion and the character variety. We here summarize our results briefly.

Let KK be a two-bridge knot given by the Schubert normal form (p,q)(p,q). Here p>0p>0 and qq are relatively prime odd integers with p<q<p-p<q<p. The knot group of KK has a presentation

π1(M)=g1,g2|wg1w1g21\pi_{1}(M)=\langle g_{1},g_{2}\,|\,wg_{1}w^{-1}g_{2}^{-1}\rangle

where w=g1ϵ1g2ϵ2g1ϵp2g2ϵp1w=g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{p-2}}g_{2}^{\epsilon_{p-1}} and ϵi=(1)iqp\epsilon_{i}=(-1)^{\lfloor\frac{iq}{p}\rfloor}. Here x\lfloor x\rfloor means the greatest integer less than or equal to xx\in\mathbb{R}. Up to conjugation, an irreducible representation ρ:π1(M)SL2()\rho:\pi_{1}(M)\rightarrow\mathrm{SL}_{2}(\mathbb{C}) is given by

ρ(g1)=(m10m1),ρ(g2)=(m0um1)\rho(g_{1})=\begin{pmatrix}m&1\\ 0&m^{-1}\end{pmatrix},\ \rho(g_{2})=\begin{pmatrix}m&0\\ -u&m^{-1}\end{pmatrix}

for some m0m\neq 0 and u0u\neq 0 satisfying the Riley polynomial of KK, [Ril84]. We denote by χρ\chi_{\rho} the character of ρ\rho and let ϕg:=g11(mm1)g12\phi_{g}:=g_{11}-(m-m^{-1})g_{12} for gπ1(M)g\in\pi_{1}(M) where gij[m±1,u]g_{ij}\in\mathbb{Z}[m^{\pm 1},u] is the (i,j)(i,j)-entry of ρ(g)\rho(g). Note that the Riley polynomial of KK is equal to ϕw\phi_{w}.

Theorem 1.1 (Theorem 3.1).

Suppose that χρ\chi_{\rho} is μ\mu-regular with m±1m\neq\pm 1, ϕwm0\frac{\partial\phi_{w}}{\partial m}\neq 0, and ϕwu0\frac{\partial\phi_{w}}{\partial u}\neq 0. Then we have

𝕋μ(χρ)=±mϵk+12(m21)w11v11ϕvϕwu\mathbb{T}_{\mu}(\chi_{\rho})=\pm\,\frac{m^{\epsilon_{k}+1}}{2(m^{2}-1)}\frac{w_{11}}{v_{11}\phi_{v}}\frac{\partial\phi_{w}}{\partial u}

where v=g1ϵ1g2ϵ2g1ϵk2g2ϵk1v=g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{k-2}}g_{2}^{\epsilon_{k-1}} and 0<k<p0<k<p is a unique (odd) integer satisfying kq±1(mod 2p)kq\equiv\pm 1\ (\mathrm{mod}\ 2p). Here the sign ±\pm does not depend on the choice of χρ\chi_{\rho}.

The adjoint torsion has intriguing and fruitful interactions with quantum field theory, see e.g. [Wit89, Guk05]. Recently, it is showed in [BGZ, GKZ] that a certain sum of adjoint torsions realizes the so-called Witten index. As a mathematical byproduct, it is conjectured that if a hyperbolic knot KK has the character variety XKX_{K} of irreducible SL2()\mathrm{SL}_{2}(\mathbb{C})-representations consisting of 1-dimensional components (which is the case of hyperbolic two-bridge knots), then we have

(1) χρtrμ1(c)1𝕋μ(χρ)=0for generic c\sum_{\chi_{\rho}\in\mathrm{tr}_{\mu}^{-1}(c)}\frac{1}{\mathbb{T}_{\mu}(\chi_{\rho})}=0\quad\textrm{for generic }c\in\mathbb{C}

where trμ:XK\mathrm{tr}_{\mu}:X_{K}\rightarrow\mathbb{C} is the trace function of a meridian μ\mu. See [GKY] for details.

Theorem 1.2 (Theorem 3.4).

The equation (1) holds for all hyperbolic two-bridge knots.

The paper is organized as follows. We review some backgrounds on the adjoint Reidemeister torsion in Section 2. We state our main results, Theorems 3.1 and 3.4, in Section 3 and give proofs of Theorems 3.1 and 3.4 in Sections 3.1 and 3.2, respectively.

Acknowledgment

The author is supported by a KIAS Individual Grant (MG073801) at Korea Institute for Advanced Study.

2. Preliminaries: the adjoint Reidemeister torsion

2.1. Basic definitions

We briefly recall some basic definitions and known results that we need in the following sections. We mainly follow [Dub06] and refer to [Por97, Tur02, Dub03] for details. In what follows, we denote by 𝔤\mathfrak{g} the Lie algebra of SL2()\mathrm{SL}_{2}(\mathbb{C}) with a standard basis

(2) e1=(0100),e2=(1001),e3=(0010)e_{1}=\begin{pmatrix}0&1\\ 0&0\end{pmatrix},\ e_{2}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix},\ e_{3}=\begin{pmatrix}0&0\\ 1&0\end{pmatrix}

and ,𝔤\langle\cdot,\cdot\rangle_{\mathfrak{g}} the Killing form of 𝔤\mathfrak{g}

(bacb),(bacb)𝔤=8bb+4(ac+ca).\left\langle\begin{pmatrix}b&a\\ c&-b\end{pmatrix},\begin{pmatrix}b^{\prime}&a^{\prime}\\ c^{\prime}&-b^{\prime}\end{pmatrix}\right\rangle_{\mathfrak{g}}=8bb^{\prime}+4(ac^{\prime}+ca^{\prime}).

Let C=(0CnC00)C_{\ast}=(0\rightarrow C_{n}\rightarrow\cdots\rightarrow C_{0}\rightarrow 0) be a chain complex of \mathbb{C}-vector spaces with boundary maps i:CiCi1\partial_{i}:C_{i}\rightarrow C_{i-1}. For a given basis cc_{\ast} of CC_{\ast} and a given basis hh_{\ast} of the homology group H(C)H_{\ast}(C_{\ast}), the algebraic torsion tor(C,c,h)\mathrm{tor}(C_{\ast},c_{\ast},h_{\ast}) is defined as follows. For each 0in0\leq i\leq n let bib_{i} be any sequence of vectors in CiC_{i} such that i(bi)\partial_{i}(b_{i}) is a basis of Imi\mathrm{Im}\,\partial_{i} and let h~i\widetilde{h}_{i} be any representative of hih_{i} in CiC_{i}. We then obtain a new basis of CiC_{i} by combining i+1(bi+1)\partial_{i+1}(b_{i+1}), h~i\widetilde{h}_{i}, and bib_{i} in order and let

(3) tor(C,c,h):=i=0n[(i+1(bi+1),h~i,bi)/ci](1)i+1.\mathrm{tor}(C_{\ast},c_{\ast},h_{\ast}):=\prod_{i=0}^{n}\left[(\partial_{i+1}(b_{i+1}),\widetilde{h}_{i},b_{i})/c_{i}\right]^{(-1)^{i+1}}\in\mathbb{C}^{\ast}.

Here [x/y][x/y] means the determinant of the transition matrix sending the basis yy to the other basis xx.

Let KK be a knot in S3S^{3} and MM be the knot exterior with a fixed triangulation. For an irreducible representation ρ:π1(M)SL2()\rho:\pi_{1}(M)\rightarrow\mathrm{SL}_{2}(\mathbb{C}) we consider the twisted cochain complex

C(M;𝔤ρ):=Hom[π1M](C(M~;),𝔤)C^{\ast}(M;\mathfrak{g}_{\rho}):=\mathrm{Hom}_{\mathbb{Z}[\pi_{1}M]}\left(C_{\ast}(\widetilde{M};\mathbb{Z}),\mathfrak{g}\right)

where M~\widetilde{M} is the universal cover of MM with the induced triangulation and 𝔤\mathfrak{g} is endowed with a [π1(M)]\mathbb{Z}[\pi_{1}(M)]-module structure via the adjoint representation of ρ\rho. We fix an orientation of each cell in MM and denote the cells of MM by c1,,cmc_{1},\cdots,c_{m}. Once we choose a lift c~j\widetilde{c}_{j} of each cjc_{j} to M~\widetilde{M} (1jm)(1\leq j\leq m), we denote by 𝐜\mathbf{c}^{\ast} a basis of C(M;𝔤ρ)C^{\ast}(M;\mathfrak{g}_{\rho}) given as

(4) 𝐜=(c11,c12,c13,,cj1,cj2,cj3,,cm1,cm2,cm3)\mathbf{c}^{\ast}=\left(c_{1}^{1},c_{1}^{2},c_{1}^{3},\cdots,c_{j}^{1},c_{j}^{2},c_{j}^{3},\cdots,c_{m}^{1},c_{m}^{2},c_{m}^{3}\right)

where cjic_{j}^{i} sends the cell c~j\widetilde{c}_{j} to eie_{i} and the other cells c~k\widetilde{c}_{k} (1kjm)(1\leq k\neq j\leq m) to 0.

