Exact integral formulas for volumes of two-bridge knot cone-manifolds


Abstract

We provide exact integral formulas for hyperbolic and spherical volumes of cone-manifolds whose underlying space is the \(3\)-sphere and whose singular set belongs to three infinite families of two-bridge knots: \(C(2n,2)\) (twist knots), \(C(2n,3)\), and \(C(2n,-2n)\) for any non-zero integer \(n\). Our formulas express volumes as integrals of explicit rational functions involving Chebyshev polynomials of the second kind, with integration limits determined by roots of algebraic equations. This extends previous work where only implicit formulas requiring numerical approximation were known.

1 2

1 Introduction↩︎

An \(n\)-dimensional cone-manifold is a simplicial complex \(M\) which can be triangulated so that the link of each simplex is piecewise-linear homeomorphic to a standard \((n-1)\)-sphere and \(M\) is equipped with a complete path metric such that the restriction of the metric to each simplex is isometric to a geodesic simplex of constant curvature \(\kappa\). The cone-manifold is hyperbolic, Euclidean, or spherical if \(\kappa\) is \(-1\), \(0\), or \(+1\) respectively.

The singular locus \(\Sigma\) of a cone-manifold \(M\) consists of the points in \(M\) with no neighborhood isometric to a ball in a Riemannian manifold. Then \(\Sigma\) is a union of totally geodesic closed simplices of dimension \(n-2\). At each point of \(\Sigma\) in an open \((n-2)\)-simplex, there is a cone angle which is the sum of dihedral angles of \(n\)-simplices containing the point. In general, the cone angle may vary from point to point within a simplex. The regular set \(M \setminus \Sigma\) is a dense open subset of \(M\) and has a smooth Riemannian metric of constant curvature \(\kappa\), but this metric is incomplete if \(\Sigma \not= \emptyset\).

In this paper, we will only consider \(3\)-dimensional cone-manifolds whose underlying space \(M\) is the \(3\)-sphere \(\mathbb{S}^3\) and whose singular set is a knot \(K\) with constant cone angle \(\alpha \in (0,2\pi]\). We will denote these cone manifolds by \(K(\alpha)\).

A two-bridge knot, also known as a rational knot, is a knot that admits a projection with two maxima and two minima. In the Conway notation, a two-bridge knot corresponds to a continued fraction \[[a_1, a_2, \dots, a_k] = a_1 + \dfrac{1}{a_2 + \dfrac{1}{\ddots + \dfrac{1}{a_k}}}\] and denoted by \(C(a_1, a_2, \dots, a_k)\). Its diagram is shown in Figure 1. In the \(a_i\) box, \(|a_i|\) denotes the number of signed half-twists and the sign of each half-twist is the same as the sign of \(a_i \in \mathbb{Z}\). Here, we use the convention that the sign of the right-handed half-twist \(\raisebox{-5pt}{ \epsfig{file=leftcross.pdf}}\) in the \(a_i\) box is positive for odd \(i\) and negative for even \(i\). In the Schubert notation, \(C(a_1, a_2, \dots, a_k)\) is the two-bridge knot \(\mathfrak b(p,q)\) where \(\frac{p}{q} = [a_1, a_2, \dots, a_k]\).

Figure 1: The rational knot C(a_1,a_2, \ldots,a_k). The upper/lower one corresponds to odd/even k.

Geometric transitions of cone-manifolds↩︎

For a two-bridge knot \(K\), Kojima [1] and Porti [2] established that there exists a critical angle \(\alpha_K \in [\frac{2\pi}{3}, \pi)\) such that the cone manifold \(K(\alpha)\) undergoes geometric transitions:

  • hyperbolic structure for \(\alpha \in (0, \alpha_K)\),

  • Euclidean structure for \(\alpha = \alpha_K\),

  • spherical structure for \(\alpha \in (\alpha_K, 2\pi - \alpha_K)\).

Previous work on volume formulas↩︎

Hilden, Lozano, and Montesinos-Amilibia [3] introduced a method for calculating volumes of two-bridge knot cone manifolds, but without providing explicit formulas. Integral formulas for hyperbolic volumes have been obtained for specific families:

  • \(C(2n, 2)\) (twist knots) by Ham, Mednykh, and Petrov [4],

  • \(C(2n,3)\) by Ham and Lee [5],

  • \(C(2n,k)\) (double twist knots \(J(-2n, k)\)) by Tran [6].

However, these formulas involve implicitly defined integrands and are primarily useful for numerical approximation. The only exact integral formulas previously known were those given by Mednykh [7] for two-bridge knots with up to seven crossings.

Our contribution↩︎

In this paper, we extend Mednykh’s approach to obtain exact integral formulas for infinite families of two-bridge knots. We will give exact integral formulas for hyperbolic and spherical volumes of cone-manifolds along two-bridge knots \(C(2n, 2)\), \(C(2n, 3)\) and \(C(2n, -2n)\), where \(n\) is a non-zero integer.

To state our main result, we introduce the Chebychev polynomials of the second kind \(S_k(z)\). They are recursively defined by \(S_0(z)=1\), \(S_1(z)=z\) and \(S_{k}(z) = z S_{k-1}(z) - S_{k-2}(z)\) for all integers \(k\).

For \(K=C(2n,3)\), \(C(2n, 2)\) or \(C(2n, -2n)\) we let \[\begin{align} f_n(y) &=& \frac{2S_n(y)-yS_{n-1}(y)}{(y-2)S_{n-1}(y)}, \\ g_n(y) &=& \begin{cases} - \displaystyle{\frac{(S_n(y) - S_{n-1}(y))^2}{(y-2)^3S^4_{n-1}(y)}} & \quad \text{if } K=C(2n,3), \\ - \displaystyle{\frac{S_n(y) - S_{n-1}(y)}{(y-2)^2S^3_{n-1}(y)}} & \quad \text{if } K=C(2n,2), \\ \displaystyle{\frac{1}{(y-2)^2S^4_{n-1}(y)}} & \quad \text{if } K=C(2n, -2n). \end{cases} \end{align}\] Then the hyperbolic and spherical volumes of the cone-manifold \(K(\alpha)\) are given as follows.

Theorem 1. For \(\alpha \in (0, \alpha_K)\) we have \[\mathrm{Vol}(K(\alpha)) = i \int_{\overline{y_0}}^{y_0} \log \left( \frac{f^2_n(y)+ A^2}{(1+A^2)g_n(y)} \right) \frac{f'_n(y)}{f^2_n(y)-1} dy ,\] where \(y_0\), with \(\mathrm{Im} (f_n(y_0)) > 0\), is a root of \(f^2_n(y)+ A^2= (1+A^2)g_n(y)\) and \(A=\cot \frac{\alpha}{2}\).

Theorem 2. For \(\alpha \in (\alpha_K, 2\pi - \alpha_K)\) we have \[\mathrm{Vol}(K(\alpha)) = \int_{y_+}^{y_-} \log \left( \frac{f^2_n(y)+ A^2}{(1+A^2)g_n(y)} \right) \frac{f'_n(y)}{f^2_n(y)-1} dy,\] where \(y_{\pm}\), with \(f_n(y_{\pm}) \in \mathbb{R}\), are roots of \(f^2_n(y)+ A^2= (1+A^2)g_n(y)\) and \(A=\cot \frac{\alpha}{2}\).

