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

Constructing an infinite family of quandles from a quandle

Pedro Lopes  and  Manpreet Singh Center for Mathematical Analysis, Geometry, and Dynamical Systems,
Department of Mathematics,
Instituto Superior Técnico, Universidade de Lisboa
1049-001 Lisbon, Portugal
[email protected] [email protected]
Abstract.

Quandles are self-distributive, right-invertible, idempotent algebras. A group with conjugation for binary operation is an example of a quandle. Given a quandle (Q,)(Q,\ast) and a positive integer nn, define anb=((ab))bna\ast_{n}b=(\cdots(a\ast\underbrace{b)\ast\cdots)\ast b}_{n}, where a,bQa,b\in Q. Then, (Q,n)(Q,\ast_{n}) is again a quandle. We set forth the following problem. “Find (Q,)(Q,\ast) such that the sequence {(Q,n):n+}\{(Q,\ast_{n})\,:\,n\in\mathbb{Z}^{+}\} is made up of pairwise non-isomorphic quandles.” In this article we find such a quandle (Q,)(Q,\ast). We study the general linear group of 22-by-22 matrices over \mathbb{C} as a quandle under conjugation. Its (algebraically) connected components, that is, its conjugacy classes, are subquandles of it. We show the latter are connected as quandles and prove rigidity results about them such as the dihedral quandle of order 33 is not a subquandle for most of them. Then we consider the quandle which is the projective linear group of 22-by-22 matrices over \mathbb{C} with conjugation, and prove it solves the problem above. In the course of this work we prove a sufficient and necessary condition for a quandle to be latin. This will reduce significantly the complexity of algorithms for ascertaining if a quandle is latin.

Key words and phrases:
Quandle, connected quandle, 22-connected quandle, nn-quandle, general linear group, projective linear group
2020 Mathematics Subject Classification:
20N02, 57K12

Introduction

Quandles are binary algebras whose axioms are the interpretation of Reidemeister moves of oriented link diagrams. Any group with conjugation for binary operation is a quandle. In the 1980s, Joyce [23, 24] and Matveev [31] proved that the fundamental quandle of an oriented knot is a complete knot invariant up to the orientation of the ambient space and that of the knot. This result extends to orientable non-split links. Since then quandles have been studied extensively and are used to construct new invariants. We refer the readers to [20, 33, 25] for more details.

Besides knot theory, the importance of quandles paved into other areas of mathematics, such as Hopf algebras [1], solutions to the set-theoretic quantum Yang-Baxter equation [1, 22, 17] and Riemannian symmetric spaces [29]. Concurrently, a study of quandles and racks from the algebraic point of view emerged. A (co)homology theory for quandles and racks has been developed in [13, 12] that led to stronger invariants for knots and knotted surfaces. Algebraic and combinatorial properties of quandles have been investigated, for example, in [4, 2, 7, 8, 34, 9, 15, 32, 32, 27, 26, 30]. In the last few years new aspects of quandle theory were introduced like covering theory [16, 18], extension theory [11, 3], commutator theory [10] and ring theory [19, 6].

Given a quandle (Q,)(Q,\ast) and a positive integer nn, we define anb=((ab))bna\ast_{n}b=(\cdots(a\ast\underbrace{b)\ast\cdots)\ast b}_{n} for a,bQa,b\in Q. Then, (Q,n)(Q,\ast_{n}) is again a quandle (see [5]). In this article we present a solution to the following problem.

Problem.

Find a quandle (Q,)(Q,\ast) such that the sequence {(Q,n):n+}\{(Q,\ast_{n})\,:\,n\in\mathbb{Z}^{+}\} is made up of pairwise non-isomomorphic quandles.