We assume that ρ\rho is μ\mu-regular for a meridian μ\mu of KK, see [Por97] for the definition of μ\mu-regularity. It follows that the twisted cohomology group Hi(M;𝔤ρ)H^{i}(M;\mathfrak{g}_{\rho}) has dimension 1 for i=1,2i=1,2 and is trivial, otherwise. Furthermore, choosing a non-trivial element P𝔤P\in\mathfrak{g} invariant under the adjoint action of ρ(g)\rho(g) for all gπ1(M)g\in\pi_{1}(\partial M), we have isomorphisms Fi:Hi(M;𝔤ρ)F_{i}:H^{i}(M;\mathfrak{g}_{\rho})\rightarrow\mathbb{C} for i=1,2i=1,2 (see [Por97, Dub06] for details) defined by

F1(v)=P,v(μ~)𝔤,F2(v)=P,v(M~)𝔤F_{1}(v)=\langle P,v(\widetilde{\mu})\rangle_{\mathfrak{g}},\ F_{2}(v)=\langle P,v(\widetilde{\partial M})\rangle_{\mathfrak{g}}

where μ~C1(M~;)\widetilde{\mu}\in C_{1}(\widetilde{M};\mathbb{Z}) and M~C2(M~;)\widetilde{\partial M}\in C_{2}(\widetilde{M};\mathbb{Z}) represent lifts of μ\mu and M\partial M to M~\widetilde{M}, respectively, having the same base point. Here μ\mu and M\partial M are regarded as to be coherently oriented, see [Dub06] for details. Letting 𝐡i\mathbf{h}^{i} be a basis of Hi(M;𝔤ρ)H^{i}(M;\mathfrak{g}_{\rho}) satisfying Fi(𝐡i)=1F_{i}(\mathbf{h}^{i})=1 for i=1,2i=1,2, the adjoint Reidemeister torsion 𝕋μ(ρ)\mathbb{T}_{\mu}(\rho) is defined as

(5) 𝕋μ(ρ)=ϵtor(C(M;𝔤ρ),𝐜,𝐡).\mathbb{T}_{\mu}(\rho)=\epsilon\cdot\mathrm{tor}\left(C^{\ast}(M;\mathfrak{g}_{\rho}),\mathbf{c}^{\ast},\mathbf{h}^{\ast}\right)\in\mathbb{C}^{\ast}.

Here ϵ{±1}\epsilon\in\{\pm 1\} is the sign determined by so-called Turaev’s sign trick and the choice of homology orientation, see [Tur01] for details.

Remark 2.1.

It is known that the notion of μ\mu-regularity and the adjoint torsion are invariant under conjugation, so the adjoint Reidemeister torsion is also well-defined for μ\mu-regular characters. Recall that the character of an irreducible representation ρ\rho determines and is determined by the conjugacy class of ρ\rho, see [CS83, Proposition 1.5.2].

2.2. Infinitesimal deformation

We now fix a finite presentation of the knot group π1(M)\pi_{1}(M) of deficiency 11

π1(M)=g1,,gn|r1,,rn1\pi_{1}(M)=\langle g_{1},\cdots,g_{n}\,|\,r_{1},\cdots,r_{n-1}\rangle

and let YY be the corresponding 2-dimensional cell complex. Recall that YY has one 0-cell pp, nn 1-cells g1,,gng_{1},\cdots,g_{n}, and n1n-1 2-cells r1,,rn1r_{1},\cdots,r_{n-1}. Since we may use YY instead of the knot exterior MM to compute the adjoint torsion, we consider the twisted cochain complex of YY

0C0(Y;𝔤ρ)δ0C1(Y;𝔤ρ)δ1C2(Y;𝔤ρ)0.0\rightarrow C^{0}(Y;\mathfrak{g}_{\rho})\overset{\delta^{0}}{\longrightarrow}C^{1}(Y;\mathfrak{g}_{\rho})\overset{\delta^{1}}{\longrightarrow}C^{2}(Y;\mathfrak{g}_{\rho})\rightarrow 0.

Once we fix a lift of the base point pp to the universal cover of YY, each cell of YY admits a unique lift correspondingly. Denoting these lifts by using the usual symbol tilde p~\widetilde{p}, g~i\widetilde{g}_{i}, and r~j\widetilde{r}_{j}, it specifies the basis 𝐜\mathbf{c}^{\ast} of C(Y;𝔤ρ)C^{\ast}(Y;\mathfrak{g}_{\rho}) given as in the equation (4). With respect to the basis 𝐜\mathbf{c}^{\ast}, it is known that

δ0=(Φ(g11)Φ(gn1)),δ1=(Φ(r1g1)Φ(r1gn)Φ(rn1g1)Φ(rn1gn))\delta^{0}=\begin{pmatrix}\Phi(g_{1}-1)\\ \vdots\\ \Phi(g_{n}-1)\end{pmatrix},\quad\delta^{1}=\begin{pmatrix}\Phi(\frac{\partial r_{1}}{\partial g_{1}})&\cdots&\Phi(\frac{\partial r_{1}}{\partial g_{n}})\\ \vdots&\ddots&\vdots\\ \Phi(\frac{\partial r_{n-1}}{\partial g_{1}})&\cdots&\Phi(\frac{\partial r_{n-1}}{\partial g_{n}})\end{pmatrix}

where Φ:[π1(M)]M3,3()\Phi:\mathbb{Z}[\pi_{1}(M)]\rightarrow M_{3,3}(\mathbb{C}) is the \mathbb{Z}-linear extension of the adjoint representation of ρ\rho and rj/gi\partial r_{j}/\partial g_{i} is the Fox free differential. Here M3,3()M_{3,3}(\mathbb{C}) is the set of all 3-by-3 matrices. We refer to [Kit96, DHY09] for details.

Let FnF_{n} be the free group generated by g1,,gng_{1},\cdots,g_{n} and suppose that we have a one-parameter family of representations ρt:FnSL2()\rho_{t}:F_{n}\rightarrow\mathrm{SL}_{2}(\mathbb{C}) (parameterized by tt\in\mathbb{R}) such that ρt0(g1)=ρ(g1),,ρt0(gn)=ρ(gn)\rho_{t_{0}}(g_{1})=\rho(g_{1}),\cdots,\rho_{t_{0}}(g_{n})=\rho(g_{n}) for some t0t_{0}\in\mathbb{R}. Assuming that the entries of ρt(g1),,ρt(gn)\rho_{t}(g_{1}),\cdots,\rho_{t}(g_{n}) are differentiable at t=t0t=t_{0}, we define AρtC1(Y;𝔤ρ)A_{\rho_{t}}\in C^{1}(Y;\mathfrak{g}_{\rho}) by

(6) Aρt(g~i)=ddt|t=t0ρt(gi)ρt0(gi)1for 1in.A_{\rho_{t}}(\widetilde{g}_{i})=\left.\frac{d}{dt}\right|_{t=t_{0}}\rho_{t}(g_{i})\rho_{t_{0}}(g_{i})^{-1}\quad\textrm{for }1\leq i\leq n.

It follows from [Wei64] (see also [Sik12]) that δ1(Aρt)C2(Y;𝔤ρ)\delta^{1}(A_{\rho_{t}})\in C^{2}(Y;\mathfrak{g}_{\rho}) satisfies (and hence is determined by)

(7) (δ1Aρt)(r~j)=ddt|t=t0ρt(rj)for 1jn1.(\delta^{1}A_{\rho_{t}})(\widetilde{r}_{j})=\left.\frac{d}{dt}\right|_{t=t_{0}}\rho_{t}(r_{j})\quad\textrm{for }1\leq j\leq n-1.

Note that we have ρt0(r1)==ρt0(rn1)=I\rho_{t_{0}}(r_{1})=\cdots=\rho_{t_{0}}(r_{n-1})=I.

3. Adjoint Reidemeister torsion of two-bridge knots

Let KK be a two-bridge knot given by the Schubert normal form (p,q)(p,q). Here p>0p>0 and qq are relatively prime odd integers with p<q<p-p<q<p. We refer to [BZ03, Chapter 12] for details on two-bridge knot. Let MM be the knot exterior of KK. The knot group π1(M)\pi_{1}(M) of KK has a presentation

(8) π1(M)=g1,g2|wg1w1g21\pi_{1}(M)=\langle g_{1},g_{2}\,|\,wg_{1}w^{-1}g_{2}^{-1}\rangle

where w=g1ϵ1g2ϵ2g1ϵp2g2ϵp1w=g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{p-2}}g_{2}^{\epsilon_{p-1}} and ϵi=(1)iqp\epsilon_{i}=(-1)^{\lfloor\frac{iq}{p}\rfloor}. Here x\lfloor x\rfloor means the greatest integer less than or equal to xx\in\mathbb{R}. Up to conjugation, an irreducible representation ρ:π1(M)SL2()\rho:\pi_{1}(M)\rightarrow\mathrm{SL}_{2}(\mathbb{C}) is given by

(9) ρ(g1)=(m10m1),ρ(g2)=(m0um1)\rho(g_{1})=\begin{pmatrix}m&1\\ 0&m^{-1}\end{pmatrix},\ \rho(g_{2})=\begin{pmatrix}m&0\\ -u&m^{-1}\end{pmatrix}

for some m0m\neq 0 and u0u\neq 0 satisfying the Riley polynomial of KK, see the equation (10) below. We denote by χρ\chi_{\rho} the character of ρ\rho and let ϕg:=g11(mm1)g12\phi_{g}:=g_{11}-(m-m^{-1})g_{12} for gπ1(M)g\in\pi_{1}(M) where gij[m±1,u]g_{ij}\in\mathbb{Z}[m^{\pm 1},u] is the (i,j)(i,j)-entry of ρ(g)\rho(g). It is known that the Riley polynomial of KK is equal to

(10) ϕw=w11(mm1)w12[m±1,u]\phi_{w}=w_{11}-(m-m^{-1})w_{12}\in\mathbb{Z}[m^{\pm 1},u]

and the character variety XKX_{K} of irreducible SL2()\mathrm{SL}_{2}(\mathbb{C})-representations is given by

XK={(m,u)2:ϕw(m,u)=0,m0,u0}/X_{K}=\left\{(m,u)\in\mathbb{C}^{2}:\phi_{w}(m,u)=0,\ m\neq 0,\ u\neq 0\right\}/_{\sim}

where the quotient \sim means that we identify (m,u)(m,u) and (m1,u)(m^{-1},u). We refer to [Ril84] for details.

