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

Quandles over a hyperboloid of one sheet and the longitudinal mapping knot invariant for SL(2,)SL(2,\mathbb{R})

Kentaro Yonemura the 2010 MSC: 57K10, 57K12, E-mail: [email protected]
Abstract

This paper aims to consider algebraic structures of quandles defined over a hyperboloid of one sheet and compute the related longitudinal mapping for SL(2,)SL(2,\mathbb{R}).

keywords:
quandles, the longitudinal mapping, a hyperboloid of one sheet

1 Introduction

A quandle is an algebraic system defined by Joyce [7] and Matveev [9] independently. Joyce and Matveev was motivated by knot theory and constructed the almost complete knot invariant called the knot quandles or the fundamental quandles of knots. In this paper, we deal with a smooth quandle, which is a differentiable manifold with a smooth quandle-operation, defined over a hyperboloid of one sheet. See Ishikawa [6] and Nosaka [12] for more details on a smooth quandle.

We deal with two problems in this paper.

First, we see that subquandles of conjugacy quandles are generally different from the quandle composed by the Azcan-Fenn [1], even if both of them have the same topological structure. We consider this issue to show that a fact is a special case: Azcan-Fenn [1] defined the spherical quandle using Euclidean inner product. On the other hand, Clark-Saito [4] defined a family of quandles on conjugacy classes of SU(2)SU(2) and called them spherical quandles. The auther [14] proved that the two types of spherical quandles are compatible, that is, the spherical quandle defined by Azcan-Fenn is isomorphic to one of the spherical quandle defined by Clark-Saito.

They defined a knot invariant called a longitudinal mapping and calculate it in the case of SU(2)SU(2) using the quandle.

Secondly, we determine the value of the longitudinal mapping Gx(K)\mathcal{L}_{G}^{x}(K) under limited condition: When GG is SL(2,)SL(2,\mathbb{R}), KK is a (2,n)(2,n)-torus knot, and xx is conjugate with

(er00er)SL(2,).\begin{pmatrix}e^{r}&0\\ 0&e^{-r}\end{pmatrix}\in SL(2,\mathbb{R}).

The longitudinal mapping is a knot invariant defined by Clark-Saito [4]. Clark and Saito explained that their invariant is a extention of the knot colouring polynomial [5], which is a generalisation of the quandle cocycle invariant [3]. Our approach is a little different from Clark-Saito [4]. Clark-Saito [4] presented the elemtnts of SU(2)SU(2) as unitquaternions and used Python to determine the value of the longitudinal mapping for SU(2)SU(2). On the other hand, we present the elements of SL(2,)SL(2,\mathbb{R}) as matrices and use linear algebra to determine the value of the longitudinal mapping for SL(2,)SL(2,\mathbb{R}). Our approach may be a useful example when considering the value of the longitudinal mapping for highter dimentional Lie groups.

This paper is organized as follows. In section 2, the basic notation and facts on quandles and SL(2,)SL(2,\mathbb{R}) are presented. In section 3, we discuss on the algebraic structure of quandles over a hyperboloid of one sheet. In section 4, we determine XX-colorings with respect to a diagram of (2,n)(2,n)-torus knots for the case XX is a subquandle of conjugacy quandles. In the section 5, we apply the argument considered in section 4. In the section 6, we introduce a result of Nosaka [10]. In the section 7,we apply the result introduced in section 6. In section 8, we determine the value of the longitudinal mapping for SL(2,)SL(2,\mathbb{R}).

2 Preliminaries

We recall definitions and facts using in this paper without proofs.

2.1 Quandle

We see the definition of a quandle and some basic facts. See Kamada [8] and Nosaka [11] for more details.

Definition 2.1 (Joyce [7], Matveev [9]).

A quandle is a set XX with a binary operation :X×XX\triangleright:X\times X\to X satisfying the three conditions:

(Q1) xx=xx\triangleright x=x for any xXx\in X.

(Q2) The map Sy:XXS_{y}:X\to X defined by xxyx\mapsto x\triangleright y is bijective for any yXy\in X.

(Q3) (xy)z=(xz)(yz)(x\triangleright y)\triangleright z=(x\triangleright z)\triangleright(y\triangleright z) for any x,y,zXx,y,z\in X.

A subset YY of quandle XX is said to be a subquandle if the quandle operation of XX is closed in YY. A map f:XYf:X\to Y between quandles is said to be a quandle homomorphism if f(xy)=f(x)f(y)f(x\triangleright y)=f(x)\triangleright f(y) for any x,yXx,y\in X. A quandle homomorphism is said to be a quandle isomorphism if it is bijective.

We see examples of quandles.

Example 2.2.

Suppose XX is a set and xy=xx\triangleright y=x for any x,yXx,y\in X. Then (X,)(X,\triangleright) is a quandle called a trivial quandle.

Example 2.3.

Suppose X=/nX=\mathbb{Z}/n\mathbb{Z} and xy=2yxx\triangleright y=2y-x for any x,yXx,y\in X. Then (X,)(X,\triangleright) is a quandle called a dihedral quandle [13].

Example 2.4.

Let GG be a group and \triangleright be a binary operation of GG defined by gh=h1ghg\triangleright h=h^{-1}gh for any g,hGg,h\in G. Then (G,)(G,\triangleright) is a quandle called conjugacy quandle. We denote this quandle Conj(G)\operatorname{Conj}(G).

Example 2.5.

Let GG be a group, XX a set on which GG acts from the right, and κ:XG\kappa:X\to G a map satisfying the two conditions:

  1. 1.

    κ(xg)=g1κ(x)g\kappa(x\cdot g)=g^{-1}\kappa(x)g for any xXx\in X and gGg\in G.

  2. 2.

    xκ(x)=xx\cdot\kappa(x)=x for any xXx\in X.

Suppose xy=xκ(y)x\triangleright y=x\cdot\kappa(y) for any x,yXx,y\in X. Then (X,)(X,\triangleright) is a quandle called an augmented quandle [7]. We denote this quandle by (X,G,κ)(X,G,\kappa) or simply XX. The quandle XX is said to be faithful if the map κ\kappa is injective.

Example 2.6.