In the current article we also address the conjugation quandle on GL(2,)\operatorname{GL}(2,\mathbb{C}) and on PGL(2,)\operatorname{PGL}(2,\mathbb{C}), the latter eventually providing an answer to the problem above (Theorem 5.16). We let Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) stand for the conjugation quandle on GL(2,)\operatorname{GL}(2,\mathbb{C}). We prove a number of results about Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) and about its infinite (algebraically) connected components which are subquandles of it. (We note that since these are topological groups then they are also topological quandles.) Our results are new even from the point of view of conjugacy classes of GL(2,)\operatorname{GL}(2,\mathbb{C}). For instance, we prove that the non-trivial connected components of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) are 22-connected (Theorem 4.4, Theorem 4.11, Theorem 4.15). This means that some pair of elements in the quandle, say (a,b)(a,b) requires two elements, say x,yx,y, to solve the equation b=(((az1)z2))znb=(\cdots((a\ast z_{1})\ast z_{2})\cdots)\ast z_{n} and no (a,b)(a,b) in the quandle requires more than two x,yx,y to solve said equation. In the course of our work on Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})), we prove a general result concerning nn-connectedness from which we obtain that a quandle is latin if and only if one of its left multiplications is bijective (Theorem 2.1). This latter result reduces considerably the complexity of the algorithms for checking if a quandle is latin.

Resuming the discussion of our work on Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})), we also provide a partial classification of the quandles which are the connected components of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) (Theorem 6.1, Theorem 6.2, Theorem 6.4). In particular, we prove that there are infinitely many non-isomporphic such quandles. Furthermore, we prove results about rigidity of these quandles. For instance, we prove that R3\operatorname{R}_{3} cannot be a subquandle of most of them and also that certain strings of two elements do not fix elements by right multiplication (Theorem 6.10, Theorem 6.11).

The present article is organized as follows. In Section 1, we recall the basics of quandles that are used later. In Section 2, we find a sufficient and necessary condition for a quandle to be Latin (respect., nn-connected). In Section 3, we note down some results of generating quandles by multiplying a quandle operation with itself. In Section 4, we prove that all the non-trivial connected components of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) are 22-connected. In Section 5, we prove that, for n1n\geq 1, the nn-conjugation quandles over PGL(2,)\operatorname{PGL}(2,\mathbb{C}) are pairwise distinct. In Section 6, we provide a partial classification of the connected components of Conj(GL(2,)\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C}). In Section 7, we provide some ideas for future work.

1. Preliminaries

In this section, we recall the definition of quandles and the associated terms with some examples.

Definition 1.1.

A quandle is a non-empty set QQ together with a binary operation * satisfying the following axioms:

  1. Q1

    xx=xx*x=x  for all xQx\in Q.

  2. Q2

    For each x,yQx,y\in Q, there exists a unique zQz\in Q such that x=zyx=z*y.

  3. Q3

    (xy)z=(xz)(yz)(x*y)*z=(x*z)*(y*z)  for all x,y,zQx,y,z\in Q.

The axiom Q2 is equivalent to the bijectivity of the right multiplication by each element of QQ. This gives a dual binary operation 1*^{-1} on QQ defined as x1y=zx*^{-1}y=z if x=zyx=z*y. Thus, the axiom Q2 is equivalent to saying that

(xy)1y=xand(x1y)y=x(x*y)*^{-1}y=x\qquad\textrm{and}\qquad\left(x*^{-1}y\right)*y=x

for all x,yQx,y\in Q, and hence it allows us cancellation from the right. The axioms Q1 and Q3 are referred to as idempotency and distributivity axioms, respectively. The idempotency and cancellation law gives x1x=xx*^{-1}x=x for all xQx\in Q.

Examples.

The following are some examples of quandles.

  • Let SS be any non-empty set. Then define a binary operation * on SS as xy=xx*y=x for all x,ySx,y\in S. Then (S,)(S,*) becomes a quandle and is called trivial quandle.