Theorem 3.1.

Suppose that χρ\chi_{\rho} is μ\mu-regular with m±1m\neq\pm 1, ϕwm0\frac{\partial\phi_{w}}{\partial m}\neq 0, and ϕwu0\frac{\partial\phi_{w}}{\partial u}\neq 0. Then we have

𝕋μ(χρ)=±mϵk+12(m21)w11v11ϕvϕwu\mathbb{T}_{\mu}(\chi_{\rho})=\pm\,\frac{m^{\epsilon_{k}+1}}{2(m^{2}-1)}\frac{w_{11}}{v_{11}\phi_{v}}\frac{\partial\phi_{w}}{\partial u}

where v=g1ϵ1g2ϵ2g1ϵk2g2ϵk1v=g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{k-2}}g_{2}^{\epsilon_{k-1}} and 0<k<p0<k<p is a unique (odd) integer satisfying kq±1(mod 2p)kq\equiv\pm 1\ (\mathrm{mod}\ 2p). Here the sign ±\pm does not depend on the choice of χρ\chi_{\rho}.

Remark 3.2.

The Riley polynomial ϕw\phi_{w} has no repeated factor, see [Ril84, Lemma 3]. It follows that the number of μ\mu-regular characters excluded in Theorem 3.1 is finite. In addition, the author does not know whether there exists an irreducible μ\mu-regular character with either ϕwm=0\frac{\partial\phi_{w}}{\partial m}=0 or ϕwu=0\frac{\partial\phi_{w}}{\partial u}=0.

Remark 3.3.

Othsuki and Takata [OT15] gave a diagrammatic formula for computing the adjoint torsion for a geometric representation of a hyperbolic two-bridge knot. Their formula can be directly applied to any irreducible μ\mu-regular character of a two-bridge knot with m=±1m=\pm 1.

Let trμ:XK\mathrm{tr}_{\mu}:X_{K}\rightarrow\mathbb{C} be the trace function of a meridian μ\mu, i.e., trμ(χρ)=trρ(μ)\mathrm{tr}_{\mu}(\chi_{\rho})=\mathrm{tr}\,\rho(\mu). It is known that non-μ\mu-regular characters in XKX_{K} are contained in the set of zeros of the differential of trμ:XK\mathrm{tr}_{\mu}:X_{K}\rightarrow\mathbb{C}, see [Por97, Proposition 3.26]. It follows that the set trμ1(c)\mathrm{tr}_{\mu}^{-1}(c) consists of μ\mu-regular characters for generic cc\in\mathbb{C}. As a consequence of Theorem 3.1, we prove the following.

Theorem 3.4.

Suppose that KK is a hyperbolic two-bridge knot. Then we have

χρtrμ1(c)1𝕋μ(χρ)=0\sum_{\chi_{\rho}\in\mathrm{tr}_{\mu}^{-1}(c)}\frac{1}{\mathbb{T}_{\mu}(\chi_{\rho})}=0

for generic cc\in\mathbb{C}.

Remark 3.5.

It is proved in [Sch54] that a two-bridge knot KK is hyperbolic if q±1q\neq\pm 1 and is a torus knot of the type (2,n)(2,n), otherwise. For a non-hyperbolic two-bridge knot (q=±1)(q=\pm 1) we have k=1k=1 and

𝕋μ(χρ)=mq+1w112(m21)ϕwu\mathbb{T}_{\mu}(\chi_{\rho})=\frac{m^{q+1}\,w_{11}}{2(m^{2}-1)}\frac{\partial\phi_{w}}{\partial u}

from Theorem 3.1. One checks that the inverse sum of adjoint torsions as in Theorem 3.4 is numerically 2q-2q for generic cc\in\mathbb{C}.

3.1. A proof of Theorem 3.1

Let YY be the 2-dimensional cell complex corresponding to the presentation (8) of the knot group π1(M)\pi_{1}(M). Recall that YY has one 0-cell pp, two 1-cells g1g_{1}, g2g_{2}, and one 2-cell rr and that we have the basis 𝐜\mathbf{c}^{\ast} of C(Y;𝔤ρ)C^{\ast}(Y;\mathfrak{g}_{\rho}) as in the equation (4): 𝐜0=(p1,p2,p3),𝐜1=(g11,g12,g13,g21,g22,g23),𝐜2=(r1,r2,r3).\mathbf{c}^{0}=(p^{1},p^{2},p^{3}),\ \mathbf{c}^{1}=(g_{1}^{1},g_{1}^{2},g_{1}^{3},g_{2}^{1},g_{2}^{2},g_{2}^{3}),\ \mathbf{c}^{2}=(r^{1},r^{2},r^{3}).

Let F2F_{2} be the free group generated by g1g_{1} and g2g_{2}. We consider a representation ρ¯:F2M2,2([t1±1,t2±1,t3])\overline{\rho}:F_{2}\rightarrow M_{2,2}(\mathbb{Z}[t_{1}^{\pm 1},t_{2}^{\pm 1},t_{3}]) given by

(11) ρ¯(g1)=(t110t11),ρ¯(g2)=(t20t3t21)\overline{\rho}(g_{1})=\begin{pmatrix}t_{1}&1\\ 0&t_{1}^{-1}\end{pmatrix},\ \overline{\rho}(g_{2})=\begin{pmatrix}t_{2}&0\\ -t_{3}&t_{2}^{-1}\end{pmatrix}

and define one-parameter families ρt0,ρt1,ρt2,ρt3:F2SL2()\rho^{0}_{t},\rho^{1}_{t},\rho^{2}_{t},\rho^{3}_{t}:F_{2}\rightarrow\mathrm{SL}_{2}(\mathbb{C}) from ρ¯\overline{\rho} by letting

ρt0:t1=t,t2=t,t3=u,ρt1:t1=t,t2=m,t3=u,ρt2:t1=m,t2=t,t3=u,ρt3:t1=m,t2=m,t3=t.\begin{array}[]{llll}\rho^{0}_{t}:&t_{1}=t,&t_{2}=t,&t_{3}=u,\\ \rho^{1}_{t}:&t_{1}=t,&t_{2}=m,&t_{3}=u,\\ \rho^{2}_{t}:&t_{1}=m,&t_{2}=t,&t_{3}=u,\\ \rho^{3}_{t}:&t_{1}=m,&t_{2}=m,&t_{3}=t.\end{array}

Clearly, ρti\rho^{i}_{t} coincides with ρ\rho at t=mt=m for i=0,1,2i=0,1,2 and at t=ut=u for i=3i=3. We thus have AρtiC1(Y;𝔤ρ)A_{\rho_{t}^{i}}\in C^{1}(Y;\mathfrak{g}_{\rho}) defined as in the equation (6) for i=0,1,2,3i=0,1,2,3. With respect to the basis 𝐜1\mathbf{c}^{1},

(12) Aρt0=(1m100m1um2),Aρt1=(1m10000),Aρt2=(0000m1um2),Aρt3=(00000m1).A_{\rho_{t}^{0}}=\begin{pmatrix}-1\\ m^{-1}\\ 0\\ 0\\ m^{-1}\\ -um^{-2}\end{pmatrix},\ A_{\rho_{t}^{1}}=\begin{pmatrix}-1\\ m^{-1}\\ 0\\ 0\\ 0\\ 0\end{pmatrix},\ A_{\rho_{t}^{2}}=\begin{pmatrix}0\\ 0\\ 0\\ 0\\ m^{-1}\\ -um^{-2}\end{pmatrix},\ A_{\rho_{t}^{3}}=\begin{pmatrix}0\\ 0\\ 0\\ 0\\ 0\\ -m^{-1}\end{pmatrix}.

Note that we have Aρt0=Aρt1+Aρt2A_{\rho^{0}_{t}}=A_{\rho^{1}_{t}}+A_{\rho^{2}_{t}} which is also clear from the definition.

Recall that the Riley polynomial is equal to ϕw=w11(mm1)w12\phi_{w}=w_{11}-(m-m^{-1})w_{12} where wij[m±1,u]w_{ij}\in\mathbb{Z}[m^{\pm 1},u] is the (i,j)(i,j)-entry of ρ(w)\rho(w). Similarly, we let ϕ¯w:=w¯11(t1t11)w¯12\overline{\phi}_{w}:=\overline{w}_{11}-(t_{1}-t_{1}^{-1})\overline{w}_{12} where w¯ij[t1±1,t2±1,t3]\overline{w}_{ij}\in\mathbb{Z}[t_{1}^{\pm 1},t_{2}^{\pm 1},t_{3}] is the (i,j)(i,j)-entry of ρ¯(w)\overline{\rho}(w).

Lemma 3.6.