Suppose GG be a group, XX be a subquandle of Conj(G)\operatorname{Conj}(G), and iX:XGi_{X}:X\to G be an inclusion map. Then GG acts on XX via conjugation and (X,G,iX)(X,G,i_{X}) is a faithfull augumented quandle. The identity map idX:XX\operatorname{id}_{X}:X\to X induces a quandle isomorphism from XX to (X,G,iX)(X,G,i_{X}).

Example 2.7.

Let KK be a tame knot in the 3-sphere, πK=π1(S3K)\pi_{K}=\pi_{1}(S^{3}\setminus K) the knot group of KK, and H=𝔪,𝔩H=\langle\mathfrak{m},\mathfrak{l\rangle} the subgroup of πK\pi_{K} generated by a meridian 𝔪πK\mathfrak{m}\in\pi_{K} and the preferred longitude 𝔩πK\mathfrak{l}\in\pi_{K} of KK. Suppose X=H\πKX=H\backslash\pi_{K} and HxHy=H𝔪xy1𝔪yHx\triangleright Hy=H\mathfrak{m}xy^{-1}\mathfrak{m}y for x,yπKx,y\in\pi_{K}. Then, the algebraic system (X,)(X,\triangleright) is a quandle and called the knot quandle of KK or the fundamental quandle of KK [7, 9]. We denote this quandle by QKQ_{K}.

The knot quandle, we defined in Example 2.7, is a complete knot invariant. See [7, 9, 8, 11] for more details.

A quandle is said to be an involutory quandle or a kei if (xy)y=x(x\triangleright y)\triangleright y=x for any x,yx,y. A dihedral quandle is an involutory quandle for example.

We introduce an XX-coloring and provide some examples.

Definition 2.8 (XX-colorings).

Let XX be a quandle, and DD be an oriented knot diagram. An XX-coloring of DD is a map C:{arcs of D}XC:\{\mbox{arcs of }D\}\to X satisfying the condition C(ατ)C(βτ)=C(γτ)C(\alpha_{\tau})\triangleright C(\beta_{\tau})=C(\gamma_{\tau}) in Fig. 1 at each crossing τ\tau of DD. We denote the set of XX-colorings of DD by ColX(D)\operatorname{Col}_{X}(D).

For a knot KK with a diagram DD and a quandle XX, it is known that an XX-coloring of DD induces a quandle homomorphism from QKQ_{K} to XX.

Refer to caption
Figure 1: Crossings τ\tau of a knot diagram DD

We see some special quandle colorings.

Example 2.9.

Suppose DD be an oriented knot diagram and Y={}Y=\{*\} be a trivial quandle. Then, any map C:{arcs of D}YC:\{\mbox{arcs of }D\}\to Y is a YY-colorings.

Definition 2.10.

In this paper, an XX-coloring CC is said to be trivial if the image of CC is a trivial subquandle of XX.

Remark 2.11.

Some literature defines a trivial coloring as a coloring considered in Example 2.9.

2.2 Properties of SL(2,)SL(2,\mathbb{R})

Consider the Lie group

SL(2,)={(abcd):a,b,c,d,adbc=1}.SL(2,\mathbb{R})=\left\{\begin{pmatrix}a&&b\\ c&&d\end{pmatrix}\ :\ {}a,b,c,d\in\mathbb{R},\ ad-bc=1\right\}.

The Lie group SL(2,)SL(2,\mathbb{R}) acts on itself via conjugation. Let S12(r)S^{2}_{1}(r) be a conjugacy class of SL(2,)SL(2,\mathbb{R})

S12(r)={g1(er00er)gSL(2,):gSL(2,)}S^{2}_{1}(r)=\left\{g^{-1}\begin{pmatrix}e^{r}&&0\\ 0&&e^{-r}\end{pmatrix}g\in{SL}(2,\mathbb{R})\ :\ {}g\in{SL}(2,\mathbb{R})\right\}

for r>0r>0. The conjugacy class S12(r)S^{2}_{1}(r) is diffeomorfic to a hyperboloid of one sheet. We recall the following known facts.

Proposition 2.12.

For any r>0r>0, S12(r)S^{2}_{1}(r) is an SL(2,)SL(2,\mathbb{R})-orbit.

Proposition 2.13.

For any r>0r>0,

S12(r)={gSL(2,):traceg=2coshr}.S^{2}_{1}(r)=\{g\in SL(2,\mathbb{R})\ :\ {}\operatorname{trace}g=2\cosh{r}\}.
Proposition 2.14.

Let rr be positive, and

D(r)=(er00er)S12(r).D(r)=\begin{pmatrix}e^{r}&&0\\ 0&&e^{-r}\end{pmatrix}\in S^{2}_{1}(r).

Then, both of the isotropy subgroup of SL(2,)SL(2,\mathbb{R}) with respect to D(r)SL(2,)D(r)\in SL(2,\mathbb{R}) is the subgroup consisting of the entire diagonal matrices of SL(2,)SL(2,\mathbb{R}).

3 Quandles over a hyperboloid of one sheet

We consider two types of quandles S12(r)S^{2}_{1}(r) and S12{S^{2}_{1}}_{\mathbb{R}} over a hyperboloid of one sheet. We see that they are not isomorphic though one of the spherical quandle defined in Clark-Saito [4] , which is similar to S12(r)S^{2}_{1}(r), is isomorphic to the spherical quandle defined by Azcan-Fenn [1], which is similar to S12{S^{2}_{1}}_{\mathbb{R}}.

First of all, we see a quandle S12(r)S^{2}_{1}(r) and its property.

Proposition 3.1.

For r>0r>0, the conjugacy class S12(r)S^{2}_{1}(r) is a subquandle of Conj(SL(2,))\operatorname{Conj}(SL(2,\mathbb{R})).

Proof.