  • If GG is a group and nn\in\mathbb{Z}, then the binary operation xy=ynxynx*y=y^{-n}xy^{n} turns GG into the quandle Conjn(G)\operatorname{Conj}_{n}(G) called the nn-conjugation quandle of GG. For n=1n=1, the quandle is simply denoted by Conj(G)\operatorname{Conj}(G).

  • A group GG with the binary operation xy=yx1yx*y=yx^{-1}y turns GG into the quandle Core(G)\operatorname{Core}(G) called the core quandle of GG. In particular, if GG is a cyclic group of order nn, then it is called the dihedral quandle and is denoted by Rn\operatorname{R}_{n}. Usually, one writes Rn={0,1,,n1}\operatorname{R}_{n}=\{0,1,\ldots,n-1\} with ij=2jimodni*j=2j-i\mod n.

In Section 3 we elaborate about Conjn\operatorname{Conj}_{n} and Core\operatorname{Core} as functors from the category of groups 𝐆𝐫𝐩\boldsymbol{\operatorname{Grp}} to the category of quandles 𝐐𝐧𝐝\boldsymbol{\operatorname{Qnd}}.

A homomorphism of quandles PP and QQ is a map ϕ:PQ\phi:P\to Q with ϕ(xy)=ϕ(x)ϕ(y)\phi(x*y)=\phi(x)*\phi(y) for all x,yPx,y\in P. By the cancellation law in PP and QQ, we get ϕ(x1y)=ϕ(x)1ϕ(y)\phi(x*^{-1}y)=\phi(x)*^{-1}\phi(y) for all x,yPx,y\in P. The quandle axioms are equivalent to saying that for each yQy\in Q, the map Ry:QQR_{y}:Q\to Q given by Ry(x)=xyR_{y}(x)=x*y is an automorphism of QQ fixing xx. The group generated by such automorphisms is called the group of inner automorphisms of QQ, denoted Inn(Q)\operatorname{Inn}(Q).

We recall some relevant definitions. A quandle QQ is said to be

  • connected if the group Inn(Q)\operatorname{Inn}(Q) acts transitively on QQ. For example, the dihedral quandle R2n+1\operatorname{R}_{2n+1} is connected whereas R2n\operatorname{R}_{2n} is not.

  • involutory if x1y=xyx*^{-1}y=x*y for all x,yQx,y\in Q. For example, the core quandle Core(G)\operatorname{Core}(G) of any group GG is involutory.

  • latin if the left multiplication Lx:QQL_{x}:Q\to Q defined by Lx(y)=xyL_{x}(y)=x*y is a bijection for each xQx\in Q.

  • 11-connected if LxL_{x} is a surjective map for each xQx\in Q. Each latin quandle is obviously 11-connected, but the converse is not true. For example, Core()\operatorname{Core}(\mathbb{C}^{*}) is 11-connected but not latin.

  • nn-connected if it is not kk-connected for 1kn11\leq k\leq n-1 and for any x,yQx,y\in Q, we have x=yz1zlx=y*z_{1}*\cdots*z_{l} and y=xz1zly=x*z_{1}^{\prime}*\cdots*z_{l^{\prime}}^{\prime} where 1l,ln1\leq l,l^{\prime}\leq n and zi,zjQz_{i},z_{j}^{\prime}\in Q for all 1il1\leq i\leq l and 1jl1\leq j\leq l^{\prime}. Note that this property is a quandle invariant.

The associated group As(Q)\operatorname{As}(Q) of a quandle QQ is the group with the set of generators as {ex|xQ}\{e_{x}~{}|~{}x\in Q\} and the defining relations as

exy=ey1exeye_{x*y}=e_{y}^{-1}e_{x}e_{y}

for all x,yQx,y\in Q. For example, if QQ is a trivial quandle, then As(Q)\operatorname{As}(Q) is the free abelian group of rank equal to the cardinality of QQ. The natural map

η:QConj(As(Q))\eta:Q\to\operatorname{Conj}(\operatorname{As}(Q))

given by η(x)=ex\eta(x)=e_{x} is a quandle homomorphism. The map η\eta is not injective in general.