We have

(13) (δ1Aρt0)(r~)\displaystyle(\delta^{1}A_{\rho^{0}_{t}})(\widetilde{r}) =ϕwmM,\displaystyle=\frac{\partial\phi_{w}}{\partial m}M,
(14) (δ1Aρt1)(r~)\displaystyle(\delta^{1}A_{\rho^{1}_{t}})(\widetilde{r}) =(m100m1)+α1M+(000),\displaystyle=\begin{pmatrix}m^{-1}&0\\ 0&-m^{-1}\end{pmatrix}+\alpha_{1}M+\begin{pmatrix}0&0\\ \ast&0\end{pmatrix},
(15) (δ1Aρt2)(r~)\displaystyle(\delta^{1}A_{\rho^{2}_{t}})(\widetilde{r}) =(m100m1)+α2M+(000),\displaystyle=\begin{pmatrix}-m^{-1}&0\\ 0&m^{-1}\end{pmatrix}+\alpha_{2}M+\begin{pmatrix}0&0\\ \ast&0\end{pmatrix},
(16) (δ1Aρt3)(r~)\displaystyle(\delta^{1}A_{\rho^{3}_{t}})(\widetilde{r}) =ϕwuM\displaystyle=\frac{\partial\phi_{w}}{\partial u}M

where

M=(w11uw21m1w11mw21u+w22um1w21m),α1=ϕ¯wt1|t1=t2=m,t3=u,α2=ϕ¯wt2|t1=t2=m,t3=u.M=\begin{pmatrix}w_{11}u-w_{21}m^{-1}&w_{11}m\\ w_{21}u+w_{22}um^{-1}&w_{21}m\end{pmatrix},\ \alpha_{1}=\left.\frac{\partial\overline{\phi}_{w}}{\partial t_{1}}\right|_{t_{1}=t_{2}=m,t_{3}=u},\ \alpha_{2}=\left.\frac{\partial\overline{\phi}_{w}}{\partial t_{2}}\right|_{t_{1}=t_{2}=m,t_{3}=u}.

Note that we have α1+α2=ϕwm\alpha_{1}+\alpha_{2}=\frac{\partial\phi_{w}}{\partial m} from the definition (also from the fact that Aρt0=Aρt1+Aρt2A_{\rho^{0}_{t}}=A_{\rho^{1}_{t}}+A_{\rho^{2}_{t}}).

Proof.

A straightforward computation shows that

ρ¯(r)\displaystyle\overline{\rho}(r) =ρ¯(wg1w1g21)\displaystyle=\overline{\rho}(wg_{1}w^{-1}g_{2}^{-1})
=(t1t2100t11t2)+ϕ¯w(w¯11t3w¯21t21w¯11t2w¯21t3+w¯22t3t21w¯21t2)\displaystyle=\begin{pmatrix}t_{1}t_{2}^{-1}&0\\ 0&t_{1}^{-1}t_{2}\end{pmatrix}+\overline{\phi}_{w}\begin{pmatrix}\overline{w}_{11}t_{3}-\overline{w}_{21}t_{2}^{-1}&\overline{w}_{11}t_{2}\\ \overline{w}_{21}t_{3}+\overline{w}_{22}t_{3}t_{2}^{-1}&\overline{w}_{21}t_{2}\end{pmatrix}
+(t11t21)(00t30)(w¯12t3+w¯21)(00w¯21(t1t11)w¯220).\displaystyle\quad\quad+\left(t_{1}^{-1}-t_{2}^{-1}\right)\begin{pmatrix}0&0\\ t_{3}&0\end{pmatrix}-(\overline{w}_{12}t_{3}+\overline{w}_{21})\begin{pmatrix}0&0\\ \overline{w}_{21}-(t_{1}-t_{1}^{-1})\overline{w}_{22}&0\end{pmatrix}.

Taking t1=t2=mt_{1}=t_{2}=m and t3=ut_{3}=u, the above equation reduces to the equation

(17) ρ(r)=ρ(wg1w1g21)=I+ϕw(w11uw21m1w11mw21u+w22um1w21m)=I+ϕwM.\rho(r)=\rho(wg_{1}w^{-1}g_{2}^{-1})=I+\phi_{w}\begin{pmatrix}w_{11}u-w_{21}m^{-1}&w_{11}m\\ w_{21}u+w_{22}um^{-1}&w_{21}m\end{pmatrix}=I+\phi_{w}M.

Note that the fact that w12u+w21=0w_{12}u+w_{21}=0 is used, see [Ril84]. We then obtain the equations (13)–(16) immediately from the equation (7). For instance,

(δ1Aρt0)(r~)\displaystyle(\delta^{1}A_{\rho^{0}_{t}})(\widetilde{r}) =ddt|t=mρt0(r)=ρ(r)m=ϕwmM+ϕwMm=ϕwmM,\displaystyle=\left.\frac{d}{dt}\right|_{t=m}\rho^{0}_{t}(r)=\frac{\partial\rho(r)}{\partial m}=\frac{\partial\phi_{w}}{\partial m}M+\phi_{w}\frac{\partial M}{\partial m}=\frac{\partial\phi_{w}}{\partial m}M,
(δ1Aρt1)(r~)\displaystyle(\delta^{1}A_{\rho^{1}_{t}})(\widetilde{r}) =ddt|t=mρt1(r)=ρ¯(r)t1|t1=t2=m,t3=u\displaystyle=\left.\frac{d}{dt}\right|_{t=m}\rho^{1}_{t}(r)=\left.\frac{\partial\overline{\rho}(r)}{\partial t_{1}}\right|_{t_{1}=t_{2}=m,t_{3}=u}
=(m100m1)+ϕ¯wt1|t1=t2=m,t3=uM+(000).\displaystyle=\begin{pmatrix}m^{-1}&0\\ 0&-m^{-1}\end{pmatrix}+\left.\frac{\partial\overline{\phi}_{w}}{\partial t_{1}}\right|_{t_{1}=t_{2}=m,t_{3}=u}M+\begin{pmatrix}0&0\\ \ast&0\end{pmatrix}.

We compute (δ1Aρt2)(r~)(\delta^{1}A_{\rho^{2}_{t}})(\widetilde{r}) and (δ1Aρt3)(r~)(\delta^{1}A_{\rho^{3}_{t}})(\widetilde{r}), similarly.

Claim 1.

The following vector BC1(Y;𝔤ρ)B\in C^{1}(Y;\mathfrak{g}_{\rho}) represents a generator of H1(Y;𝔤ρ)H^{1}(Y;\mathfrak{g}_{\rho}).

(18) B=Aρt0ϕw/mϕw/uAρt3=Aρt1+Aρt2ϕw/mϕw/uAρt3.B=A_{\rho^{0}_{t}}-\frac{\partial\phi_{w}/\partial m}{\partial\phi_{w}/\partial u}A_{\rho^{3}_{t}}=A_{\rho^{1}_{t}}+A_{\rho^{2}_{t}}-\frac{\partial\phi_{w}/\partial m}{\partial\phi_{w}/\partial u}A_{\rho^{3}_{t}}.
Proof.

Lemma 3.6 immediately implies that BKerδ1B\in\mathrm{Ker}\,\delta^{1}. The fact that BImδ0B\notin\mathrm{Im}\,\delta^{0} can be easily checked from the explicit expressions of AρtiA_{\rho^{i}_{t}} (see the equation (12)) and δ0\delta^{0} with respect to the basis 𝐜\mathbf{c}^{\ast}

(19) δ0=(m212m100m100m21m2100um00u22um1m21).\delta^{0}=\begin{pmatrix}m^{2}-1&-2m&-1\\ 0&0&m^{-1}\\ 0&0&m^{-2}-1\\ m^{2}-1&0&0\\ um&0&0\\ -u^{2}&-2um^{-1}&m^{-2}-1\end{pmatrix}.

On the zero set of the Riley polynomial ϕw=w11(mm1)w12\phi_{w}=w_{11}-(m-m^{-1})w_{12} with m±1m\neq\pm 1, we have w110w_{11}\neq 0 and w120w_{12}\neq 0. Otherwise, we have w11=w12=0w_{11}=w_{12}=0 which contradicts to detρ(w)=1\det\rho(w)=1.

Recall Section 2 that to compute the adjoint torsion we need to specify a sequence bib^{i} of vectors in Ci(Y;𝔤ρ)C^{i}(Y;\mathfrak{g}_{\rho}) such that δi(bi)\delta^{i}(b^{i}) is a basis of Imδi\mathrm{Im}\ \delta^{i} for i=0,1i=0,1 and a basis 𝐡i\mathbf{h}^{i} of Hi(Y;𝔤ρ)H^{i}(Y;\mathfrak{g}_{\rho}) satisfying Fi(𝐡i)=1F_{i}(\mathbf{h}^{i})=1 for i=1,2i=1,2.

Claim 2.

The following choice of bib^{i} satisfies that δi(bi)\delta^{i}(b^{i}) is a basis of Imδi\mathrm{Im}\,\delta^{i}.

(20) b0=𝐜0,b1=(β1Aρt1+β2Aρt2,Aρt3)b^{0}=\mathbf{c}^{0},\quad b^{1}=(\beta_{1}A_{\rho_{t}^{1}}+\beta_{2}A_{\rho_{t}^{2}},\ A_{\rho_{t}^{3}})

for any β1β2\beta_{1}\neq\beta_{2}\in\mathbb{C}.

Proof.

It is clear that δ0(b0)\delta^{0}(b^{0}) is a basis of Imδ0\mathrm{Im}\,\delta^{0}, since δ0\delta^{0} is injective.