As in [7], the proofs of Theorems 1 and 2 are based on

  • trigonometric identity between the cone angle \(\alpha\) and the complex length \(\gamma_\alpha\) of the singular geodesic \(K\) in the cone-manifold \(K(\alpha)\), and

  • the Schläfli formula \[\kappa \, d \mathrm{Vol}(K(\alpha)) = \frac{1}{2} l_\alpha d\alpha,\] where \(l_\alpha = \mathrm{Re} \, \gamma_\alpha > 0\) is the real length of \(K\).

The paper is organized as follows. In Section 2 we briefly review holonomy representations of hyperbolic and spherical knot cone-manifolds. In Section 3 we first study \(\mathrm{SL}_2(\mathbb{C})\)-representations of \(C(2n, 2p+1)\), then prove trigonometric identity between the cone angle and the complex length of the singular geodesic, and finally give a proof of Theorems 1 and 2 for \(C(2n,3)\). In Section 4 we carry out the same things for \(C(2n,2)\) and \(C(2n, -2n)\).

2 Knot cone-manifolds↩︎

Recall that \(K(\alpha)\) denotes the \(3\)-dimensional cone manifolds whose underlying space \(M\) is the \(3\)-sphere \(S^3\) and whose singular set is a knot \(K\) with constant cone angle \(\alpha \in (0,2\pi]\). Let \(G(K) := \pi_1(S^3 \setminus K)\) be the knot group, which is the fundamental group of the knot exterior. Choose the canonical meridian-longitude pair \((\mu, \lambda)\) in \(G(K)\) such that \(\mu\) is an oriented boundary of meridian disk of \(K\) and \(\lambda\) is null-homologous outside \(K\).

If \(K(\alpha)\) is hyperbolic, then let \(\rho_\alpha: G(K) \to \mathrm{Isom}^+(\mathbb{H}^3) \cong \mathrm{PSL}_2(\mathbb{C})\) be the holonomy representation. Then \(\rho_\alpha\) admits two liftings to \(\mathrm{SL}_2(\mathbb{C})\). Up to conjugation in \(\mathrm{SL}_2(\mathbb{C})\), we can assume that \[\rho_\alpha (\mu) = \pm \left[ \begin{array}{cc} e^{i\alpha/2} & 0 \\ 0 & e^{-i\alpha/2} \end{array} \right], \quad \rho_\alpha(\lambda) = \left[ \begin{array}{cc} e^{\gamma_\alpha/2} & 0\\ 0 & e^{-\gamma_\alpha/2} \end{array} \right]\] where \(\gamma_\alpha = l_\alpha + i \varphi_\alpha\), \(l_\alpha\) is the length of \(K\), and \(\varphi_\alpha \in [-2\pi, 2\pi)\) is the angle of the lifted holonomy of \(K\). We call \(\gamma_\alpha = l_\alpha + i \varphi_\alpha\) the complex length of the singular geodesic \(K\).

If \(K(\alpha)\) is spherical, then let \(\rho_\alpha: G(K) \to \mathrm{Isom}^+(\mathbb{S}^3) \cong\mathrm{SO}(4)\) be the holonomy representation. Then \(\rho_\alpha\) admits two liftings to \(\mathrm{SU}(2) \times \mathrm{SU}(2)\). Up to conjugation in \(\mathrm{SU}(2) \times \mathrm{SU}(2)\), we can assume that \[\begin{align} \rho_\alpha(\mu) &=& \left( \pm \left[ \begin{array}{cc} e^{i\alpha/2} & 0 \\ 0 & e^{-i\alpha/2} \end{array} \right], \pm \left[ \begin{array}{cc} e^{i\alpha/2} & 0 \\ 0 & e^{-i\alpha/2} \end{array} \right] \right), \\ \rho_\alpha(\lambda) &=& \left( \left[ \begin{array}{cc} e^{i\gamma} & 0 \\ 0 & e^{-i\gamma} \end{array} \right], \left[ \begin{array}{cc} e^{i\phi} & 0 \\ 0 & e^{-i\phi} \end{array} \right]\right). \end{align}\] In this case \(l_\alpha = \gamma - \phi\) is the length of the knot \(K\), and \(\varphi_\alpha = \gamma + \phi \in [-2\pi, 2\pi)\) is the angle of the lifted holonomy of \(K\). Note that \(\gamma = \frac{1}{2}(\varphi_\alpha + l_\alpha)\) and \(\phi = \frac{1}{2}(\varphi_\alpha - l_\alpha)\).

3 \(C(2n, 2p+1)\)↩︎

3.1 Knot group↩︎

Proposition 1. We have \(G(C(2n, 2p+1)) = \langle a, b \mid \omega a = b \omega\rangle\) where \[\omega = (ab)^n [(a^{-1}b^{-1})^{n}(ab)^n]^p.\]

Figure 2: C(2n,2p+1)

Proof. Starting from the left hand section of \(2n\) crossings, by induction we have \[\begin{align} a_k &= (b_0 a_0)^{-k} a_0 (b_0 a_0)^{k}, \\ b_k &= (b_0 a_0)^{-k} b_0 (b_0 a_0)^{k}. \end{align}\] Similarly, in the right hand section of \(l\) crossings we have \[\begin{align} c_k &= (c_0^{-1}d_0)^{-k} c_0 (c_0^{-1}d_0)^{k}, \\ d_k &= (c_0^{-1}d_0)^{-k} d_0 (c_0^{-1}d_0)^{k}. \end{align}\]

By using the identity \(a_0 = c_{p+1}\) we have \[\begin{align} && a_0 = (c_0^{-1}d_0)^{-p-1} c_0 (c_0^{-1}d_0)^{p+1} \\ &\Longrightarrow& a_0 = (b_n^{-1}b_0)^{-p-1} b_n (b_n^{-1}b_0)^{p+1} \\ &\Longrightarrow& a_0 = (b_n^{-1}b_0)^{-p-1} (b_0 a_0)^{-n} b_0 (b_0 a_0)^{n} (b_n^{-1}b_0)^{p+1} \\ &\Longrightarrow& (b_0 a_0)^{n} (b_n^{-1}b_0)^{p+1} a_0 = b_0 (b_0 a_0)^{n} (b_n^{-1}b_0)^{p+1}. \end{align}\] Hence \(\omega a_0 = b_0 \omega\) where \(\omega = (b_0 a_0)^{n} (b_n^{-1}b_0)^{n+1}\).

Let \(a=a_0\) and \(b = b_0\). Then \(b_n^{-1}b_0 = (b a)^{-n} b^{-1} (b a)^{n}b = (b a)^{-n} (ab)^n\). Hence \[\omega = (b_0 a_0)^{n} (b_n^{-1}b_0)^{p+1}= (ba)^n [(b a)^{-n} (ab)^n]^{p+1} = (ab)^n [(b a)^{-n} (ab)^n]^p.\] This completes the proof. ◻

Note that the knot group presentation in Proposition 1 is different from the one in [8][10] (where \(C(2n, 2p+1)\) is denoted by \(J(-2n, 2p+1)\)), but it can be applied to find exact integral formulas for volumes of cone-manifolds along \(C(2n, 2p+1)\).

3.2 \(\mathrm{SL}_2(\mathbb{C})\)-representations↩︎

Suppose \(\rho\colon G(C(2n, 2p+1))\to \mathrm{SL}_2(\mathbb{C})\) is a nonabelian representation. Up to conjugation, we may assume that \[\label{repn} A:=\rho(a) = \left[ \begin{array}{cc} m & 1\\ 0 & m^{-1} \end{array} \right] \quad \text{and} \quad B:=\rho(b) = \left[ \begin{array}{cc} m & 0\\ y - m^2 - m^{-2} & m^{-1} \end{array} \right]\tag{1}\] where \((m,y) \in \mathbb{C}^2\) satisfies \(\rho(\omega a) = \rho(b \omega)\). Note that \(y=\mathop{\mathrm{\mathrm tr}}\rho(ab)\).