If QQ is a quandle, then by [35, Lemma 4.4.7], we can write

xd(yez)=((xez)dy)ez(called the left association identity)x*^{d}\left(y*^{e}z\right)=\left(\left(x*^{-e}z\right)*^{d}y\right)*^{e}z\quad\textrm{(called the {\it left association identity})}

for all x,y,zQx,y,z\in Q and d,e{1,1}d,e\in\{-1,1\}. Henceforth, we will write a left-associated product

((((a0e1a1)e2a2)e3)en1an1)enan\left(\left(\cdots\left(\left(a_{0}*^{e_{1}}a_{1}\right)*^{e_{2}}a_{2}\right)*^{e_{3}}\cdots\right)*^{e_{n-1}}a_{n-1}\right)*^{e_{n}}a_{n}

simply as

a0e1a1e2enan.a_{0}*^{e_{1}}a_{1}*^{e_{2}}\cdots*^{e_{n}}a_{n}.

A repeated use of left association identity gives the following result [35, Lemma 4.4.8].

Lemma 1.1.

The product

(a0d1a1d2dmam)e0(b0e1b1e2enbn)\left(a_{0}*^{d_{1}}a_{1}*^{d_{2}}\cdots*^{d_{m}}a_{m}\right)*^{e_{0}}\left(b_{0}*^{e_{1}}b_{1}*^{e_{2}}\cdots*^{e_{n}}b_{n}\right)

of two left-associated forms a0d1a1d2dmama_{0}*^{d_{1}}a_{1}*^{d_{2}}\cdots*^{d_{m}}a_{m} and b0e1b1e2enbnb_{0}*^{e_{1}}b_{1}*^{e_{2}}\cdots*^{e_{n}}b_{n} in a quandle can again be written in a left-associated form as

a0d1a1d2dmamenbnen1bn1en2e1b1e0b0e1b1e2enbn.a_{0}*^{d_{1}}a_{1}*^{d_{2}}\cdots*^{d_{m}}a_{m}*^{-e_{n}}b_{n}*^{-e_{n-1}}b_{n-1}*^{-e_{n-2}}\cdots*^{-e_{1}}b_{1}*^{e_{0}}b_{0}*^{e_{1}}b_{1}*^{e_{2}}\cdots*^{e_{n}}b_{n}.

Thus, any product of elements of a quandle QQ can be expressed in the canonical left-associated form a0e1a1e2enana_{0}*^{e_{1}}a_{1}*^{e_{2}}\cdots*^{e_{n}}a_{n}, where a0a1a_{0}\neq a_{1}, and for i=1,2,,n1i=1,2,\ldots,n-1, ei=ei+1e_{i}=e_{i+1} whenever ai=ai+1a_{i}=a_{i+1}.

Let QQ be a quandle, and x,yQx,y\in Q. Then for n+n\in\mathbb{Z}^{+}, we let xϵny=xϵyϵyϵϵyn times x*^{\epsilon n}y=x\underbrace{*^{\epsilon}y*^{\epsilon}y*^{\epsilon}\cdots*^{\epsilon}y}_{n\textrm{ times }}, where ϵ=±1\epsilon=\pm 1. We say QQ is nn-quandle if xny=xx*^{n}y=x for all x,yQx,y\in Q. For example, core quandles are 22-quandles.

Remark 1.1.

Note that if QQ is nn-quandle, then it is also nmnm-quandle, where m+m\in\mathbb{Z}^{+}.

Definition 1.2.

For n+n\in\mathbb{Z}^{+}, a quandle QQ is said to be of type nn, if nn is the least number satisfying pnq=pp*^{n}q=p for all p,qQp,q\in Q. We denote the type of QQ by Type(Q)\operatorname{Type}(Q). If there does not exist any such nn, then we say the quandle is of type \infty.