Since dimH2(Y;𝔤ρ)=1\dim_{\mathbb{C}}H^{2}(Y;\mathfrak{g}_{\rho})=1, we have dimImδ1=2\dim_{\mathbb{C}}\mathrm{Im}\,\delta^{1}=2. Suppose δ1(b1)\delta^{1}(b^{1}) fails to span Imδ1\mathrm{Im}\,\delta^{1}. Then there is a linear combination CC of β1Aρt1+β2Aρt2\beta_{1}A_{\rho_{t}^{1}}+\beta_{2}A_{\rho_{t}^{2}} and Aρt3A_{\rho_{t}^{3}} such that δ1(C)=0\delta^{1}(C)=0. Recall that w110w_{11}\neq 0, ϕwu0\frac{\partial\phi_{w}}{\partial u}\neq 0, and

(21) (δ1Aρt3)(r~)=ϕwu(w11uw21m1w11mw21u+w22um1w21m).(\delta^{1}A_{\rho_{t}^{3}})(\widetilde{r})=\dfrac{\partial\phi_{w}}{\partial u}\begin{pmatrix}w_{11}u-w_{21}m^{-1}&w_{11}m\\ w_{21}u+w_{22}um^{-1}&w_{21}m\end{pmatrix}.

It follows that δ1(Aρt3)0\delta^{1}(A_{\rho_{t}^{3}})\neq 0 and thus CC has a non-trivial coefficient for β1Aρt1+β2Aρt2\beta_{1}A_{\rho^{1}_{t}}+\beta_{2}A_{\rho^{2}_{t}}. Since Aρt1,Aρt2,A_{\rho_{t}^{1}},A_{\rho_{t}^{2}}, and Aρt3A_{\rho_{t}^{3}} are linearly independent (see the equation (12)), so are the vector BB in Claim 1 and CC. Since BB generates H1(Y;𝔤ρ)H^{1}(Y;\mathfrak{g}_{\rho}), we have CkBImδ0C-kB\in\mathrm{Im}\,\delta^{0} for some kk\in\mathbb{C}. It follows that there exists a1Aρt1+a2Aρt2+a3Aρt3Imδ0a_{1}A_{\rho^{1}_{t}}+a_{2}A_{\rho^{2}_{t}}+a_{3}A_{\rho^{3}_{t}}\in\mathrm{Im}\,\delta^{0} for some a1,a2,a3a_{1},a_{2},a_{3}\in\mathbb{C} with a1a2a_{1}\neq a_{2}. However, one easily checks that such a vector can not exist from the explicit expressions (12) and (19). ∎

We fix a meridian μ\mu by g1g_{1} (see Figure 1 (left)) and choose P0𝔤P\neq 0\in\mathfrak{g} invariant under the adjoint action for all peripheral curves as

P=(12(mm1)1012(mm1)).P=\begin{pmatrix}\frac{1}{2}(m-m^{-1})&1\\ 0&-\frac{1}{2}(m-m^{-1})\end{pmatrix}.
Claim 3.

The following choice of 𝐡~1C1(Y;𝔤ρ)\widetilde{\mathbf{h}}^{1}\in C^{1}(Y;\mathfrak{g}_{\rho}) represents the basis 𝐡1\mathbf{h}^{1} of H1(Y;𝔤ρ)H^{1}(Y;\mathfrak{g}_{\rho}) satisfying F1(𝐡1)=1F_{1}(\mathbf{h}^{1})=1.

𝐡~1=14(1m2)(Aρt0ϕw/mϕw/uAρt3)\widetilde{\mathbf{h}}^{1}=\dfrac{1}{4(1-m^{-2})}\left(A_{\rho^{0}_{t}}-\frac{\partial\phi_{w}/\partial m}{\partial\phi_{w}/\partial u}A_{\rho^{3}_{t}}\right)
Proof.

The proof is immediately followed from Claim 1 with

P,B(g~1)=(12(mm1)1012(mm1)),(m110m1)𝔤=4(1m2).\langle P,B(\widetilde{g}_{1})\rangle=\left\langle\begin{pmatrix}\frac{1}{2}(m-m^{-1})&1\\ 0&-\frac{1}{2}(m-m^{-1})\end{pmatrix},\begin{pmatrix}m^{-1}&-1\\ 0&-m^{-1}\end{pmatrix}\right\rangle_{\mathfrak{g}}=4(1-m^{-2}).

In the Schubert normal form, the relator r=wg1w1g21r=wg_{1}w^{-1}g_{2}^{-1} is represented by a loop that travels along the knot diagram as in Figure 1 (right).

Refer to caption
g1g_{1}
g2g_{2}
Figure 1. The generators g1,g2g_{1},g_{2} and the loop representing the relator rr for (p,q)=(5,3)(p,q)=(5,3).

More precisely, if we follow the loop from the square-dot as in Figure 1 (right), we obtain the word wg1w1g21wg_{1}w^{-1}g_{2}^{-1}. Similarly, if we start from the circle-dot as in Figure 1 (right), we obtain the word g1wg21(w)1g_{1}w^{\dagger}g_{2}^{-1}(w^{\dagger})^{-1} along the loop, where w=g2ϵ1g1ϵ2g1ϵp1w^{\dagger}=g_{2}^{\epsilon_{1}}g_{1}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{p-1}} is the word obtained from ww by exchanging g1g_{1} and g2g_{2}. Considering the base-point change from the square-dot to the circle-dot, we obtain

(22) v1(wg1w1g21)v=g1wg21(w)1where v={g1ϵ1g2ϵ2g1ϵk2g2ϵk1ifϵk=1g2g1ϵ1g2ϵ2g2ϵk1g1ϵkifϵk=1v^{\prime-1}(wg_{1}w^{-1}g_{2}^{-1})v^{\prime}=g_{1}w^{\dagger}g_{2}^{-1}(w^{\dagger})^{-1}\quad\textrm{where }v^{\prime}=\left\{\begin{array}[]{ll}g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{k-2}}g_{2}^{\epsilon_{k-1}}&\mathrm{if}\ \epsilon_{k}=1\\ g_{2}g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{2}^{\epsilon_{k-1}}g_{1}^{\epsilon_{k}}&\mathrm{if}\ \epsilon_{k}=-1\end{array}\right.

Here 0<k<p0<k<p is a unique odd integer satisfying kq±1kq\equiv\pm 1 in modulo 2p2p. Since the loop www^{\dagger}w is a longitude of KK (see [Ril72, Ril84]), the boundary torus M\partial M of KK is represented by v1r~+wr~C2(Y~;)-v^{\prime-1}\cdot\widetilde{r}+w^{\dagger}\cdot\widetilde{r}\in C_{2}(\widetilde{Y};\mathbb{Z}) as in Figure 2.

Refer to caption
ww^{\dagger}
ww
g1g_{1}
ww
ww^{\dagger}
g2g_{2}
g1g_{1}
Refer to caption
wr~w^{\dagger}\cdot\widetilde{r}
v1r~-v^{\prime-1}\cdot\widetilde{r}
Figure 2. The boundary torus M\partial M.
Claim 4.

The following choice of 𝐡~2C2(Y;𝔤ρ)\widetilde{\mathbf{h}}^{2}\in C^{2}(Y;\mathfrak{g}_{\rho}) represents the basis 𝐡2\mathbf{h}^{2} of H2(Y;𝔤ρ)H^{2}(Y;\mathfrak{g}_{\rho}) satisfying F2(𝐡2)=1F_{2}(\mathbf{h}^{2})=1.

(23) 𝐡~2=(0014v11ϕv)T\widetilde{\mathbf{h}}^{2}=\begin{pmatrix}0&0&-\dfrac{1}{4v^{\prime}_{11}\phi_{v^{\prime}}}\end{pmatrix}^{T}

Here we write an element of 𝔤\mathfrak{g} as an element of 3\mathbb{C}^{3} with respect to the basis (2).

Proof.

Recall Lemma 3.6 that α1+α2=ϕw/m0\alpha_{1}+\alpha_{2}=\partial\phi_{w}/\partial m\neq 0 and thus both α1\alpha_{1} and α2\alpha_{2} can not be zero. Without loss of generality, we assume α10\alpha_{1}\neq 0. Considering the e1e_{1}- and e2e_{2}-coordinates of (δ1Aρt0)(r~)(\delta^{1}A_{\rho_{t}^{0}})(\widetilde{r}) and (δ1Aρt1)(r~)(\delta^{1}A_{\rho_{t}^{1}})(\widetilde{r}) given in Lemma 3.6, it follows that any element ηH2(Y;𝔤ρ)\eta\in H^{2}(Y;\mathfrak{g}_{\rho}) has a representative η~C2(Y;𝔤ρ)\widetilde{\eta}\in C^{2}(Y;\mathfrak{g}_{\rho}) such that η~(r~)=(0,0,c)T\widetilde{\eta}(\widetilde{r})=(0,0,c)^{T} for some cc\in\mathbb{C}. On the other hand, one can easily check that any vector C𝔤C\in\mathfrak{g} satisfying P,C𝔤=1\langle P,C\rangle_{\mathfrak{g}}=1 is of the form

(24) C=(0,0,1/4)T+a(1,0,0)T+b(0,1,mm1)TC=(0,0,1/4)^{T}+a(1,0,0)^{T}+b(0,\,-1,\,m-m^{-1})^{T}

for some a,ba,b\in\mathbb{C}.