We now solve the matrix equation \(\rho(\omega a) = \rho(b \omega)\). Recall that \(S_k(z)\)’s are the Chebychev polynomials defined by \(S_0(z)=1\), \(S_1(z)=z\) and \(S_{k}(z) = z S_{k-1}(z) - S_{k-2}(z)\) for \(k \in \mathbb{Z}\). Note that \(S_k(z) = (s^{k+1} - s^{-k-1})/(s - s^{-1})\) if \(z = s + s^{-1}\).

The following lemmas are elementary, see e.g. [10] and references therein.

Lemma 1. For any integer \(k\) we have \[S^2_k(z) - z S_{k}(z) S_{k-1}(z) + S^2_{k-1}(z) =1.\]

Lemma 2. Suppose \(M \in \mathrm{SL}_2(\mathbb{C})\) and \(z= \mathop{\mathrm{\mathrm tr}}M.\) For any integer \(k\) we have \[M^k = S_k(z) I - S_{k-1}(z) M^{-1}.\]

Let \(x= \mathop{\mathrm{\mathrm tr}}\rho(a)=\mathop{\mathrm{\mathrm tr}}\rho(b) = m+m^{-1}\). Let \(U=\rho((a^{-1} b^{-1})^{n} (ab)^{n})\) and \(u = \mathop{\mathrm{\mathrm tr}}U\).

Proposition 2. We have \[u = 2 + (y-2) (y+2-x^2) S^2_{n-1}(y).\]

Proof. Recall that \(A = \rho(a)\) and \(B= \rho(b)\). Since \(\mathop{\mathrm{\mathrm tr}}A^{-1} B^{-1} = y\) and \(\mathop{\mathrm{\mathrm tr}}AB = y\), by Lemma 2 we have \[\begin{align} U &= (A^{-1} B^{-1})^{n} (AB)^{n} \\ &= (S_n(y)I - S_{n-1}(y)BA) (S_n(y) I - S_{n-1}(y) B^{-1}A^{-1})\\ &= S_n^2(y) I + S_{n-1}^2(y) BA B^{-1}A^{-1} - S_n(y)S_{n-1}(y) (B^{-1}A^{-1} + BA). \end{align}\] Taking trace we obtain \[u= \mathop{\mathrm{\mathrm tr}}U = 2S_n^2(y) + (\mathop{\mathrm{\mathrm tr}}BA B^{-1}A^{-1}) S_{n-1}^2(y) - 2y S_n(y)S_{n-1}(y).\]

By Lemma 1 we have \(S_n^2(y) - y S_n(y) S_{n-1}(y) + S_{n-1}^2(y) =1\). This implies that \[u = 2 + (\mathop{\mathrm{\mathrm tr}}BA B^{-1}A^{-1}-2) S_{n-1}^2(y).\] Finally, by a direct calculation using the matrix form 1 we have \[\mathop{\mathrm{\mathrm tr}}BA B^{-1}A^{-1}-2 = (y-2) (y - m^2 - m^{-2}) = (y-2) (y+2-x^2).\] Hence \(u = 2 + (y-2) (y+2-x^2) S^2_{n-1}(y)\). ◻

Proposition 3. We have \[\rho(\omega a) - \rho(b \omega) = \left[ \begin{array}{cc} 0 & \Phi_{C(2n, 2p+1)}(x,y)\\ (x^2-2-y) \Phi_{C(2n, 2p+1)}(x,y) & 0 \end{array} \right]\] where \[\Phi_{C(2n, 2p+1)}(x,y) = (S_n(y) - S_{n-1}(y)) S_p(u) - (S_{n-1}(y) - S_{n-2}(y)) S_{p-1}(u).\]

Proof. Let \(W= \rho(\omega)\). Then \(W = \rho((ab)^n [(b a)^{-n} (ab)^n]^p)= (AB)^n U^p\). Since \(\mathop{\mathrm{\mathrm tr}}U = u\), by Lemma 2 we have \[\begin{align} W &=& (AB)^n (S_p(u) I - S_{p-1}(u) U^{-1}) \\ &=& S_p(u) (AB)^n - S_{p-1}(u) (BA)^n \\ &=& S_p(u) (S_n(y) I - S_{n-1}(y) B^{-1}A^{-1}) - S_{p-1}(u) (S_n(y) I - S_{n-1}(y) A^{-1}B^{-1}). \end{align}\] Hence \[\begin{align} WA - BW &=& S_p(u) [S_n(y) (A-B) - S_{n-1}(y) (B^{-1}-A^{-1})] \\ && - \, S_{p-1}(u) [S_n(y) (A-B) - S_{n-1}(y) (A^{-1}B^{-1}A - B A^{-1}B^{-1})]. \end{align}\] By direct calculations using the matrix form 1 we have \[\begin{align} A - B &=& \left[ \begin{array}{cc} 0 & 1\\ m^2 + m^{-2} - y & 0 \end{array} \right], \\ B^{-1}-A^{-1} &=& \left[ \begin{array}{cc} 0 & 1\\ m^2 + m^{-2} - y & 0 \end{array} \right], \\ A^{-1}B^{-1}A - B A^{-1}B^{-1} &=& \left[ \begin{array}{cc} 0 & y-1\\ (y-1)(m^2 + m^{-2} - y) & 0 \end{array} \right]. \end{align}\] Hence \(WA - B W = \left[ \begin{array}{cc} 0 & \Phi\\ (m^2 + m^{-2} -y) \Phi & 0 \end{array} \right]\) where \[\Phi = S_p(u) (S_n(y) - S_{n-1}(y)) - S_{p-1} (u) (S_n(y) - (y-1)S_{n-1}(y)).\] Finally, since \(S_n(y) - (y-1)S_{n-1}(y) = S_{n-1}(y) - S_{n-2}(y)\) the proposition follows. ◻

Proposition 3 implies that \(\rho(\omega a) = \rho(b \omega)\) if and only if \(\Phi_{C(2n, 2p+1)}(x,y) =0\).

Remark 4. The polynomial \(\Phi_{C(2n, 2p+1)}(x,y)\) is called the Riley polynomial of the two-bridge knot \(C(2n, 2p+1)\), see [11].

3.3 Longitude and trignometric identity↩︎

If we choose the meridian to be \(\mu=a\) then the canonical longitude is \(\lambda = \omega \omega^* a^{-4n}\), where \(\omega^*\) is the word obtained from \(\omega\) by writing the letters in \(\omega\) in reversed order.

Since \(\rho(\mu) = \left[ \begin{array}{cc} m & 1\\ 0 & m^{-1} \end{array} \right]\) we have \(\rho(\lambda) = \left[ \begin{array}{cc} l & *\\ 0 & l^{-1} \end{array} \right]\). By [8] we have \(lm^{4n} = - \frac{\widetilde{W_{12}}}{W_{12}}\), where \(W_{12}\) is the \((1,2)\)-entry of \(W=\rho(\omega)\) and \(\widetilde{W_{12}}\) is obtained from \(W_{12}\) by replacing \(m\) by \(m^{-1}\). Note that \(W_{12}\) is a function in \(m\) and \(y\).