It is easy to see the result in light of Proposition 2.12. ∎

Secondly, we see the definition of a quandle S12{S^{2}_{1}}_{\mathbb{R}}. Let ,:3×3\langle-,-\rangle:\mathbb{R}^{3}\times\mathbb{R}^{3}\to\mathbb{R} be a bilinear map defined by

(x0,x1,x2),(y0,y1,y2)=x0y0+x1y1+x2y2,\langle(x_{0},x_{1},x_{2}),(y_{0},y_{1},y_{2})\rangle=-x_{0}y_{0}+x_{1}y_{1}+x_{2}y_{2},

S12={𝒙3:𝒙,𝒙=1}S^{2}_{1}=\{\bm{x}\in\mathbb{R}^{3}\ :\ \langle\bm{x},\bm{x}\rangle=1\} a hyperboloid of one sheet, and :S12×S12S12\triangleright:S^{2}_{1}\times S^{2}_{1}\to S^{2}_{1} a binary operation defined by 𝒙𝒚=2𝒙,𝒚𝒚𝒙\bm{x}\triangleright\bm{y}=2\langle\bm{x},\bm{y}\rangle\bm{y}-\bm{x} for all 𝒙,𝒚S12\bm{x},\bm{y}\in S^{2}_{1}.

Proposition 3.2 (Azcan-Fenn [1]).

The algebraic system (S12,)(S^{2}_{1},\triangleright) is an involutory quandle. We denote this quandle as S12{S^{2}_{1}}_{\mathbb{R}}.

Finally, we see that S12{S^{2}_{1}}_{\mathbb{R}} is different from S12(r){S^{2}_{1}}(r) for any r>0r>0.

Lemma 3.3.

For r>0r>0, the quandle S12(r)S^{2}_{1}(r) is not an involutory quandle.

Proof.

We prove by contradiction. Assume S12(r)S^{2}_{1}(r) be an ivolutory quandle. For D(r)D(r), which is defined in Proposition 2.14, and any yS12(r)y\in S^{2}_{1}(r),

(D(r)y)y=D(r),(D(r)\triangleright y)\triangleright y=D(r),

that is, D(r)y2=y2D(r)D(r)y^{2}=y^{2}D(r). By Proposition 2.14,

S12(r)={D(r)±1}.S^{2}_{1}(r)=\{D(r)^{\pm 1}\}.

This is contradiction since S12(r)S^{2}_{1}(r) is not finite. ∎

Proposition 3.4.

For r>0r>0, S12(r)S^{2}_{1}(r) is not isomorphic to S12{S^{2}_{1}}_{\mathbb{R}}.

Proof.

The quandle S12(r)S^{2}_{1}(r) is not an involutory quandle because of Lemma 3.3. On the other hand, the quandle S12{S^{2}_{1}}_{\mathbb{R}} is an involutory quandle. This is a contradiction. ∎

By Proposition 3.4, subquandles of conjugacy quandles are generally different from the quandle composed by the Azcan-Fenn [1], even if both of them have the same topological structure.

4 Quandle Colorings of (2,n)(2,n)-torus knots

In this section, we discuss a coloring of (2,n)(2,n)-torus knots with respect to subquandles of conjugacy quadles.

Let GG be a group, XX be a subquandle of Conj(G)\operatorname{Conj}(G), and DD be the diagram of (2,n)(2,n)-torus knots as shown in Fig. 2. We consider XX-colorings of the diagram DD. By the definition of the quandle coloring,

ColX(D)={C:{α0,,αn1}X:i=0,1,,nC(αi+2)=C(αi+1)1C(αi)C(αi+1)}.\operatorname{Col}_{X}(D)=\left\{C:\{\alpha_{0},\cdots,\alpha_{n-1}\}\to X:\begin{array}[]{rcl}{}^{\forall}i&=&0,1,\cdots,n\\ C(\alpha_{i+2})&=&C(\alpha_{i+1})^{-1}C(\alpha_{i})C(\alpha_{i+1})\end{array}\right\}.
Lemma 4.1.

For j=0,1,,n12j=0,1,\cdots,\tfrac{n-1}{2} and CColX(D)C\in\operatorname{Col}_{X}(D),

C(α2j)=(C(α0)C(α1))jC(α0)(C(α0)C(α1))jC(α2j+1)=(C(α0)C(α1))jC(α1)(C(α0)C(α1))j.\begin{array}[]{ccc}C(\alpha_{2j})&=&\big{(}C(\alpha_{0})C(\alpha_{1})\big{)}^{-j}C(\alpha_{0})\big{(}C(\alpha_{0})C(\alpha_{1})\big{)}^{j}\\ C(\alpha_{2j+1})&=&(C(\alpha_{0})C(\alpha_{1}))^{-j}C(\alpha_{1})(C(\alpha_{0})C(\alpha_{1}))^{j}.\end{array}
Proof.

The result follows by induction on jj. ∎

For x,yXx,y\in X, we define a map Cx,y:{α0,,αn1}XC_{x,y}:\{\alpha_{0},\cdots,\alpha_{n-1}\}\to X as

Cx,y(αm)={(xy)jx(xy)jifm=2j(xy)jy(xy)jifm=2j+1.C_{x,y}(\alpha_{m})=\left\{\begin{array}[]{ccl}(xy)^{-j}x(xy)^{j}&\mbox{if}&m=2j\\ (xy)^{-j}y(xy)^{j}&\mbox{if}&m=2j+1\end{array}\right..

The set ColX(D)\operatorname{Col}_{X}(D) is presented as follows.

Proposition 4.2.

Suppose k1k\geq 1 and n=2k+1n=2k+1,

ColX(D)={Cx,y:{α0,,αn1}X:x,yX s.t. (xy)kx=y(xy)k}.\operatorname{Col}_{X}(D)=\left\{C_{x,y}:\{\alpha_{0},\cdots,\alpha_{n-1}\}\to X\ :\ {}x,y\in X\mbox{ s.t. }(xy)^{k}x=y(xy)^{k}\right\}.
Proof.

The result follows from Lemma 4.1 and α0=αn=α2k+1\alpha_{0}=\alpha_{n}=\alpha_{2k+1}. ∎

Refer to caption
Figure 2: The diagram of (2,n)(2,n)-torus knots and its arcs α0,,αn1\alpha_{0},\cdots,\alpha_{n-1}.

5 S12(r)S^{2}_{1}(r)-Colorings of (2,n)(2,n)-torus knots

Suppose r>0r>0 and DD be the diagram of the (2,n)(2,n)-torus knot as shown in Fig. 2. We determine S12(r)S^{2}_{1}(r)-colorings of DD.