Recall that we showed (cf. Figure 2) that an element η~C2(Y;𝔤ρ)\widetilde{\eta}\in C^{2}(Y;\mathfrak{g}_{\rho}) with η~(r~)=(0,0,c)T\widetilde{\eta}(\widetilde{r})=(0,0,c)^{T} satisfies

η~(M~)=(Adρ(v1)+Adρ(w))(0,0,c)T.\widetilde{\eta}(\widetilde{\partial M})=\left(-\mathrm{Ad}_{\rho(v^{\prime-1})}+\mathrm{Ad}_{\rho(w^{\dagger})}\right)(0,0,c)^{T}.

A simple computation shows that η~(M~)\widetilde{\eta}(\widetilde{\partial M}) in the above equation has the form (24) if and only if

c=14(w22(w22+(mm1)w12)v11(v11(mm1)v12)).c=\dfrac{1}{4(w_{22}^{\dagger}(w_{22}^{\dagger}+(m-m^{-1})w_{12}^{\dagger})-v^{\prime}_{11}(v^{\prime}_{11}-(m-m^{-1})v^{\prime}_{12}))}.

This completes the proof, since w22+(mm1)w12w_{22}^{\dagger}+(m-m^{-1})w_{12}^{\dagger} coincides with the Riley polynomial ϕw\phi_{w}, see [Ril84]. ∎

We now compute the determinant of the transition matrices. From Claim 2 and 3, we have [b0/𝐜0]=1[b^{0}/\mathbf{c}^{0}]=1 and

[(δ0(b0),𝐡~1,b1)𝐜1]\displaystyle\left[\dfrac{(\delta^{0}(b^{0}),\widetilde{\mathbf{h}}^{1},b^{1})}{\mathbf{c}^{1}}\right] =14(1m2)[(δ0(𝐜0),Aρt0,β1Aρt1+β2Aρt2,Aρt3)𝐜1]\displaystyle=\frac{1}{4(1-m^{-2})}\left[\dfrac{(\delta^{0}(\mathbf{c}^{0}),\,A_{\rho^{0}_{t}},\,\beta_{1}A_{\rho^{1}_{t}}+\beta_{2}A_{\rho^{2}_{t}},\,A_{\rho_{t}^{3}})}{\mathbf{c}^{1}}\right]
=14(1m2)det(m212m11β1000m1m1β1m1000m21000m2100000um00m1β2m10u22um1m21um2β2um2m1)\displaystyle=\frac{1}{4(1-m^{-2})}\,\det\begin{pmatrix}m^{2}-1&-2m&-1&-1&-\beta_{1}&0\\ 0&0&m^{-1}&m^{-1}&\beta_{1}m^{-1}&0\\ 0&0&m^{-2}-1&0&0&0\\ m^{2}-1&0&0&0&0&0\\ um&0&0&m^{-1}&\beta_{2}m^{-1}&0\\ -u^{2}&-2um^{-1}&m^{-2}-1&-um^{-2}&-\beta_{2}um^{-2}&-m^{-1}\end{pmatrix}
=(1m2)(β1β2)2.\displaystyle=\frac{(1-m^{-2})(\beta_{1}-\beta_{2})}{2}.

We choose the constants β1β2\beta_{1}\neq\beta_{2} such that α1β1+α2β2=0\alpha_{1}\beta_{1}+\alpha_{2}\beta_{2}=0. Note that the fact α1+α2=ϕwm0\alpha_{1}+\alpha_{2}=\frac{\partial\phi_{w}}{\partial m}\neq 0 implies that β1β2\beta_{1}\neq\beta_{2}. Then Lemma 3.6 gives that

(δ1(β1Aρt1+β2Aρt2))(r~)=(β1β2m0β1β2m)(\delta^{1}(\beta_{1}A_{\rho_{t}^{1}}+\beta_{2}A_{\rho_{t}^{2}}))(\widetilde{r})=\begin{pmatrix}\frac{\beta_{1}-\beta_{2}}{m}&0\\ \ast&-\frac{\beta_{1}-\beta_{2}}{m}\end{pmatrix}

and from Claims 2 and 4 we obtain

[δ1(b1),𝐡~2𝐜2]\displaystyle\left[\frac{\delta^{1}(b^{1}),\widetilde{\mathbf{h}}^{2}}{\mathbf{c}^{2}}\right] =[δ1(β1Aρt1+β2Aρt2),δ1(Aρt3),𝐡~2𝐜2]\displaystyle=\left[\frac{\delta^{1}(\beta_{1}A_{\rho_{t}^{1}}+\beta_{2}A_{\rho_{t}^{2}}),\,\delta^{1}(A_{\rho_{t}^{3}}),\,\widetilde{\mathbf{h}}^{2}}{\mathbf{c}^{2}}\right]
=det(0ϕwuw11m0β1β2m014v11ϕv)\displaystyle=\det\begin{pmatrix}0&\frac{\partial\phi_{w}}{\partial u}w_{11}m&0\\ \frac{\beta_{1}-\beta_{2}}{m}&\ast&0\\ \ast&\ast&-\frac{1}{4v^{\prime}_{11}\phi_{v^{\prime}}}\end{pmatrix}
=(β1β2)w114v11ϕvϕwu,\displaystyle=\frac{(\beta_{1}-\beta_{2})w_{11}}{4v^{\prime}_{11}\phi_{v^{\prime}}}\frac{\partial\phi_{w}}{\partial u},

where the second column is obtained from the equation (21). Since one easily checks that v11ϕv=v11ϕvv^{\prime}_{11}\phi_{v^{\prime}}=v_{11}\phi_{v} if ϵk=1\epsilon_{k}=1 and v11ϕv=m2v11ϕvv^{\prime}_{11}\phi_{v^{\prime}}=m^{2}v_{11}\phi_{v} if ϵk=1\epsilon_{k}=-1 from the equation (22), we obtain Theorem 3.1

(25) 𝕋μ(χρ)=±[b0𝐜0][δ0(b0),𝐡~1,b1𝐜1]1[δ1(b1),𝐡~2𝐜2]=±w112(1m2)v11ϕvϕwu=±mϵk+1w112(m21)v11ϕvϕwu.\mathbb{T}_{\mu}(\chi_{\rho})=\pm\left[\frac{b^{0}}{\mathbf{c}^{0}}\right]\left[\dfrac{\delta^{0}(b^{0}),\widetilde{\mathbf{h}}^{1},b^{1}}{\mathbf{c}^{1}}\right]^{-1}\left[\frac{\delta^{1}(b^{1}),\widetilde{\mathbf{h}}^{2}}{\mathbf{c}^{2}}\right]=\pm\frac{w_{11}}{2(1-m^{-2})v^{\prime}_{11}\phi_{v^{\prime}}}\frac{\partial\phi_{w}}{\partial u}=\pm\frac{m^{\epsilon_{k}+1}w_{11}}{2(m^{2}-1)v_{11}\phi_{v}}\frac{\partial\phi_{w}}{\partial u}.

Since changing the character χρ\chi_{\rho}, equivalently, changing the pair (m,u)(m,u), does not change any combinatorial data of the computation, the sign ±\pm in the equation (25) does not depend on the choice of χρ\chi_{\rho} (see [Tur01] for details).

3.2. A proof of Theorem 3.4

Recall that the character variety XKX_{K} of a two-bridge knot KK is given by

XK={(m,u)2:ϕw(m,u)=0,m0,u0}/X_{K}=\{(m,u)\in\mathbb{C}^{2}:\phi_{w}(m,u)=0,\ m\neq 0,\ u\neq 0\}/_{\sim}

where the quotient \sim means that we identify (m,u)(m,u) and (m1,u)(m^{-1},u). Since the trace function of μ\mu is simply m+m1m+m^{-1}, Theorem 3.1 implies that for generic cc\in\mathbb{C}

(26) χρtrμ1(c)1𝕋μ(χρ)=±dϵk+12(d21)ϕw(d,u)=0v11(d,u)ϕv(d,u)w11(d,u)ϕwu(d,u)\sum_{\chi_{\rho}\in\mathrm{tr}_{\mu}^{-1}(c)}\frac{1}{\mathbb{T}_{\mu}(\chi_{\rho})}=\pm\frac{d^{\epsilon_{k}+1}}{2(d^{2}-1)}\sum_{\phi_{w}(d,u)=0}\frac{v_{11}(d,u)\phi_{v}(d,u)}{w_{11}(d,u)\frac{\partial\phi_{w}}{\partial u}(d,u)}

where dd\in\mathbb{C}^{\ast} satisfies d+d1=cd+d^{-1}=c. Hereafter we regard every element in [m±1,u]\mathbb{C}[m^{\pm 1},u] as an element of [u]\mathbb{C}[u] by letting m=dm=d. For instance, ϕw=ϕw(d,u)[u]\phi_{w}=\phi_{w}(d,u)\in\mathbb{C}[u]; in particular, ϕw\phi^{\prime}_{w} means the derivative of ϕw\phi_{w} with respect to the variable uu.