Proposition 5. We have \[W_{12} = \left(m^{-1} - m \frac{S_n(y) - S_{n-1}(y)}{S_{n-1}(y) - S_{n-2}(y)} \right) S_{p}(u)S_{n-1}(y).\]

Proof. From the proof of Proposition 3 we have \[W = S_p(u) (S_n(y) I - S_{n-1}(y) B^{-1}A^{-1}) - S_{p-1}(u) (S_n(y) I - S_{n-1}(y) A^{-1}B^{-1}).\] Taking the \((1,2)\)-entry we have \[W_{12} = - S_p(u) S_{n-1}(y) (B^{-1}A^{-1})_{12} + S_{p-1}(u) S_{n-1}(y) (A^{-1}B^{-1})_{12}.\] Since \((B^{-1}A^{-1})_{12} = - m^{-1}\) and \((A^{-1}B^{-1})_{12} = -m\), we obtain \[W_{12} = (m^{-1} S_p(u) - m S_{p-1}(u)) S_{n-1}(y).\]

We now simplify \(W_{12}\) by using \(\Phi_{C(2n, 2p+1)}(x,y) =0\). Since \((S_n(y) - S_{n-1}(y)) S_p(u) - (S_{n-1}(y) - S_{n-2}(y)) S_{p-1}(u)=0\), we have \(S_{p-1}(u) = \frac{S_n(y) - S_{n-1}(y)}{S_{n-1}(y) - S_{n-2}(y)} S_p(u)\). Hence \[W_{12} = \left(m^{-1} - m \frac{S_n(y) - S_{n-1}(y)}{S_{n-1}(y) - S_{n-2}(y)} \right) S_{p}(u)S_{n-1}(y)\] as claimed. ◻

Proposition 6. Let \(f_n(y) = \frac{2S_n(y)-yS_{n-1}(y)}{(y-2)S_{n-1}(y)}\). Then \[f_n(y) = - \frac{\ell^{1/2} +\ell^{-1/2}}{\ell^{1/2}-\ell^{-1/2}} \cdot \frac{m+m^{-1}}{m-m^{-1}},\] where \(\ell = lm^{4n}\).

Proof. Since \(\ell = lm^{4n}= -\widetilde{W_{12}} /W_{12}\), by Proposition 5 we have \[\ell = - \frac{m - m^{-1} r}{m^{-1} - m r} = \frac{m^2 - r}{m^2 r-1}\] where \(r =\frac{S_n(y) - S_{n-1}(y)}{S_{n-1}(y) - S_{n-2}(y)}\). This implies that \[\frac{\ell+1}{\ell-1} = - \frac{(m^2-1)(r+1)}{(m^2+1)(r-1)}.\] Hence \[\frac{\ell+1}{\ell-1} \cdot \frac{m^2+1}{m^2-1} = - \frac{r+1}{r-1} = - \frac{S_n(y)-S_{n-2}(y)}{(y-2)S_{n-1}(y)} = - f_n(y).\] ◻

Hyperbolic case: Let \(K(\alpha)\) be a hyperbolic 3-dimensional cone-manifold whose singular set is \(K=C(2n, 2p+1)\) with cone angle \(\alpha \in (0, 2\pi]\). Up to conjugation in \(\mathrm{SL}_2(\mathbb{C})\), \[\rho_\alpha (\mu) = \pm \left[ \begin{array}{cc} e^{i\alpha/2} & 0 \\ 0 & e^{-i\alpha/2} \end{array} \right], \quad \rho_\alpha(\lambda) = \left[ \begin{array}{cc} e^{\gamma_\alpha/2} & 0\\ 0 & e^{-\gamma_\alpha/2} \end{array} \right]\] where \(\gamma_\alpha = l_\alpha + i \varphi_\alpha\) is the complex length of the singular geodesic \(K\) in \(K(\alpha)\), \(l_\alpha > 0\) is the real length of \(K\), and \(\varphi_\alpha \in [-2\pi, 2\pi)\) is the angle of the lifted holonomy of \(K\).

Proposition 7. In the hyperbolic case we have \[i \coth \left( \frac{\gamma_\alpha + 4n i \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right) = f_n(y).\] In particular, we have \(\mathrm{Im} (f_n(y)) > 0\).

Proof. Since \(m = e^{i \alpha/2}\) and \(\ell = lm^{4n} = e^{(\gamma_\alpha + 4ni \alpha)/2}\), we obtain \[\begin{align} f_n(y) &=& - \frac{\ell^{1/2} +\ell^{-1/2}}{\ell^{1/2}-\ell^{-1/2}} \cdot \frac{m+m^{-1}}{m-m^{-1}} \\ &=& i \coth \left( \frac{\gamma_\alpha + 4n i \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right). \end{align}\] Note that \(\cot \left( \frac{\alpha}{2} \right) > 0\) and \(\mathrm{Re} (\gamma_\alpha + 4n i \alpha) = l_\alpha>0\). Hence \[\mathrm{Re}(-i f_n(y)) = \cot \left( \frac{\alpha}{2} \right) \mathrm{Re} \coth \left( \frac{\gamma_\alpha + 4n i \alpha}{4} \right) >0.\] This implies that \(\mathrm{Im} (f_n(y)) > 0\). ◻

Spherical case: Let \(K(\alpha)\) be a spherical 3-dimensional cone-manifold whose singular set is \(K=C(2n, 2p+1)\) with cone angle \(\alpha \in (0, 2\pi]\). Up to conjugation in \(\mathrm{SU}(2) \times \mathrm{SU}(2)\), we can assume that \[\begin{align} \rho_\alpha(\mu) &=& \left( \pm \left[ \begin{array}{cc} e^{i\alpha/2} & 0 \\ 0 & e^{-i\alpha/2} \end{array} \right], \pm \left[ \begin{array}{cc} e^{i\alpha/2} & 0 \\ 0 & e^{-i\alpha/2} \end{array} \right] \right), \\ \rho_\alpha(\lambda) &=& \left( \left[ \begin{array}{cc} e^{i\gamma} & 0 \\ 0 & e^{-i\gamma} \end{array} \right], \left[ \begin{array}{cc} e^{i\phi} & 0 \\ 0 & e^{-i\phi} \end{array} \right]\right). \end{align}\] In this case \(l_\alpha = \gamma - \phi\) is the length of the knot \(K\), and \(\varphi_\alpha = \gamma + \phi \in [-2\pi, 2\pi)\) is the angle of the lifted holonomy of \(K\). Note that \(\gamma = \frac{1}{2}(\varphi_\alpha + l_\alpha)\) and \(\phi = \frac{1}{2}(\varphi_\alpha - l_\alpha)\). Hence \(m = e^{i\alpha/2}\) and \(\ell=e^{i(\varphi_\alpha \pm l_\alpha)/2}\).

Proposition 8. In the spherical case we have \[\cot \left( \frac{\varphi_\alpha \pm l_\alpha+ 4n \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right) = f_n(y_{\pm}).\] In particular, we have \(f_n(y_{\pm}) \in \mathbb{R}\).