For an inner automorphism ρInnSL(2,)\rho\in\operatorname{Inn}SL(2,\mathbb{R}) induces a quandle automorphism of S12(r)S^{2}_{1}(r). We identify ρ\rho and the induced quandle automorphism in this section. The fact induces the action of InnSL(2,)\operatorname{Inn}SL(2,\mathbb{R}) on ColS12(r)(D)\operatorname{Col}_{S^{2}_{1}(r)}(D).

By Proposition 4.2, we find pairs x,yS12(r)x,y\in S^{2}_{1}(r) satisfying

(xy)kx=y(xy)k,(xy)^{k}x=y(xy)^{k}, (5.1)

where k=n12k=\tfrac{n-1}{2}, to determine the S12(r)S^{2}_{1}(r)-colorings. Considering the action of InnSL(2,)\operatorname{Inn}SL(2,\mathbb{R}), it does not lose its generality as

x=D(r)=(er00er),y=(abcd).x=D(r)=\begin{pmatrix}e^{r}&0\\ 0&e^{-r}\end{pmatrix},\ {}y=\begin{pmatrix}a&b\\ c&d\end{pmatrix}.

However, by Proposition 2.13, let aa, bb, cc, and dd be real numbers that satisfy the following conditions: adbc=1ad-bc=1 and a+d=coshra+d=\cosh{r}.

Let λ\lambda, μ\mu\in\mathbb{C} be the two solutions of the characteristic equation of yy

t2(aer+der)t+1=0.t^{2}-(ae^{r}+de^{-r})t+1=0.

We are able to determine x,yx,y satisfying equation (5.1) in light of the argument in section C.

(In the case b=0b=0 or c=0c=0) In light of lemma B.2 and Lemma C.2, the real numbers aa, bb, cc, and dd satisfying equation (5.1) are

(a,b,c,d)=(er,0,0,er).(a,b,c,d)=(e^{r},0,0,e^{-r}).

Then, xx and yy induces a trivial coloring in the sense of Example 2.9. We denote the S12S^{2}_{1}-coloring as C0C_{0}.

(In the case b0b\neq 0 and c0c\neq 0 and λμ\lambda\neq\mu) In light of Lemma C.3, the real numbers aa, bb, cc, and dd satisfying equation (5.1) satisfy

(a,d)\displaystyle(a,d) =\displaystyle= (ercoshr+cos2θjsinhr,ercoshrcos2θjsinhr),\displaystyle\left(\frac{-e^{-r}\cosh{r}+\cos{2\theta_{j}}}{\sinh{r}},\frac{e^{r}\cosh{r}-\cos{2\theta_{j}}}{\sinh{r}}\right),
bc\displaystyle bc =\displaystyle= 4sin2θj(sin2θj+sinh2r)sinh2r,\displaystyle-\frac{4\sin^{2}{\theta_{j}}(\sin^{2}{\theta_{j}}+\sinh^{2}{r})}{\sinh^{2}{r}},

where θj=πj2n\theta_{j}=\tfrac{\pi j}{2n} and j=1,3,,n12j=1,3,\cdots,\frac{n-1}{2}. We denote the S12(r)S^{2}_{1}(r)-coloring induced by x,yx,y as Cj,b,cC_{j,b,c}.

(In the case b0b\neq 0 and c0c\neq 0 and λ=μ\lambda=\mu) There is no aa, bb, cc, or dd satisfying equation in (5.1) light of Lemma C.4.

We summarize the discussion in this section like the following theorem.

Theorem 5.1.

Suppose DD be the diagram of (2,n)(2,n)-torus knots as shown in Fig. 2. Then,

ColS12(r)(D)=ρInnSL(2,){ρC0}{ρCj,b,c:j=1,3,,n2,b,cbc=4sin2θj(sin2θj+sinh2r)sinh2r}.\displaystyle\operatorname{Col}_{S^{2}_{1}(r)}(D)=\bigcup_{\rho\in\operatorname{Inn}SL(2,\mathbb{R})}\left\{\rho\circ C_{0}\right\}\cup\left\{\rho\circ C_{j,b,c}\ :\ {}\begin{matrix}j=1,3,\cdots,n-2,\ b,c\in\mathbb{R}\\ bc=-\frac{4\sin^{2}{\theta_{j}}(\sin^{2}{\theta_{j}}+\sinh^{2}{r})}{\sinh^{2}{r}}\end{matrix}\right\}.

Finally, we end this section by preparing a lemma used in section 8.

Lemma 5.2.

Suppose α0\alpha_{0}, α1\alpha_{1} are arcs of the diagram of (2,n)(2,n)-torus knots as shown in Fig. 2. Then

(Cj,b,c(α0)Cj,b,c(α1))n=(1001).(C_{j,b,c}(\alpha_{0})C_{j,b,c}(\alpha_{1}))^{n}=\begin{pmatrix}-1&0\\ 0&-1\end{pmatrix}.
Proof.

By Lemma C.1,

(Cj,b,c(α0)Cj,b,c(α1))n=\displaystyle(C_{j,b,c}(\alpha_{0})C_{j,b,c}(\alpha_{1}))^{n}=
1λμ(λn1+μn1+aer(λnμn)ber(λnμn)cer(λnμn)λn+1μn+1aer(λnμn)).\displaystyle\frac{1}{\lambda-\mu}\begin{pmatrix}-\lambda^{n-1}+\mu^{n-1}+ae^{r}(\lambda^{n}-\mu^{n})&be^{r}(\lambda^{n}-\mu^{n})\\ ce^{-r}(\lambda^{n}-\mu^{n})&\lambda^{n+1}-\mu^{n+1}-ae^{r}(\lambda^{n}-\mu^{n})\end{pmatrix}.

By Lemma B.1,

λnμnλμ=sinnθjsinθj=0\frac{\lambda^{n}-\mu^{n}}{\lambda-\mu}=\frac{\sin{n\theta_{j}}}{\sin{\theta_{j}}}=0

and

λn±1μn±1λμ=sin(n±1)θjsinθj=1.\frac{\lambda^{n\pm 1}-\mu^{n\pm 1}}{\lambda-\mu}=\frac{\sin{(n\pm 1)\theta_{j}}}{\sin{\theta_{j}}}=\mp 1.