Recall that we have w=g1ϵ1g2ϵ2g1ϵp2g2ϵp1w=g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{p-2}}g_{2}^{\epsilon_{p-1}} and v=g1ϵ1g2ϵ2g1ϵk2g2ϵk1v=g_{1}^{\epsilon_{1}}g_{2}^{\epsilon_{2}}\cdots g_{1}^{\epsilon_{k-2}}g_{2}^{\epsilon_{k-1}} where 0<k<p0<k<p is a unique odd integer satisfying kq±1kq\equiv\pm 1 in modulo 2p2p. Let y=g1ϵkg2ϵk+1g1ϵp2g2ϵp1y=g_{1}^{\epsilon_{k}}g_{2}^{\epsilon_{k}+1}\cdots g_{1}^{\epsilon_{p-2}}g_{2}^{\epsilon_{p-1}} so that w=vyw=vy. From v=wy1v=wy^{-1} we have v11=w11y22w12y21v_{11}=w_{11}y_{22}-w_{12}y_{21} and thus v11=w11(y21(dd1)y22)/(dd1)v_{11}=-w_{11}(y_{21}-(d-d^{-1})y_{22})/(d-d^{-1}) on the zero set of the Riley polynomial ϕw=w11(dd1)w12\phi_{w}=w_{11}-(d-d^{-1})w_{12}. Therefore,

(27) ϕw=0v11ϕvw11ϕw=1dd1ϕw=0(y21(dd1)y22)ϕvϕw.\sum_{\phi_{w}=0}\frac{v_{11}\phi_{v}}{w_{11}\phi^{\prime}_{w}}=-\frac{1}{d-d^{-1}}\sum_{\phi_{w}=0}\frac{(y_{21}-(d-d^{-1})y_{22})\phi_{v}}{\phi_{w}^{\prime}}.

We will claim that (y21(dd1)y22)ϕv=αϕw+h(y_{21}-(d-d^{-1})y_{22})\phi_{v}=\alpha\phi_{w}+h for some α\alpha\in\mathbb{C} and h[u]h\in\mathbb{C}[u] with deghdegϕw2\deg h\leq\deg\phi_{w}-2. Then we obtain Theorem 3.4 by applying the Euler-Jacobi theorem, saying that for any ff and g[u]g\in\mathbb{C}[u]

f=0g(u)f(u)=0\sum_{f=0}\frac{g(u)}{f^{\prime}(u)}=0

whenever ff has a non-zero constant term with no double root and gg satisfies deggdegf2\deg g\leq\deg f-2. We refer to [GH78, Chapter 5] for details. Note that the Riley polynomial ϕw\phi_{w} has a non-zero constant term with no double root for generic dd\in\mathbb{C}^{\ast}, see [Ril84, Lemma 3].

For ff and g[u]g\in\mathbb{C}[u] we write f=g+o(un)f=g+o(u^{n}) if fgf-g has degree less than n0n\geq 0; f=g+o(u0)f=g+o(u^{0}) means f=gf=g.

Lemma 3.7.

Let (f1,,f2n)(f_{1},\cdots,f_{2n}) be a sequence of fj{±1}f_{j}\in\{\pm 1\} of even length 2n>02n>0 and let hj:=g1f1g2f2g1f2j1g2f2jh_{j}:=g_{1}^{f_{1}}g_{2}^{f_{2}}\cdots g_{1}^{f_{2j-1}}g_{2}^{f_{2j}} for 1jn1\leq j\leq n. Then hjh_{j} satisfies

(28) ρ(hj)=±(uj+Ajuj1+o(uj1)f2jdf2j(uj1+o(uj1))f1df1(uj+Bjuj1+o(uj1))f1f2jdf1f2j(uj1+o(uj1)))\rho(h_{j})=\pm\begin{pmatrix}u^{j}+A_{j}u^{j-1}+o(u^{j-1})&\ -f_{2j}d^{-f_{2j}}\left(u^{j-1}+o(u^{j-1})\right)\\[3.0pt] f_{1}d^{-f_{1}}\left(u^{j}+B_{j}u^{j-1}+o(u^{j-1})\right)&\ -f_{1}f_{2j}d^{-f_{1}-f_{2j}}\left(u^{j-1}+o(u^{j-1})\right)\end{pmatrix}

for some AjA_{j} and BjB_{j}\in\mathbb{C} where the value AjBjA_{j}-B_{j} does not depend on jj.

Proof.

Recall that we have

ρ(g1)=(d10d1),ρ(g2)=(d0ud1).\rho(g_{1})=\begin{pmatrix}d&1\\ 0&d^{-1}\end{pmatrix},\ \rho(g_{2})=\begin{pmatrix}d&0\\ -u&d^{-1}\end{pmatrix}.

A simple computation shows that

ρ(g1fig2fj)=±(ufifjdfi+fjfjdfjfidfiufifjdfifj)\rho\left(g_{1}^{f_{i}}g_{2}^{f_{j}}\right)=\pm\begin{pmatrix}u-f_{i}f_{j}d^{f_{i}+f_{j}}&-f_{j}d^{-f_{j}}\\[3.0pt] f_{i}d^{-f_{i}}u&-f_{i}f_{j}d^{-f_{i}-f_{j}}\end{pmatrix}

for all fi,fj{±1}f_{i},f_{j}\in\{\pm 1\}. With the above equation, a routine induction proves that ρ(hj)\rho(h_{j}) has the form (28) with relations

Aj+1\displaystyle A_{j+1} =Ajf2jf2j+1df2jf2j+1f2j+1f2j+2df2j+1+f2j+2,\displaystyle=A_{j}-f_{2j}f_{2j+1}d^{-f_{2j}-f_{2j+1}}-f_{2j+1}f_{2j+2}d^{f_{2j+1}+f_{2j+2}},
Bj+1\displaystyle B_{j+1} =Bjf2jf2j+1df2jf2j+1f2j+1f2j+2df2j+1+f2j+2.\displaystyle=B_{j}-f_{2j}f_{2j+1}d^{-f_{2j}-f_{2j+1}}-f_{2j+1}f_{2j+2}d^{f_{2j+1}+f_{2j+2}}.

In particular, we have Aj+1Bj+1=AjBjA_{j+1}-B_{j+1}=A_{j}-B_{j}. ∎

From Lemma 3.7 we have

(29) ρ(v)\displaystyle\rho(v) =±(uk12+V11uk32+o(uk32)ϵk1dϵk1(uk32+o(uk32))ϵ1dϵ1(uk12+V21uk32+o(uk32))ϵ1ϵk1dϵ1ϵk1(uk32+o(uk32)))\displaystyle=\pm\begin{pmatrix}u^{\frac{k-1}{2}}+V_{11}u^{\frac{k-3}{2}}+o(u^{\frac{k-3}{2}})&\ -\epsilon_{k-1}d^{-\epsilon_{k-1}}\left(u^{\frac{k-3}{2}}+o(u^{\frac{k-3}{2}})\right)\\[3.0pt] \epsilon_{1}d^{-\epsilon_{1}}\left(u^{\frac{k-1}{2}}+V_{21}u^{\frac{k-3}{2}}+o(u^{\frac{k-3}{2}})\right)&\ -\epsilon_{1}\epsilon_{k-1}d^{-\epsilon_{1}-\epsilon_{k-1}}\left(u^{\frac{k-3}{2}}+o(u^{\frac{k-3}{2}})\right)\end{pmatrix}
(30) ρ(y)\displaystyle\rho(y) =±(upk2+Y11upk22+o(upk22)ϵp1dϵp1(upk22+o(upk22))ϵkdϵk(upk2+Y21upk22+o(upk22))ϵkϵp1dϵkϵp1(upk22+o(upk22)))\displaystyle=\pm\begin{pmatrix}u^{\frac{p-k}{2}}+Y_{11}u^{\frac{p-k-2}{2}}+o(u^{\frac{p-k-2}{2}})&\ -\epsilon_{p-1}d^{-\epsilon_{p-1}}\left(u^{\frac{p-k-2}{2}}+o(u^{\frac{p-k-2}{2}})\right)\\[3.0pt] \epsilon_{k}d^{-\epsilon_{k}}\left(u^{\frac{p-k}{2}}+Y_{21}u^{\frac{p-k-2}{2}}+o(u^{\frac{p-k-2}{2}})\right)&\ -\epsilon_{k}\epsilon_{p-1}d^{-\epsilon_{k}-\epsilon_{p-1}}\left(u^{\frac{p-k-2}{2}}+o(u^{\frac{p-k-2}{2}})\right)\end{pmatrix}
(31) ρ(w)\displaystyle\rho(w) =±(up12+W11up32+o(up32)ϵp1dϵp1(up32+o(up32))ϵ1dϵ1(up12+W21up32+o(up32))ϵ1ϵp1dϵ1ϵp1(up32+o(up32)))\displaystyle=\pm\begin{pmatrix}u^{\frac{p-1}{2}}+W_{11}u^{\frac{p-3}{2}}+o(u^{\frac{p-3}{2}})&\ -\epsilon_{p-1}d^{-\epsilon_{p-1}}\left(u^{\frac{p-3}{2}}+o(u^{\frac{p-3}{2}})\right)\\[3.0pt] \epsilon_{1}d^{-\epsilon_{1}}\left(u^{\frac{p-1}{2}}+W_{21}u^{\frac{p-3}{2}}+o(u^{\frac{p-3}{2}})\right)&\ -\epsilon_{1}\epsilon_{p-1}d^{-\epsilon_{1}-\epsilon_{p-1}}\left(u^{\frac{p-3}{2}}+o(u^{\frac{p-3}{2}})\right)\end{pmatrix}

for some Vij,Yij,WijV_{ij},Y_{ij},W_{ij}\in\mathbb{C}. Note that k1k\neq 1 (and thus k3k\geq 3) since q±1q\neq\pm 1 and that

(32) ϕw=±(up12+(W11+ϵp1dϵp1(dd1))up32+o(up32)).\phi_{w}=\pm\left(u^{\frac{p-1}{2}}+(W_{11}+\epsilon_{p-1}d^{-\epsilon_{p-1}}(d-d^{-1}))u^{\frac{p-3}{2}}+o(u^{\frac{p-3}{2}})\right).