Proof. Since \(m = e^{i \alpha/2}\) and \(\ell = lm^{4n} = e^{i(\varphi_\alpha \pm l_\alpha + 4n\alpha)/2}\), we obtain \[\begin{align} f_n(y_{\pm}) &=& - \frac{\ell^{1/2} +\ell^{-1/2}}{\ell^{1/2}-\ell^{-1/2}} \cdot \frac{m+m^{-1}}{m-m^{-1}} \\ &=& \cot \left( \frac{\varphi_\alpha \pm l_\alpha+ 4n \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right). \end{align}\] ◻

3.4 Proof of Theorems 1 and 2 for \(C(2n,3)\)↩︎

Suppose \(p=1\). By Propositions 2 and 3 we have \(u = 2 + (y-2) (y+2-x^2) S^2_{n-1}(y)\) and \[\begin{align} \Phi_{C(2n,3)}(x,y) &=& (S_n(y) - S_{n-1}(y)) u - (S_n(y) - (y-1)S_{n-1}(y)) \\ &=& (y-2) (y+2-x^2) S^2_{n-1}(y) (S_n(y) - S_{n-1}(y)) \\ && + \, 2 (S_n(y) - S_{n-1}(y)) - (S_n(y) - (y-1)S_{n-1}(y)). \end{align}\]

Lemma 3. Suppose \(\Phi(x,y) = b - a x^2\) where \(a, b \in \mathbb{C}(y)\). Let \(A = \cot \frac{\alpha}{2}\). Then, for any \(c \in \mathbb{C}(y)\), the equation \(\Phi(2\cos \frac{\alpha}{2}, y)=0\) is equivalent to \(c^2+ A^2= (1+A^2)d\) where \(d = 1 + (c^2-1)(1 - \frac{b}{4a})\).

Proof. Let \(x = 2\cos \frac{\alpha}{2}\). We have \(4-x^2 = 4 \sin^2 \frac{\alpha}{2} = \frac{4}{A^2+1}\). Hence \[\begin{align} \Phi(x, y)=0 &\Longleftrightarrow& x^2 =\frac{b}{a} \\ &\Longleftrightarrow& \frac{4}{A^2+1} = 4 - \frac{b}{a} = \frac{4(d-1)}{c^2 - 1} \\ &\Longleftrightarrow& (A^2 + 1) (d-1) = c^2 -1 \\ &\Longleftrightarrow& c^2+ A^2= (1+A^2)d. \end{align}\] This proves the lemma. ◻

We write \(\Phi_{C(2n,3)}(x,y) = b - a x^2\), where \[\begin{align} a &=& (y-2) S^2_{n-1}(y) (S_n(y) - S_{n-1}(y)), \\ b &=& S_n(y) + (y-3) S_{n-1}(y) + (y-2) (y+2) S^2_{n-1}(y) (S_n(y) - S_{n-1}(y)). \end{align}\] Note that \(b - 4a = S_n(y) + (y-3) S_{n-1}(y) + (y-2)^2 S^2_{n-1}(y) (S_n(y) - S_{n-1}(y))\).

Choose \(c = f_n(y)\). By Lemma 3, the equation \(\Phi_{C(2n,3)}(2\cos \frac{\alpha}{2}, y)=0\) is equivalent to \(c^2+ A^2= (1+A^2)d\) where \(d = 1 + (c^2-1)(1 - \frac{b}{4a})\). Since \[c^2 -1 = \frac{4(S_n(y)-S_{n-1}(y)) (S_n(y) - (y-1) S_{n-1}(y))}{(y-2)^2 S^2_{n-1}(y)},\] we have \[\begin{align} (c^2-1) \left( 1 - \frac{b}{4a} \right) &=& - \frac{S_n(y) - (y-1) S_{n-1}(y)}{(y-2)^3 S^4_{n-1}(y)} [ S_n(y) + (y-3) S_{n-1}(y) \\ && \qquad \qquad \qquad \qquad + (y-2)^2 S^2_{n-1}(y) (S_n(y) - S_{n-1}(y)) ] \\ &=& - \frac{1}{(y-2)^3 S^4_{n-1}(y)} [ (S_n(y) - S_{n-1}(y))^2 - (y-2)^2 S^2_{n-1}(y) \\ && + \, (y-2)^2 S^2_{n-1}(y) \left( S^2_n(y) - y S_n(y) S_{n-1}(y) + (y-1) S^2_{n-1}(y) \right) ]. \end{align}\]

By Lemma 1 we have \(S^2_n(y) - y S_n(y) S_{n-1}(y) + S^2_{n-1}(y) = 1\). This implies that \(S^2_n(y) - y S_n(y) S_{n-1}(y) + (y-1) S^2_{n-1}(y) = 1 + (y-2) S^2_{n-1}(y)\). Hence \[\begin{align} (c^2-1) \left( 1 - \frac{b}{4a} \right) &=& - \frac{(S_n(y) - S_{n-1}(y))^2 + (y-2)^3 S^4_{n-1}(y)}{(y-2)^3 S^4_{n-1}(y)} \end{align}\] and \(d = 1 + (c^2-1) \left( 1 - \frac{b}{4a} \right) = - \frac{(S_n(y) - S_{n-1}(y))^2}{(y-2)^3 S^4_{n-1}(y)}\).

In summary, we have proved the following.

Proposition 9. Let \[f_n(y) = \frac{2S_n(y)-yS_{n-1}(y)}{(y-2)S_{n-1}(y)}, \qquad g_n(y) = - \frac{(S_n(y) - S_{n-1}(y))^2}{(y-2)^3S^4_{n-1}(y)}.\] Let \(x = 2\cos \frac{\alpha}{2}\) and \(A = \cot \frac{\alpha}{2}\). Then the equation \(\Phi_{C(2n, 3)}(x,y) = 0\) is equivalent to \(f^2_n(y)+ A^2= (1+A^2)g_n(y)\).

For a two-bridge knot \(K\), there exists an angle \(\alpha_K \in [\frac{2\pi}{3},\pi)\) such that \(K(\alpha)\) is hyperbolic for \(\alpha \in (0, \alpha_K)\), Euclidean for \(\alpha =\alpha_K\), and spherical for \(\alpha \in (\alpha_K, 2\pi - \alpha_K)\).

3.4.1 Hyperbolic case↩︎

For \(\alpha \in (0, \alpha_K)\), by the Schläfli formula we have \[\frac{d\mathrm{Vol}(K(\alpha))}{d\alpha} = - \frac{1}{2} l_\alpha\] where \(l_\alpha = \mathrm{Re}(\gamma_\alpha) > 0\) is the real length of \(K \subset K(\alpha)\). Note that \(K(\alpha)\) is Euclidean at \(\alpha = \alpha_K\), so \(\mathrm{Vol}(K(\alpha)) \to 0\) as \(\alpha \to \alpha_K\). Let \[\label{F} F(\alpha)= i \int_{\overline{y_0}}^{y_0} \log \left( \frac{f^2_n(y)+ A^2}{(1+A^2)g_n(y)} \right) \frac{f'_n(y)dy}{f^2_n(y)-1}.\tag{2}\] Then Theorem 1 is equivalent to \(\mathrm{Vol}(K(\alpha)) = F(\alpha)\).

We first claim that \(F(\alpha) \to 0\) as \(\alpha \to \alpha_K\). Indeed, as \(\alpha \to \alpha_K\) we have \(l_\alpha \to 0\) and so \(\gamma_\alpha = \ell_\alpha + i \varphi_\alpha \to i \varphi_{\alpha_K}\). Then, by the trigonometric identity (Proposition 7) we obtain \[\begin{align} f_n(y_0) &=& i \coth \left( \frac{\gamma_\alpha + 4n i \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right) \\ &\to& i \coth\left( \frac{i \varphi_{\alpha_K} + 4n i \alpha_K}{4} \right) \cot \left( \frac{\alpha_K}{2} \right) \\ &=& \cot \left( \frac{\varphi_{\alpha_K} + 4n \alpha_K}{4} \right) \cot \left( \frac{\alpha_K}{2} \right), \end{align}\] where we used \(\coth (iz) = -i \cot(z)\). Then \(\mathrm{Im} \, f_n(y_0) \to 0\). Hence \(f_n(\overline{y_0}) - f_n(y_0) = \overline{ f_n(y_0)} - f_n(y_0) = - 2 i \, \mathrm{Im} \, f_n(y_0) \to 0\).