Thus the result follows. ∎

6 Quandle colorings and representation of knot groups

We introduce Nosaka’s work to see the one-to-one correspondence between a quandle coloring and a representation of knot groups. See Nosaka [10, 11] for more details.

Let KK be a tame knot in the 3-sphere S3S^{3} with a diagram DD, π1(S3K)\pi_{1}(S^{3}\setminus K) the knot group of KK and QKQ_{K} the knot quandle. For any augmented quandle (X,G,κ)(X,G,\kappa), we define a set

R(K,G)={fHom(π1(S3K),G):xX,f(𝔪)=κ(x)}.R(K,G)=\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),G)\ :\ {}{}^{\exists}x\in X,\ f(\mathfrak{m})=\kappa(x)\}.
Theorem 6.1 (Nosaka [10]).

Let (X,G,κ)(X,G,\kappa) be a faithful augmented quandle. Then, there is a bijection Ψ:ColX(D)R(K,G).\Psi:\operatorname{Col}_{X}(D)\xrightarrow{\sim}R(K,G).

The bijection Ψ\Psi is given as follows. It is known that π1(S3K)\pi_{1}(S^{3}\setminus K) is generated by the elements corresponding to the arcs of DD (Wirtinger presentation, see [2]). For any XX-coloring CC, ΨC:π1(S3K)G\Psi C:\pi_{1}(S^{3}\setminus K)\to G is a group homomorphism satisfying Ψf(α)=κf(α)\Psi f(\alpha)=\kappa\circ f(\alpha) for any α\alpha of π1(S3K)\pi_{1}(S^{3}\setminus K) corresponding to an arc of DD.

We give a few facts about the bijection Ψ\Psi.

Lemma 6.2.
  1. 1.

    The action of GG on XX induces the action of GG on ColX(D)\operatorname{Col}_{X}(D). XX-colorings C1,C2:QKXC_{1},C_{2}:Q_{K}\to X are in the same GG-orbit if and only if ΨC1\Psi C_{1} and ΨC2\Psi C_{2} are conjugate.

  2. 2.

    An XX-coloring CC is trivial in the sense of Definition 2.10 if and only if ΨC\Psi C is an abelian representation.

Corollary 6.3.

Suppose α0\alpha_{0} is an arc of DD, xXx\in X, and 𝔪\mathfrak{m} is equal to α0\alpha_{0} as an element of the knot group π1(S3K)\pi_{1}(S^{3}\setminus K). Then the bijection Ψ\Psi induces the one-to-one correspondence

{fHom(π1(S3K),G):f(𝔪)=κ(x)}{CColX(D):C(α0)=x}.\displaystyle\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),G)\ :\ {}f(\mathfrak{m})=\kappa(x)\}\simeq\{C\in\operatorname{Col}_{X}(D)\ :\ {}C(\alpha_{0})=x\}.

7 S12(r)S^{2}_{1}(r)-colorings and hyperbolic representation of knot groups

Suppose DD is the diagram of (2,n)(2,n)-torus knots as shown in Fig. 2.

Lemma 7.1.

Let KK be a knot and 𝔪π1(S3K)\mathfrak{m}\in\pi_{1}(S^{3}\setminus K) be a meridian. Then,

{fHom(π1(S3K),SL(2,)):xS12(r),f(𝔪)=x}\displaystyle\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),SL(2,\mathbb{R}))\ :\ {}{}^{\exists}x\in S^{2}_{1}(r),\ f(\mathfrak{m})=x\}
=\displaystyle= {fHom(π1(S3K),SL(2,)):tracef(𝔪)=2coshr}.\displaystyle\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),SL(2,\mathbb{R}))\ :\ {}\operatorname{trace}f(\mathfrak{m})=2\cosh{r}\}.
Proof.

By Proposition 2.13, this lemma is true. ∎

Proposition 7.2.

Suppose KK be a knot with a diagram DD, and r>0r>0. Then, there is a bijection

ΨK,r:ColS12(r)(D){fHom(π1(S3K),SL(2,)):tracef(𝔪)=2coshr}\Psi_{K,r}:\operatorname{Col}_{S_{1}^{2}(r)}(D)\xrightarrow{\sim}\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),SL(2,\mathbb{R}))\ :\ {}\operatorname{trace}f(\mathfrak{m})=2\cosh r\}
Proof.

The result follows from Example 2.6, Theorem 6.1, and Lemma 7.1. ∎

By Lemma 6.2, we have the following properties.

  1. 1.

    The action of SL(2,)SL(2,\mathbb{R}) on S12(r)S^{2}_{1}(r) induces the action of SL(2,)SL(2,\mathbb{R}) on ColS12(r)(D)\operatorname{Col}_{S^{2}_{1}(r)}(D). Then, S12(r)S^{2}_{1}(r)-colorings C1,C2C_{1},C_{2} are in the same SL(2,)SL(2,\mathbb{R})-orbit if and only if ΦK,rC1\Phi_{K,r}C_{1} and ΦK,rC2\Phi_{K,r}C_{2} are conjugate.

  2. 2.

    A S12(r)S^{2}_{1}(r)-coloring CC is trivial in the sense of Definition 2.10 if and only if ΦK,rC\Phi_{K,r}C is abelian.

Finally, we end this section by preparing a proposition used in section 8.

Proposition 7.3.

Let KK be (2,n)(2,n)-torus knots with a diagram DD as shown in Fig. 2, 𝔪π1(S3K)\mathfrak{m}\in\pi_{1}(S^{3}\setminus K) be a meridian, and D(r)D(r) be an S12(r)S^{2}_{1}(r)-element defined in Proposition 2.14. For ρInnSL(2,)\rho\in\operatorname{Inn}SL(2,\mathbb{R}),

{fHom(π1(S3K),SL(2,)):f(𝔪)=ρ(D(r))}\displaystyle\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),SL(2,\mathbb{R}))\ :\ {}f(\mathfrak{m})=\rho(D(r))\}
\displaystyle\simeq {CColS12(r)(D):C(α0)=ρ(D(r))}\displaystyle\{C\in\operatorname{Col}_{S^{2}_{1}(r)}(D)\ :\ {}C(\alpha_{0})=\rho(D(r))\}
=\displaystyle= {ρC0}{ρCj,b,c:j=1,3,,n2,b,cbc=4sin2θj(sin2θj+sinh2r)sinh2r}\displaystyle\{\rho\circ C_{0}\}\cup\left\{\rho\circ C_{j,b,c}\ :\ {}\begin{matrix}j=1,3,\cdots,n-2,\ b,c\in\mathbb{R}\\ bc=-\frac{4\sin^{2}{\theta_{j}}(\sin^{2}{\theta_{j}}+\sinh^{2}{r})}{\sinh^{2}{r}}\end{matrix}\right\}
Proof.