If for a quandle QQ, Type(Q)\operatorname{Type}(Q) is finite, then we say QQ is of finite type. Clearly finite quandles and nn-quandles are of finite type.

A subset PP of a quandle QQ is called subquandle if it is a quandle under the same binary operation of QQ. The terms defined for quandles in this section are similarly defined for subquandles, for example, if QQ is a quandle and PP its subquandle, then we say PP is an nn-subquandle, if for all x,yPx,y\in P we have xny=xx*^{n}y=x.

2. A sufficient and necessary condition for a quandle to be latin (resp., nn-connected)

In this section, we find a sufficient and necessary condition for a quandle to be latin and nn-connected.

Theorem 2.1.

Let QQ be a quandle. Then QQ is latin if and only if there exists xQx\in Q such that Lx:QQL_{x}:Q\to Q is bijective.

Proof.

By definition of latin quandle the necessity is obvious.

Now we will prove the converse. Let yQy\in Q and yxy\neq x. Since LxL_{x} is surjective, there exists αQ\alpha\in Q such that Lx(α)=yL_{x}(\alpha)=y, which implies x=Rα1(y)x=R_{\alpha}^{-1}(y).

Claim.

LyL_{y} is surjective.

Let aQa\in Q. We need to find bQb\in Q such that Ly(b)=aL_{y}(b)=a. Consider a=Rα1(a)a^{\prime}=R_{\alpha}^{-1}(a). Again the surjectivity of LxL_{x} implies the existence of βQ\beta\in Q such that

Lx(β)\displaystyle L_{x}(\beta) =a\displaystyle=a^{\prime}
\displaystyle\iff Rα1(y)β\displaystyle R_{\alpha}^{-1}(y)*\beta =Rα1(a)\displaystyle=R_{\alpha}^{-1}(a)
\displaystyle\iff Rα(Rα1(y)β)\displaystyle R_{\alpha}(R_{\alpha}^{-1}(y)*\beta) =a\displaystyle=a
\displaystyle\iff yRα(β)\displaystyle y*R_{\alpha}(\beta) =a.\displaystyle=a.

Taking b=Rα(β)b=R_{\alpha}(\beta) implies that Ly(b)=aL_{y}(b)=a, and thus proves the claim.

Claim.

LyL_{y} is injective.

Let b1,b2Qb_{1},b_{2}\in Q. If Ly(b1)=a=Ly(b2)L_{y}(b_{1})=a=L_{y}(b_{2}), then Rα1(yb1)=Rα1(a)=Rα1(yb2)R_{\alpha}^{-1}(y*b_{1})=R_{\alpha}^{-1}(a)=R_{\alpha}^{-1}(y*b_{2}) which is equivalent to Lx(Rα1(b1))=Rα1(a)=Lx(Rα1(b2))L_{x}(R_{\alpha}^{-1}(b_{1}))=R_{\alpha}^{-1}(a)=L_{x}(R_{\alpha}^{-1}(b_{2})). But since LxL_{x} is injective, we have Rα1(b1)=Rα1(b2)R_{\alpha}^{-1}(b_{1})=R_{\alpha}^{-1}(b_{2}), which further implies that b1=b2b_{1}=b_{2}. Thus LyL_{y} is injective. ∎

In the HAP package [21], finite Latin quandles are recognized by testing whether all left translations are permutations or not. As an application of Theorem 2.1, the time complexity of detecting Latin quandles can be reduced.

Using [27, Proposition 3.3] and Theorem 2.1, we have the following result.

Theorem 2.2.

Let QQ be a finite quandle and xQx\in Q. Suppose that the cycles of RxR_{x} have pairwise distinct lengths and for every yQy\in Q, RyR_{y} has a unique fixed point. Then QQ is Latin.

In the following theorem, we find a sufficient and necessary condition for a quandle to be nn-connected.

Theorem 2.3.