Computing the first and second leading terms of (y21(dd1)y22)ϕv(y_{21}-(d-d^{-1})y_{22})\phi_{v},

(y21(dd1)y22)ϕv\displaystyle(y_{21}-(d-d^{-1})y_{22})\phi_{v}
=±ϵkdϵk(upk2+(Y21+ϵp1dϵp1(dd1))upk22+o(upk22))\displaystyle=\pm\epsilon_{k}d^{-\epsilon_{k}}\left(u^{\frac{p-k}{2}}+(Y_{21}+\epsilon_{p-1}d^{-\epsilon_{p-1}}(d-d^{-1}))u^{\frac{p-k-2}{2}}+o(u^{\frac{p-k-2}{2}})\right)
(uk12+(V11+ϵk1dϵk1(dd1))uk32+o(uk32))\displaystyle\quad\cdot\left(u^{\frac{k-1}{2}}+(V_{11}+\epsilon_{k-1}d^{-\epsilon_{k-1}}(d-d^{-1}))u^{\frac{k-3}{2}}+o(u^{\frac{k-3}{2}})\right)
=±ϵkdϵk(up12+(V11+Y21+ϵk1dϵk1(dd1)+ϵp1dϵp1(dd1))up32+o(up32)).\displaystyle=\pm\epsilon_{k}d^{-\epsilon_{k}}\left(u^{\frac{p-1}{2}}+(V_{11}+Y_{21}+\epsilon_{k-1}d^{-\epsilon_{k-1}}(d-d^{-1})+\epsilon_{p-1}d^{-\epsilon_{p-1}}(d-d^{-1}))u^{\frac{p-3}{2}}+o(u^{\frac{p-3}{2}})\right).
Claim 5.

W11=V11+Y21+ϵk1dϵk1(dd1)W_{11}=V_{11}+Y_{21}+\epsilon_{k-1}d^{-\epsilon_{k-1}}(d-d^{-1}).

Proof.

Comparing the coefficient of the second leading terms in w11=v11y11+v12y21w_{11}=v_{11}y_{11}+v_{12}y_{21}, we obtain

W11=V11+Y11ϵk1ϵkdϵk1ϵk.W_{11}=V_{11}+Y_{11}-\epsilon_{k-1}\epsilon_{k}d^{-\epsilon_{k-1}-\epsilon_{k}}.

Thus it is enough to show that Y11Y21=ϵk1ϵkdϵk1ϵk+ϵk1dϵk1(dd1)Y_{11}-Y_{21}=\epsilon_{k-1}\epsilon_{k}d^{-\epsilon_{k-1}-\epsilon_{k}}+\epsilon_{k-1}d^{-\epsilon_{k-1}}(d-d^{-1}). On the other hand, from Lemma 3.7 and

ρ(g1ϵkg2ϵk+1)=±(uϵkϵk+1dϵk+ϵk+1ϵk+1dϵk+1ϵkdϵkuϵkϵk+1dϵkϵk+1),\rho\left(g_{1}^{\epsilon_{k}}g_{2}^{\epsilon_{k+1}}\right)=\pm\begin{pmatrix}u-\epsilon_{k}\epsilon_{k+1}d^{\epsilon_{k}+\epsilon_{k+1}}&-\epsilon_{k+1}d^{-\epsilon_{k+1}}\\ \epsilon_{k}d^{-\epsilon_{k}}u&-\epsilon_{k}\epsilon_{k+1}d^{-\epsilon_{k}-\epsilon_{k+1}}\end{pmatrix},

we obtain Y11Y21=ϵkϵk+1dϵk+ϵk+1Y_{11}-Y_{21}=-\epsilon_{k}\epsilon_{k+1}d^{\epsilon_{k}+\epsilon_{k+1}}. It follows that we need to check the equation

(33) ϵk1ϵkdϵk1ϵk+ϵk1dϵk1(dd1)+ϵkϵk+1dϵk+ϵk+1=0.\epsilon_{k-1}\epsilon_{k}d^{-\epsilon_{k-1}-\epsilon_{k}}+\epsilon_{k-1}d^{-\epsilon_{k-1}}(d-d^{-1})+\epsilon_{k}\epsilon_{k+1}d^{\epsilon_{k}+\epsilon_{k+1}}=0.

Since the integer kk satisfies kq±1(mod 2p)kq\equiv\pm 1\ (\mathrm{mod}\ 2p) with p<q<p-p<q<p, the sequence (ϵk1,ϵk,ϵk+1)(\epsilon_{k-1},\epsilon_{k},\epsilon_{k+1}) can not be one of (1,1,1)(1,1,1), (1,1,1)(-1,-1,-1), (1,1,1)(1,-1,1), and (1,11)(-1,1-1). It leaves four possible cases, (ϵk1,ϵk,ϵk+1)=(1,1,1)(\epsilon_{k-1},\epsilon_{k},\epsilon_{k+1})=(1,1,-1), (1,1,1)(1,-1,-1), (1,1,1)(-1,1,1), and (1,1,1)(-1,-1,1). One can easily check that the equation (33) holds for these four cases. ∎

This completes the proof, since Claim 5 implies that (y21(dd1)y22)ϕv=±ϵkdϵkϕw+h(y_{21}-(d-d^{-1})y_{22})\phi_{v}=\pm\epsilon_{k}d^{-\epsilon_{k}}\phi_{w}+h for some h[u]h\in\mathbb{C}[u] with deghp122\deg h\leq\frac{p-1}{2}-2.

References

  • [BGZ] F. Benini, D. Gang, and L. A. Pando Zayas. Rotating Black Hole Entropy from M5 Branes. preprint arXiv:1909.11612.
  • [BZ03] G. Burde and H. Zieschang. Knots, volume 2. Walter de gruyter, 2003.
  • [CS83] M. Culler and P. B. Shalen. Varieties of group representations and splittings of 3-manifolds. Annals of Mathematics, pages 109–146, 1983.
  • [DHY09] J. Dubois, V. Huynh, and Y. Yamaguchi. Non-abelian Reidemeister torsion for twist knots. Journal of Knot Theory and Its Ramifications, 18(03):303–341, 2009.
  • [Dub03] J. Dubois. Torsion de Reidemeister non abélienne et forme volume sur l’espace des représentations du groupe d’un noeud. PhD thesis, Université Blaise Pascal-Clermont-Ferrand II, 2003.
  • [Dub06] J. Dubois. Non abelian twisted Reidemeister torsion for fibered knots. Canadian Mathematical Bulletin, 49(1):55–71, 2006.
  • [GH78] P. Griffiths and J. Harris. Principles of algebraic geometry. John Wiley & Sons, 1978.
  • [GKY] D. Gang, S. Kim, and S. Yoon. Adjoint Reidemeister torsions from wrapped M5-branes. preprint arXiv:1911.10718.
  • [GKZ] D. Gang, N. Kim, and L. A Pando Zayas. Precision microstate counting for the entropy of wrapped M5-branes. preprint arXiv:1905.01559.
  • [Guk05] S. Gukov. Three-dimensional quantum gravity, Chern-Simons theory, and the A-polynomial. Communications in mathematical physics, 255(3):577–627, 2005.
  • [Kit96] T. Kitano. Twisted Alexander polynomial and Reidemeister torsion. Pacific Journal of Mathematics, 174(2):431–442, 1996.
  • [OT15] T. Ohtsuki and T. Takata. On the Kashaev invariant and the twisted Reidemeister torsion of two-bridge knots. Geometry & Topology, 19(2):853–952, 2015.
  • [Por97] J. Porti. Torsion de Reidemeister pour les variétés hyperboliques, volume 612. American Mathematical Soc., 1997.
  • [Ril72] R. Riley. Parabolic representations of knot groups, I. Proceedings of the London Mathematical Society, 3(2):217–242, 1972.
  • [Ril84] R. Riley. Nonabelian representations of 2-bridge knot groups. The Quarterly Journal of Mathematics, 35(2):191–208, 1984.
  • [Sch54] Horst Schubert. Über eine numerische knoteninvariante. Mathematische Zeitschrift, 61(1):245–288, 1954.
  • [Sik12] A. Sikora. Character varieties. Transactions of the American Mathematical Society, 364(10):5173–5208, 2012.
  • [Tra14] A. Tran. Twisted Alexander polynomials with the adjoint action for some classes of knots. Journal of Knot Theory and Its Ramifications, 23(10):1450051, 2014.
  • [Tra18] A. Tran. Twisted Alexander polynomials of genus one two-bridge knots. Kodai Mathematical Journal, 41(1):86–97, 2018.
  • [Tur01] V. Turaev. Introduction to combinatorial torsions. Springer Science & Business Media, 2001.
  • [Tur02] V. Turaev. Torsions of 3-manifolds. Geometry & Topology Monographs, 4, 2002.
  • [Wei64] André Weil. Remarks on the cohomology of groups. Annals of Mathematics, pages 149–157, 1964.
  • [Wit89] E. Witten. Quantum field theory and the Jones polynomial. Communications in Mathematical Physics, 121(3):351–399, 1989.
  • [Yam08] Y. Yamaguchi. A relationship between the non-acyclic Reidemeister torsion and a zero of the acyclic Reidemeister torsion. In Annales de l’institut Fourier, volume 58, pages 337–362, 2008.