The result follows in light of the argument in section 4 and Corollorary 6.3. ∎

8 The longitudinal mapping knot invariant for SL(2,)SL(2,\mathbb{R})

In this section, we determine the value of the longitudinal mapping for SL(2,)SL(2,\mathbb{R}) in the case xS12(r)x\in S^{2}_{1}(r) and KK are (2,n)(2,n)-torus knots.

We introduce the definition of the longitudinal mapping knot invariant defined by Clark-Saito [4]. According to Clark-Saito [4], the longitudinal mapping is an extention of the quandle cocycle invariant defined by Carter et al. [3] and the knot colouring polynomial defined by Eisermann [5].

Definition 8.1 (Clark-Saito [4]).

Let GG be a group, xx be an elemnt of GG, and KK be a knot. The longitudinal mapping is a map

Gx(K):{fHom(π1(S3K),G):f(𝔪)=x}Gff(𝔩),\mathcal{L}_{G}^{x}(K):\{f\in\operatorname{Hom}(\pi_{1}(S^{3}\setminus K),G)\ :\ {}f(\mathfrak{m})=x\}\to G\quad f\mapsto f(\mathfrak{l}),

where 𝔪π1(S3K)\mathfrak{m}\in\pi_{1}(S^{3}\setminus K) is a meridian and 𝔩π1(S3K)\mathfrak{l}\in\pi_{1}(S^{3}\setminus K) be a longitude. When there is no choice of confusion, we write Gx\mathcal{L}_{G}^{x} inplace of Gx(K)\mathcal{L}_{G}^{x}(K).

Remark 8.2.

The definition of longitudinal mapping does not depend on the meridian-longitude pair. See [4, Remark 3.3 and Theorem B.4] for more detail.

By Proposition 7.3, we identify the domain of the longitudinal mapping as a set of some quandle colorings.

Suppose KK be (2,n)(2,n)-torus knot with a diagram DD as showed in Fig. 2.

Theorem 8.3.

For j=1,3,,n2j=1,3,\cdots,n-2 and an inner automorphism ρInnG\rho\in\operatorname{Inn}G,

Gx(ΨK,r(ρC0))\displaystyle\mathcal{L}^{x}_{G}(\Psi_{K,r}(\rho\circ C_{0})) =\displaystyle= (1001),\displaystyle\begin{pmatrix}1&0\\ 0&1\end{pmatrix},
Gx(ΨK,r(ρCj,b,c))\displaystyle\mathcal{L}^{x}_{G}(\Psi_{K,r}(\rho\circ C_{j,b,c})) =\displaystyle= ρ((e2nr00e2nr)).\displaystyle\rho(\begin{pmatrix}-e^{-2nr}&0\\ 0&-e^{2nr}\end{pmatrix}).
Proof.

By Lemma A.2, for any representation f:π1(S3K)SL(2,)f:\pi_{1}(S^{3}\setminus K)\to SL(2,\mathbb{R}),

f(𝔩)=f(α0)2n(f(α0)f(α1))n.f(\mathfrak{l})=f(\alpha_{0})^{-2n}(f(\alpha_{0})f(\alpha_{1}))^{n}.

By Proposition 7.3, we should consider the following two cases.

(In the case f=ΨK,r(C0)f=\Psi_{K,r}(C_{0})) Since f(α0)=f(α1)f(\alpha_{0})=f(\alpha_{1}),

Gx(f)=f(α0)2n(f(α0)f(α0))n=(1001).\mathcal{L}_{G}^{x}(f)=f(\alpha_{0})^{-2n}(f(\alpha_{0})f(\alpha_{0}))^{n}=\begin{pmatrix}1&0\\ 0&1\end{pmatrix}.

(In the case f=ΨK,r(C0)f=\Psi_{K,r}(C_{0})) By Lemma 5.2,

Gx(f)=ρ((er00er)2n(1001))=ρ((e2nr00e2nr)).\mathcal{L}_{G}^{x}(f)=\rho(\begin{pmatrix}e^{r}&0\\ 0&e^{-r}\end{pmatrix}^{-2n}\begin{pmatrix}-1&0\\ 0&-1\end{pmatrix})=\rho(\begin{pmatrix}-e^{-2nr}&0\\ 0&-e^{2nr}\end{pmatrix}).

Appendix A A presentation of a longitude of (2,n)(2,n)-torus knots

We see a presentation of a longitude of (2,n)(2,n)-torus knots. The content of this section has already been done in [4, Lemma 6.3] essentially, but there is a fatal typographical error in the proof, so we prove it again for completeness. See Remark A.3 for more details on [4, Lemma 6.3].

Let KK be (2,n)(2,n)-torus knots with a diagram DD as showed in Fig. 2. The knot group π1(S3K)\pi_{1}(S^{3}\setminus K) has a Wirtinger presentation with respect to DD:

α0,,αn1:αi+21αi+11αiαi+1(j=0,1,,n2).\left\langle\alpha_{0},\cdots,\alpha_{n-1}\ :\ {}\alpha_{i+2}^{-1}\alpha_{i+1}^{-1}\alpha_{i}\alpha_{i+1}\quad(j=0,1,\cdots,{n-2})\right\rangle.
Lemma A.1.

For j=0,,n12j=0,\cdots,\tfrac{n-1}{2},

α2j\displaystyle\alpha_{2j} =\displaystyle= (α0α1)jα0(α0α1)j,\displaystyle(\alpha_{0}\alpha_{1})^{-j}\alpha_{0}(\alpha_{0}\alpha_{1})^{j},
α2j+1\displaystyle\alpha_{2j+1} =\displaystyle= (α0α1)jα1(α0α1)j.\displaystyle(\alpha_{0}\alpha_{1})^{-j}\alpha_{1}(\alpha_{0}\alpha_{1})^{j}.
Proof.