Let QQ be a quandle and xQx\in Q such that for any element aQa\in Q there exist xi’sQx_{i}\textrm{'s}\in Q, where 1in1\leq i\leq n such that RxiRx1(x)=aR_{x_{i}}\cdots R_{x_{1}}(x)=a. If QQ is not mm-connected, where 1mn11\leq m\leq n-1, then QQ is nn-connected.

Proof.

Let y,zQy,z\in Q and yxy\neq x. By assumption, there exist x1,,xjx_{1},\ldots,x_{j}, where jnj\leq n, such that y=RxjRxj1Rx1(x)y=R_{x_{j}}R_{x_{j-1}}\cdots R_{x_{1}}(x). Again by assumption, there exist x1,,xkx_{1}^{\prime},\ldots,x_{k}^{\prime}, where knk\leq n, such that

RxkRx1(x)\displaystyle R_{x_{k}^{\prime}}\cdots R_{x_{1}^{\prime}}(x) =Rx11Rxj1(z)\displaystyle=R_{x_{1}}^{-1}\cdots R_{x_{j}}^{-1}(z)
\displaystyle\iff RxkRx1Rx11Rxj1(y)\displaystyle R_{x_{k}^{\prime}}\cdots R_{x_{1}^{\prime}}R_{x_{1}}^{-1}\cdots R_{x_{j}}^{-1}(y) =Rx11Rxj1(z)\displaystyle=R_{x_{1}}^{-1}\cdots R_{x_{j}}^{-1}(z)
\displaystyle\iff RykRy1(y)\displaystyle R_{y_{k}}\cdots R_{y_{1}}(y) =z\displaystyle=z

where yl=RxjRx1(xl)y_{l}=R_{x_{j}}\cdots R_{x_{1}}(x_{l}^{\prime}) for 1lkn1\leq l\leq k\leq n. This completes the proof. ∎

3. Properties of multiplication of quandle operation

In this section we study properties of quandles which are generated by iterating its quandle operation a given number of times (see Proposition 3.1)-multiplication of quandle operation (see [5]).

The following result is proved in [5, Theorem 4.11] in a general setting, but we are noting it down for completeness.

Proposition 3.1.

Let (Q,)(Q,*) be a quandle. Then for each n+n\in\mathbb{Z}^{+}, (Q,n)(Q,*_{n}) is a quandle, where anb:=anba*_{n}b:=a*^{n}b for a,b(Q,n).a,b\in(Q,*_{n}).

Proof.

We just need to verify that (Q,n)(Q,*_{n}) satisfies the quandle axioms. Clearly it satisfies Q1 axiom. Let x,y(Q,n)x,y\in(Q,*_{n}) and z=xnyz=x*^{-n}y in QQ. Then x=znyx=z*_{n}y, and zz is unique because RynR_{y}^{-n} is a bijection. Thus Q2 is also satisfied. Now take x,y,z(Q,n)x,y,z\in(Q,*_{n}). We need to show that (xny)nz=(xnz)n(ynz)(x*_{n}y)*_{n}z=(x*_{n}z)*_{n}(y*_{n}z). Using Q3 recursively nn times, we note that (xny)z=(xz)n(yz)(x*_{n}y)*z=(x*z)*_{n}(y*z). Thus (xny)nz=(xnz)n(ynz)(x*_{n}y)*_{n}z=(x*_{n}z)*_{n}(y*_{n}z). ∎

Now onwards, for a quandle (Q,)(Q,*), we denote the quandle (Q,n)(Q,*_{n}) by 𝒬n(Q)\mathcal{Q}_{n}(Q).

Remark 3.1.

If QQ is a quandle and Type(Q)\operatorname{Type}(Q) is \infty, then Type(𝒬n(Q))\operatorname{Type}\big{(}\mathcal{Q}_{n}(Q)\big{)} is \infty for all n+n\in\mathbb{Z}^{+}.