For \(\alpha \in (\alpha_K - \varepsilon, \alpha_K)\), with \(\varepsilon\) a sufficiently small positive real number, we let \(s := f_n(y)\). Since \(f_n(y)\) is a rational function \(y\), we can write \(y = h(s)\) for some continuous function \(h(s)\) in a small open neighborhood of \(f_n(y_0)\). Then, by changing variable we have \[F(\alpha)= i \int_{f_n(\overline{y_0})}^{f_n(y_0)} \log \left( \frac{s^2+ A^2}{(1+A^2) (g_n \circ h)(s)} \right) \frac{ds}{s^2-1}.\] As \(\alpha \to \alpha_K\), since \(f_n(\overline{y_0}) - f_n(y_0) \to 0\) we obtain \(F(\alpha) \to 0\).

Note that we also have \(\mathrm{Vol}(K(\alpha)) \to 0\) as \(\alpha \to \alpha_K\). Hence \(\mathrm{Vol}(K(\alpha)) = F(\alpha)\) if we can show that \[\frac{dF(\alpha)}{d\alpha} = \frac{d\mathrm{Vol}(K(\alpha))}{d\alpha} = - \frac{1}{2} l_\alpha.\].

By taking derivative of 2 and noting that \(dA/d\alpha=-(1+A^2)/2\), we have \[\begin{align} \frac{dF(\alpha)}{d\alpha} &=& \log \left( \frac{f^2_n(y_0)+ A^2}{(1+A^2)g_n(y_0)} \right) \frac{i f'_n(y_0)}{f^2_n(y_0)-1} \frac{dy_0}{d\alpha} - \log \left( \frac{f^2_n(\overline{y_0})+ A^2}{(1+A^2)g_n(\overline{y_0})} \right) \frac{i f'_n(\overline{y_0})}{f^2_n(\overline{y_0})-1} \frac{d\overline{y_0}}{d\alpha} \\ && + \, i \int_{\overline{y_0}}^{y_0} \frac{\partial}{\partial A} \left( \frac{f'_n(y)}{f^2_n(y)-1} \log \left( \frac{f^2_n(y)+ A^2}{(1+A^2)g_n(y)} \right) \right) \frac{dA}{d\alpha} \, dy \\ &=& i \int_{\overline{y_0}}^{y_0}\frac{f'_n(y)}{f^2_n(y)-1} \left( \frac{2A}{f^2_n(y)+ A^2} - \frac{2A}{1+ A^2} \right) \frac{-(1+A^2)}{2} dy \\ &=& i \int_{\overline{y_0}}^{y_0} \frac{f'_n(y) A}{f^2_n(y)+ A^2} dy \\ &=&i \left( \mathrm{arccot} \frac{f_n(\overline{y_0})}{A} - \mathrm{arccotn} \frac{f_n(y_0)}{A}\right). \end{align}\]

Since \(f_n(y_0) = i \coth \left( \frac{\gamma_\alpha + 4n i \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right)\) we have \(\frac{f_n(y_0)}{A} = i \coth \left( \frac{\gamma_\alpha + 4n i \alpha}{4} \right) = \cot \left( \frac{\gamma_\alpha + 4n i \alpha}{4i} \right)\) and \(\frac{f_n(\overline{y_0})}{A} = \overline{\frac{f_n(y_0)}{A} } = \cot \left( \frac{\overline{\gamma_\alpha + 4n i \alpha}}{-4i} \right)\). Hence \[\begin{align} \frac{dF(\alpha)}{d\alpha} &=& i \left( \mathrm{arccot} \frac{f_n(\overline{y_0})}{A} - \mathrm{arccotn} \frac{f_n(y_0)}{A}\right) \\ &=& i \left( \frac{\overline{\gamma_\alpha + 4n i \alpha} }{-4i} - \frac{\gamma_\alpha + 4n i \alpha}{4i} \right) \\ &=& -\frac{\overline{\gamma_\alpha} + \gamma_\alpha }{4} = - \frac{l_\alpha}{2}. \end{align}\] This proves Theorem 1 for \(C(2n, 3)\) in the hyperbolic case.

3.4.2 Spherical case↩︎

For \(\alpha \in (\alpha_K, 2\pi - \alpha_K)\), by the Schläfli formula we have \[\frac{d\mathrm{Vol}(K(\alpha))}{d\alpha} = \frac{1}{2} l_\alpha.\]

Let \[\label{G} G(\alpha)= \int_{y_+}^{y_-} \log \left( \frac{f^2_n(y)+ A^2}{(1+A^2)g_n(y)} \right) \frac{f'_n(y)dy}{f^2_n(y)-1}.\tag{3}\] Then Theorem 2 is equivalent to \(\mathrm{Vol}(K(\alpha)) = G(\alpha)\).

We first claim that \(G(\alpha) \to 0\) as \(\alpha \to \alpha_K\). Indeed, as \(\alpha \to \alpha_K\), we have \(l_\alpha \to 0\) and \[f_n(y_{\pm}) = \cot \left( \frac{\varphi_\alpha \pm l_\alpha+ 4n \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right) \to \cot \left( \frac{\varphi_{\alpha_K} + 4n \alpha_K}{4} \right) \cot \left( \frac{\alpha_K}{2} \right).\]

For \(\alpha \in (\alpha_K - \varepsilon, \alpha_K)\), with \(\varepsilon\) a sufficiently small positive real number, we let \(s := f_n(y)\). Since \(f_n(y)\) is a rational function \(y\), we can write \(y = h(s)\) for some continuous function \(h(s)\) in a small open neighborhood of \(f_n(y_{\pm})\). Then, by changing variable we have \[G(\alpha)= \int_{f_n(y_+)}^{f_n(y_-)} \log \left( \frac{s^2+ A^2}{(1+A^2) (g_n \circ h)(s)} \right) \frac{ds}{s^2-1}.\] As \(\alpha \to \alpha_K\), since \(f_n(y_{\pm}) \to \cot \left( \frac{\varphi_{\alpha_K} + 4n \alpha_K}{4} \right) \cot \left( \frac{\alpha_K}{2} \right)\), we obtain \(G(\alpha) \to 0\). Note that \(\mathrm{Vol}(K(\alpha)) \to 0\) as \(\alpha \to \alpha_K\). Hence \(\mathrm{Vol}(K(\alpha)) = G(\alpha)\) if we can show that \[\frac{dG(\alpha)}{d\alpha} = \frac{d\mathrm{Vol}(K(\alpha))}{d\alpha} = \frac{1}{2} l_\alpha.\]