The result follows by induction on jj. ∎

Lemma A.2 (c.f. Clark-Saito [4]).

A longitude 𝔩π1(S3K)\mathfrak{l}\in\pi_{1}(S^{3}\setminus K) has the following presentation:

𝔩=α02n(α0α1)n\mathfrak{l}=\alpha_{0}^{-2n}(\alpha_{0}\alpha_{1})^{n}
Proof.

Suppose n=2k+1n=2k+1. By [2, 3.13 Remark.],

𝔩=(α1α3α2k1)(α0α2α2k)α0n.\mathfrak{l}=(\alpha_{1}\alpha_{3}\cdots\alpha_{2k-1})(\alpha_{0}\alpha_{2}\cdots\alpha_{2k})\alpha_{0}^{-n}.

By Lemma A.1,

α1α3α2k1\displaystyle\alpha_{1}\alpha_{3}\cdots\alpha_{2k-1} =\displaystyle= α0k(α0α1)k,\displaystyle\alpha_{0}^{-k}(\alpha_{0}\alpha_{1})^{k},
α0α2α2k\displaystyle\alpha_{0}\alpha_{2}\cdots\alpha_{2k} =\displaystyle= α0α1k(α0α1)k.\displaystyle\alpha_{0}\alpha_{1}^{-k}(\alpha_{0}\alpha_{1})^{k}.

Therefore, by Lemma A.1 and α0=αn=α2k+1\alpha_{0}=\alpha_{n}=\alpha_{2k+1},

𝔩=α0k(α0α1)k+1α1k1(α0α1)kα0n=α0k(α0α1)nα0nk1.\mathfrak{l}=\alpha_{0}^{-k}(\alpha_{0}\alpha_{1})^{k+1}\alpha_{1}^{-k-1}(\alpha_{0}\alpha_{1})^{k}\alpha_{0}^{-n}=\alpha_{0}^{-k}(\alpha_{0}\alpha_{1})^{n}\alpha_{0}^{-n-k-1}.

The result follows since meridians and longitudes commute (see [2, 3.14 Definition and Proposition]). ∎

Remark A.3.

The statement of [4, Lemma 6.3] is true. However, the proof of [4, Lemma 6.3] contains a fatal typographical error: Clark and Saito presented a image of a longitude of (2,n)(2,n)-torus knots as

(C)=q0n(q1q3q2k3)(q0q2q2k)\mathcal{L}(C)=q_{0}^{-n}(q_{1}q_{3}\cdots q_{2k-3})(q_{0}q_{2}\cdots q_{2k})

to prove [4, Lemma 6.3], but the correct presentation is

(C)=q0n(q1q3q2k1)(q0q2q2k).\mathcal{L}(C)=q_{0}^{-n}(q_{1}q_{3}\cdots q_{2k-1})(q_{0}q_{2}\cdots q_{2k}).

Appendix B The solutions of a equation λmμm=λm+1μm+1\lambda^{m}-\mu^{m}=\lambda^{m+1}-\mu^{m+1}

Suppose mm is a positive integer. We see properties of λ,μ\lambda,\mu\in\mathbb{C} satisfying following conditions: λμ=1\lambda\mu=1, λμ\lambda\neq\mu, and λmμm=λm+1μm+1\lambda^{m}-\mu^{m}=\lambda^{m+1}-\mu^{m+1}.

Lemma B.1.
λ{expπj12m+1:j=1,3,,2m1,2m+3,2m+5,,4m+1}.\lambda\in\left\{\operatorname{exp}\frac{\pi j\sqrt{-1}}{2m+1}\ :\ {}j=1,3,\cdots,2m-1,2m+3,2m+5,\cdots,4m+1\right\}.
Proof.

Since λ\lambda and μ\mu satisfy λμ=1\lambda\mu=1,

0\displaystyle 0 =\displaystyle= λmμmλm+1+μm+1\displaystyle\lambda^{m}-\mu^{m}-\lambda^{m+1}+\mu^{m+1}
=\displaystyle= i=0mλ2imi=1mλ2im1\displaystyle\sum_{i=0}^{m}\lambda^{2i-m}-\sum_{i=1}^{m}\lambda^{2i-m-1}
=\displaystyle= λm(λ2m+1+1)λ+1.\displaystyle\frac{\lambda^{-m}(\lambda^{2m+1}+1)}{\lambda+1}.

Considering (λ,μ)(±1,±1)(\lambda,\mu)\neq(\pm 1,\pm 1), the result follows. ∎

We get the following lemma in light of Lemma B.1.

Lemma B.2.
λ+μ{2cosπj2m+1:j=1,3,,2m1}.\lambda+\mu\in\left\{2\cos\frac{\pi j}{2m+1}\ :\ {}j=1,3,\cdots,2m-1\right\}.

Appendix C The equation 5.1

Suppose kk is a positive integer, r>0r>0 and

x=D(r)=(er00er),y=(abcd).x=D(r)=\begin{pmatrix}e^{r}&0\\ 0&e^{-r}\end{pmatrix},\ {}y=\begin{pmatrix}a&b\\ c&d\end{pmatrix}.

We determine the real numbers aa, bb, cc, and dd satisfying following conditions: a+d=2coshra+d=2\cosh{r}, adbc1ad-bc-1, and the equation 5.1, that is,

(xy)kx=y(xy)k.(xy)^{k}x=y(xy)^{k}.

Let λ\lambda, μ\mu\in\mathbb{C} be the two solutions of the characteristic equation of yy

t2(aer+der)t+1=0.t^{2}-(ae^{r}+de^{-r})t+1=0.
Lemma C.1.

For a positive integer mm,

(xy)m={(amerm0ceri=0m1(aer)i(der)mi1dmerm)ifb=0,(amermberi=0m1(aer)i(der)mi10dmerm)ifc=0,M1(m)ifb0 and c0 and λμ,M2(m)ifb0 and c0 and λ=μ,(xy)^{m}=\left\{\begin{array}[]{ccl}\begin{pmatrix}a^{m}e^{rm}&0\\ ce^{-r}\sum_{i=0}^{m-1}(ae^{r})^{i}(de^{-r})^{m-i-1}&d^{m}e^{-rm}\end{pmatrix}&\mbox{if}&b=0,\\ \begin{pmatrix}a^{m}e^{rm}&be^{r}\sum_{i=0}^{m-1}(ae^{r})^{i}(de^{-r})^{m-i-1}\\ 0&d^{m}e^{-rm}\end{pmatrix}&\mbox{if}&c=0,\\ M_{1}(m)&\mbox{if}&b\neq 0\mbox{ and }c\neq 0\mbox{ and }\lambda\neq\mu,\\ M_{2}(m)&\mbox{if}&b\neq 0\mbox{ and }c\neq 0\mbox{ and }\lambda=\mu,\end{array}\right.

where M1(m)M_{1}(m) is a matrix

1λμ(λm1+μm1+aer(λmμm)ber(λmμm)cer(λmμm)λm+1μm+1aer(λmμm))\frac{1}{\lambda-\mu}\begin{pmatrix}-\lambda^{m-1}+\mu^{m-1}+ae^{r}(\lambda^{m}-\mu^{m})&be^{r}(\lambda^{m}-\mu^{m})\\ ce^{-r}(\lambda^{m}-\mu^{m})&\lambda^{m+1}-\mu^{m+1}-ae^{r}(\lambda^{m}-\mu^{m})\end{pmatrix}

and M2(m)M_{2}(m) is a matrix

(((m+1)λ+amer)λm1bermλm1cermλm1λm1((m1)λamer)).\begin{pmatrix}\left(-(m+1)\lambda+ame^{r}\right)\lambda^{m-1}&be^{r}m\lambda^{m-1}\\ ce^{-r}m\lambda^{m-1}&-\lambda^{m-1}\left((m-1)\lambda-ame^{r}\right)\end{pmatrix}.
Proof.

(In the case b=0b=0 or c=0c=0) The result follows by the induction on mm.

(In the case b0b\neq 0 and c0c\neq 0 and λμ\lambda\neq\mu) The matrix xyxy is diagonalizable as follows:

xy=(berberλaerμaer)(λ00μ)(berberλaerμaer)1.xy=\begin{pmatrix}be^{r}&be^{r}\\ \lambda-ae^{r}&\mu-ae^{r}\end{pmatrix}\begin{pmatrix}\lambda&0\\ 0&\mu\end{pmatrix}\begin{pmatrix}be^{r}&be^{r}\\ \lambda-ae^{r}&\mu-ae^{r}\end{pmatrix}^{-1}.

Therefore, the result follows by direct computation.

(In the case b0b\neq 0 and c0c\neq 0 and λμ\lambda\neq\mu) The matrix xyxy has the Jordan normal form

xy=(ber0λaer1)(λ10λ)(ber0λaer1)1.xy=\begin{pmatrix}be^{r}&0\\ \lambda-ae^{r}&1\end{pmatrix}\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}\begin{pmatrix}be^{r}&0\\ \lambda-ae^{r}&1\end{pmatrix}^{-1}.

Thus the result follows by direct computation. ∎

The following lemmas are derived from Lemma C.1.

Lemma C.2.

b=0b=0 or c=0c=0 if and only if (a,b,c,d)=(er,0,0,er)(a,b,c,d)=(e^{r},0,0,e^{-r}).

Lemma C.3.

b0b\neq 0 and c0c\neq 0 and λμ\lambda\neq\mu if and only if λkμk=λk+1μk+1\lambda^{k}-\mu^{k}=\lambda^{k+1}-\mu^{k+1}.

Lemma C.4.

If b0b\neq 0 and c0c\neq 0 and λ=μ\lambda=\mu, there is no aa, bb, cc or dd satisfying the conditions.

Acknowledgement

The author is grateful to Professor Hiroyuki Ochiai, Kyushu University, for many valuable comments and discussions. He also thanks Michiko Yonemura, University of Miyazaki, for drawing Fig. 2 for him.

This work was supported by JST SPRING, Grant Number JPMJSP2136.

References

  • [1] H. Azcan and R. Fenn, Spherical representations of the link quandles, Turkish J. Math. 18 (1994), no. 1, 102–110.
  • [2] G. Burde, H. Zieschang, Knots, De Gruyter Studies in Mathematics, 5. Walter de Gruyter & Co., Berlin, 1985.
  • [3] J. S. Carter, D. Jelsovsky, S. Kamada, L. Langford, M. Saito, State-sum invariants of knotted curves and surfaces from quandle cohomology, Electron. Res. Announc. Amer. Math. Soc. 5 (1999), 146–156.
  • [4] W.E. Clark and M. Saito, Longitudinal mapping knot invariant for SU(2), J. Knot Theory Ramifications 27 (2018), no. 11.
  • [5] M. Eisermann, Knot colouring polynomials, Pacific J. Math. 231 (2007), no. 2, 305–336.
  • [6] K. Ishikawa, On the classification of smooth quandles, preprint.
  • [7] D. Joyce, A classifying invariant of knots, the knot quandle, J. Pure Appl. Algebra 23 (1982), no. 1, 37–65.
  • [8] S. Kamada, Surface-knots in 4-space. An introduction, Springer Monographs in Mathematics. Springer, Singapore, 2017.
  • [9] S.V. Matveev, Distributive groupoids in knot theory, (Russian) Mat. Sb. (N.S.) 119(161) (1982), no. 1, 78–88, 160.
  • [10] T. Nosaka, Homotopical interpretation of link invariants from finite quandles, Topology Appl. 193 (2015), 1–30.
  • [11] T. Nosaka, Quandles and topological pairs. Symmetry, knots, and cohomology, SpringerBriefs in Mathematics. Springer, Singapore, 2017.
  • [12] T. Nosaka, de Rham theory and cocycles of cubical sets from smooth quandles, Kodai Math. J. 42 (2019), no. 1, 111–129.
  • [13] M. Takasaki, Abstraction of symmetric transformations, (Japanese) Tôhoku Math. J. 49 (1943), 145–207.
  • [14] K. Yonemura, Note on spherical quandles, arXiv:2104.04921, preprint.

Department of Mathematics, Kyushu University, 744 Motooka, Nishi-ku, Fukuoka 819–0395, Japan