By taking derivative of 3 and noting that \(dA/d\alpha=-(1+A^2)/2\), we have \[\begin{align} \frac{dG(\alpha)}{d\alpha} &=& \log \left( \frac{f^2_n(y_-)+ A^2}{(1+A^2)g_n(y_-)} \right) \frac{f'_n(y_-)}{f^2_n(y_-)-1} \frac{dy_-}{d\alpha} - \log \left( \frac{f^2_n(y_+)+ A^2}{(1+A^2)g_n(y_+)} \right) \frac{f'_n(y_+)}{f^2_n(y_+)-1} \frac{dy_+}{d\alpha} \\ && + \, \int_{y_+}^{y_-} \frac{\partial}{\partial A} \left( \frac{f'_n(y)}{f^2_n(y)-1} \log \left( \frac{f^2_n(y)+ A^2}{(1+A^2)g_n(y)} \right) \right) \frac{dA}{d\alpha} \, dy \\ &=& \int_{y_+}^{y_-}\frac{f'_n(y)}{f^2_n(y)-1} \left( \frac{2A}{f^2_n(y)+ A^2} - \frac{2A}{1+ A^2} \right) \frac{-(1+A^2)}{2} dy \\ &=& \int_{y_+}^{y_-} \frac{f'_n(y) A}{f^2_n(y)+ A^2} dy \\ &=& \mathrm{arccot} \frac{f_n(y_+)}{A} - \mathrm{arccot} \frac{f_n(y_-)}{A}. \end{align}\]

By the trigonometric identity (Proposition 7) we have \(f_n(y_{\pm}) = \cot \left( \frac{\varphi_\alpha \pm l_\alpha+ 4n \alpha}{4} \right) \cot \left( \frac{\alpha}{2} \right)\). This implies that \(\frac{f_n(y_{\pm})}{A} = \cot \left( \frac{\varphi_\alpha \pm l_\alpha+ 4n \alpha}{4} \right)\). Hence \[\begin{align} \frac{dG(\alpha)}{d\alpha} &=& - \mathrm{arccot} \frac{f_n(y_-)}{A} + \mathrm{arccot} \frac{f_n(y_+)}{A} \\ &=& - \frac{\varphi_\alpha - l_\alpha+ 4n \alpha}{4} +\frac{\varphi_\alpha + l_\alpha+ 4n \alpha}{4} \\ &=& \frac{l_\alpha}{2}. \end{align}\] This proves Theorem 2 for \(C(2n, 3)\) in the spherical case.

4 \(C(2n, 2p)\)↩︎

4.1 Knot group↩︎

Note that \(C(2n, 2p)\) is the double twist knot \(J(-2n, 2p)\), so by [8] its knot group has the following presentation.

Proposition 10. We have \(G(C(2n, 2p)) = \langle a, b \mid \omega' a = b \omega' \rangle\) where \[\omega' = [(a^{-1}b)^{n}(ab^{-1})^n]^p.\]

Figure 3: C(2n,2p)

We can also prove the above proposition by the same structure as Proposition 1 for \(C(2n, 2p + 1)\), with appropriate modifications for the even case. Starting from the knot diagram and using the Wirtinger presentation, we trace through the crossings to obtain the stated relation. See also [9].

4.2 \(\mathrm{SL}_2(\mathbb{C})\)-representations↩︎

Suppose \(\rho\colon G(C(2n, 2p))\to \mathrm{SL}_2(\mathbb{C})\) is a nonabelian representation. Up to conjugation, we may assume that \[A:=\rho(a) = \left[ \begin{array}{cc} m & 1\\ 0 & m^{-1} \end{array} \right] \quad \text{and} \quad B:=\rho(b) = \left[ \begin{array}{cc} m & 0\\ 2-z & m^{-1} \end{array} \right]\] where \((m,z) \in \mathbb{C}^2\) satisfies \(\rho(\omega' a) = \rho(b \omega')\). Note that \(z=\mathop{\mathrm{\mathrm tr}}\rho(ab^{-1})\).

The following propositions are proved in [9], [10].

Proposition 11. Let \(V = \rho((a^{-1} b)^{n} (ab^{-1})^{n})\) and \(v = \mathop{\mathrm{\mathrm tr}}V\). Then \[v = 2 + (z-2) (z+2-x^2) S^2_{n-1}(z).\]

Proposition 12. We have \[\rho(\omega' a) - \rho(b \omega') = \left[ \begin{array}{cc} 0 & \Phi_{C(2n, 2p)}(x,z)\\ (z-2) \Phi_{C(2n, 2p)}(x,z) & 0 \end{array} \right]\] where \[\Phi_{C(2n, 2p)}(x,z) := \left[ 1 + (z+2-x^2) S_{n-1}(z) (S_{n}(z) - S_{n-1}(z)) \right] S_{p-1}(v) - S_{p-2}(v).\]

4.3 Longitude and trignometric identity↩︎

If we choose the meridian to be \(\mu=a\) then the canonical longitude is \(\lambda = \omega' (\omega')^*\), where \((\omega')^*\) is the word obtained from \(\omega'\) by writing the letters in \(\omega'\) in reversed order.

Since \(\rho(\mu) = \left[ \begin{array}{cc} m & 1\\ 0 & m^{-1} \end{array} \right]\) we have \(\rho(\lambda) = \left[ \begin{array}{cc} l & *\\ 0 & l^{-1} \end{array} \right]\), where \(m = e^{i \alpha/2}\) and \(l = e^{\gamma_\alpha/2}\). By [8] we have \(l = - \frac{\widetilde{W'_{12}}}{W'_{12}}\), where \(W'_{12}\) is the \((1,2)\)-entry of \(W'=\rho(\omega')\) and \(\widetilde{W'_{12}}\) is obtained from \(W'_{12}\) by replacing \(m\) by \(m^{-1}\). Note that \(W'_{12}\) is a function in \(m\) and \(z\).

Similar to Propositions 5 and 6 we have the following propositions.

Proposition 13. We have \[W'_{12} = \left( m (S_{n}(z) - S_{n-1}(z)) - m^{-1} (S_{n-1}(z) - S_{n-2}(z) \right) S_{n-1}(z) S_{p-1}(v).\]

Proposition 14. Let \(f_n(z) = \frac{2S_n(z)-zS_{n-1}(z)}{(z-2)S_{n-1}(z)}\). Then \[f_n(z) = - \frac{l^{1/2} +l^{-1/2}}{l^{1/2}-l^{-1/2}} \cdot \frac{m+m^{-1}}{m-m^{-1}}.\]

4.4 Proof of Theorems 1 and 2 for \(C(2n,2)\)↩︎

Suppose \(p=1\). By Propositions 11 and 12 we have \(v = 2 + (z-2) (z+2-x^2) S^2_{n-1}(z)\) and \[\begin{align} \Phi_{C(2n, 2)}(x,z) &=& 1 + (z+2-x^2) S_{n-1}(z) (S_{n}(z) - S_{n-1}(z)). \end{align}\]

We write \(\Phi_{C(2n,2)}(x,z) = b - a x^2\), where \[\begin{align} a &=& S_{n-1}(z) (S_{n}(z) - S_{n-1}(z)), \\ b &=& 1 + (z+2) S_{n-1}(z) (S_{n}(z) - S_{n-1}(z)). \end{align}\] Since \(S^2_n(z) - z S_n(z) S_{n-1}(z) + S^2_{n-1}(z) = 1\) we have \[\begin{align} b - 4a &=& 1 + (z-2) S_{n-1}(z) (S_{n}(z) - S_{n-1}(z)) \\ &=& S^2_{n}(z) - 2 S_{n}(z) S_{n-1}(z) + (3-z)S^2_{n-1}(z). \end{align}\]

Choose \(c = f_n(z)\). By Lemma 3, the equation \(\Phi_{C(2n, 2)}(2\cos \frac{\alpha}{2}, z)=0\) is equivalent to \(c^2+ A^2= (1+A^2)d\) where \(d = 1 + (c^2-1)(1 - \frac{b}{4a})\). Since \[c^2 -1 = \frac{4(S_n(z)-S_{n-1}(z)) (S_n(z) - (z-1) S_{n-1}(z))}{(z-2)^2 S^2_{n-1}(z)},\] by a direct calculation we have \[\begin{align} d &=& 1 + (c^2-1) \left( 1 - \frac{b}{4a} \right) \\ &=& 1 - \frac{\left( S_n(z) - (z-1) S_{n-1}(z) \right) \left( S^2_{n}(z) - 2 S_{n}(z) S_{n-1}(z) + (3-z)S^2_{n-1}(z) \right)}{(z-2)^2 S^3_{n-1}(z)} \\ &=& - \frac{\left( S_n(z) -S_{n-1}(z) \right)\left( S^2_n(z) + S^2_{n-1}(z) - z S_n(z) S_{n-1}(z) \right)}{(z-2)^2 S^3_{n-1}(z)} \\ &=& - \frac{ S_n(z) -S_{n-1}(z)}{(z-2)^2 S^3_{n-1}(z)}. \end{align}\]

Hence we have proved the following.

Proposition 15. Let \[f_n(z) = \frac{2S_n(z)-zS_{n-1}(z)}{(z-2)S_{n-1}(z)}, \qquad g_n(z) = - \frac{(S_n(z) - S_{n-1}(z))^2}{(z-2)^3S^4_{n-1}(z)}.\] Let \(x = 2\cos \frac{\alpha}{2}\) and \(A = \cot \frac{\alpha}{2}\). Then the equation \(\Phi_{C(2n, 2)}(x,z) = 0\) is equivalent to \(f^2_n(z)+ A^2= (1+A^2)g_n(z)\).

By using the trigonometric identity (Proposition 14) and Proposition 15, the proof of Theorems 1 and 2 for \(C(2n,2)\) is similar to that for \(C(2n,3)\).

4.5 Proof of Theorems 1 and 2 for \(C(2n, -2n)\)↩︎

Suppose \(p=-2n\). Note that \(C(2n, -2n)\) is the mirror image of the double twist knot \(J(2n, 2n)\) in [8]. By [9] the component of \(\Phi_{C(2n, -2n)}(x,z)\) containing the holonomy representation is a factor of \(v - z = (z-2) \left( -1 + (z+2-x^2) S^2_{n-1}(z) \right)\). The factor \(z-2\) corresponds to reducible representations, hence the factor \[\Phi^{\mathrm{hol}}_{C(2n, -2n)}(x,z) := -1 + (z+2-x^2) S^2_{n-1}(z)\] determines the component containing the holonomy representation.

We write \(\Phi^{\mathrm{hol}}_{C(2n, -2n)}(x,z) = b - a x^2\), where \(a = S^2_{n-1}(z)\) and \(b = -1 + (z+2) S^2_{n-1}(z)\). Since \(S^2_n(z) - z S_n(z) S_{n-1}(z) + S^2_{n-1}(z) = 1\) we have \[b - 4a = - 1 +(z-2) S^2_{n-1}(z) = - S^2_{n}(z) + z S_{n}(z) S_{n-1}(z) + (z-3)S_{n-1}(z).\]

Choose \(c = f_n(z)\). By Lemma 3, the equation \(\Phi^{\mathrm{hol}}_{C(2n, -2n)}(2\cos \frac{\alpha}{2}, z)=0\) is equivalent to \(c^2+ A^2= (1+A^2)d\) where \(d = 1 + (c^2-1)(1 - \frac{b}{4a})\). Since \[c^2 -1 = \frac{4(S_n(z)-S_{n-1}(z)) (S_n(z) - (z-1) S_{n-1}(z))}{(z-2)^2 S^2_{n-1}(z)},\] by a direct calculation we have \[\begin{align} d &=& 1 + (c^2-1) \left( 1 - \frac{b}{4a} \right) \\ &=& 1 - \frac{(S_n(z)-S_{n-1}(z)) (S_n(z) - (z-1) S_{n-1}(z)) }{(z-2)^2 S^4_{n-1}(z)} \\ && \qquad \times \left( - S^2_{n}(z) + z S_{n}(z) S_{n-1}(z) + (z-3)S_{n-1}(z) \right) \\ &=& \frac{\left( S^2_n(z) + S^2_{n-1}(z) - z S_n(z) S_{n-1}(z) \right)^2}{(z-2)^2 S^4_{n-1}(z)} \\ &=& \frac{1}{(z-2)^2 S^4_{n-1}(z)}. \end{align}\]

Hence we have proved the following.

Proposition 16. Let \[f_n(z) = \frac{2S_n(z)-zS_{n-1}(z)}{(z-2)S_{n-1}(z)}, \qquad g_n(z) = \frac{1}{(z-2)^2S^4_{n-1}(z)}.\] Let \(x = 2\cos \frac{\alpha}{2}\) and \(A = \cot \frac{\alpha}{2}\). Then the equation \(\Phi^{\mathrm{hol}}_{C(2n, -2n)}(x,z) = 0\) is equivalent to \(f^2_n(z)+ A^2= (1+A^2)g_n(z)\).

By using the trigonometric identity (Proposition 14) and Proposition 16, the proof of Theorems 1 and 2 for \(C(2n,-2n)\) is similar to that for \(C(2n,3)\).

Acknowledgements↩︎

The first author has been supported by a grant from the Simons Foundation (#708778).

References↩︎

[1]
S. Kojima, Deformations of hyperbolic 3-cone-manifolds, J. Differential Geom. 49(1998) 469–516.
[2]
J. Porti, Spherical cone structures on 2-bridge knots and links, Kobe J. Math. 21(2004) 61–70.
[3]
H. Hilden, M. Lozano, and J. Montesinos-Amilibia, Volumes and Chern-Simons invariants of cyclic coverings over rational knots, in Topology and Teichmüller spaces (Katinkulta, 1995), pages 31–55. World Sci. Publ., River Edge, NJ, 1996.
[4]
J. Ham, A. Mednykh, and V. Petrov, Identities and volumes of the hyperbolic twist knot cone-manifolds, J. Knot Theory Ramifications 23 (2014) 1450064.
[5]
J. Ham and J. Lee, The volume of hyperbolic cone-manifolds of the knot with Conway’s notation \(C(2n,3)\), J. Knot Theory Ramifications 25(2016) 1650030.
[6]
A. Tran, Volumes of hyperbolic double twist knot cone-manifolds, J. Knot Theory Ramifications 26(2017), no. 11, 1750068, 14 pp.
[7]
A. Mednykh, Volumes of two-bridge cone manifolds in spaces of constant curvature, Transform. Groups 26(2021), no.2, 601–629.
[8]
J. Hoste and P. Shanahan, A formula for the A-polynomial of twist knots, J. Knot Theory Ramifications 13(2004), no. 2, 193–209.
[9]
M. Macasieb, K. Petersen and R. van Luijk, On character varieties of two-bridge knot groups Proc. Lond. Math. Soc. (3) 103(2011), no. 3, 473–507.
[10]
T. Morifuji and A. Tran, Twisted Alexander polynomials of two-bridge knots for parabolic representations, Pacific J. Math. 269(2014), no. 2, 433–451.
[11]
R. Riley, Nonabelian representations of 2-bridge knot groups, Quart. J. Math. Oxford Ser. (2) 35(1984), 191–208.

  1. 2020 Mathematics Subject Classification. Primary 57K10, 57K32.↩︎

  2. Key words and phrases. cone-manifold, double twist knot, nonabelian representation, Riley polynomial, two-bridge knot, volume.↩︎