If QQ is a quandle such that Type(Q)(Q) is finite, say nn, then 𝒬n(Q)\mathcal{Q}_{n}(Q) is a trivial quandle. Furthermore, 𝒬m(Q)𝒬(m+n)(Q)\mathcal{Q}_{m}(Q)\cong\mathcal{Q}_{(m+n)}(Q). Thus the following questions are in order.

Question 3.2.

Let QQ be a quandle such that Type(Q)\operatorname{Type}(Q) is \infty. Does there exist n+n\in\mathbb{Z}^{+}, where n1n\neq 1, such that Q𝒬n(Q)Q\cong\mathcal{Q}_{n}(Q)?

Question 3.3.

Let QQ be a quandle such that Type(Q)\operatorname{Type}(Q) is \infty. Does there exist m,n>1m,n>1, where mnm\neq n, such that 𝒬m(Q)𝒬n(Q)\mathcal{Q}_{m}(Q)\cong\mathcal{Q}_{n}(Q)?

In Section 5, we will answer the above questions for Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})), and Conj(PGL(2,))\operatorname{Conj}(\operatorname{PGL}(2,\mathbb{C})).

We note that if f:(Q,)(Q,)f:(Q,*)\to(Q^{\prime},*^{\prime}) is a quandle homomorphism, then the map 𝒬n(f):𝒬n(Q)𝒬n(Q)\mathcal{Q}_{n}(f):\mathcal{Q}_{n}(Q)\to\mathcal{Q}_{n}(Q^{\prime}) defined as 𝒬n(f)(xny)=f(x)nf(y)\mathcal{Q}_{n}(f)(x*_{n}y)=f(x)*^{\prime}_{n}f(y) is a quandle homomorphism. Thus we have the following result.

Proposition 3.4.

Let 𝐐𝐧𝐝\boldsymbol{\operatorname{Qnd}} denote the category of quandles. Then we have a functor 𝓠𝐧:𝐐𝐧𝐝𝐐𝐧𝐝\bm{\mathcal{Q}_{n}}:{\boldsymbol{\operatorname{Qnd}}}\to{\boldsymbol{\operatorname{Qnd}}} defined as 𝓠𝐧(Q)=𝒬n(Q)\bm{\mathcal{Q}_{n}}(Q)=\mathcal{Q}_{n}(Q) and 𝓠𝐧(f:Q1Q2)\bm{\mathcal{Q}_{n}}(f:Q_{1}\to Q_{2}) as 𝒬n(f):𝒬n(Q1)𝒬n(Q2)\mathcal{Q}_{n}(f):\mathcal{Q}_{n}(Q_{1})\to\mathcal{Q}_{n}(Q_{2}).

Remark 3.2.

Using Proposition 3.4, we note that the functor Conjn:𝐆𝐫𝐩𝐐𝐧𝐝\operatorname{Conj}_{n}:\boldsymbol{\operatorname{Grp}}\to\boldsymbol{\operatorname{Qnd}} is the composite of Conj:𝐆𝐫𝐩𝐐𝐧𝐝\operatorname{Conj}:\boldsymbol{\operatorname{Grp}}\to\boldsymbol{\operatorname{Qnd}} and 𝓠𝒏:𝐐𝐧𝐝𝐐𝐧𝐝\bm{\mathcal{Q}_{n}}:{\boldsymbol{\operatorname{Qnd}}}\to{\boldsymbol{\operatorname{Qnd}}} functors, that is Conjn=𝓠𝒏Conj\operatorname{Conj}_{n}=\bm{\mathcal{Q}_{n}}\circ\operatorname{Conj}.

For each quandle QQ, there is a quandle homomorphism σ:QConj(Inn(Q))\sigma:Q\to\operatorname{Conj}(\operatorname{Inn}(Q)) defined as σ(q)=Rq\sigma(q)=R_{q}.

Corollary 3.5.

Let QQ be a quandle. Then for each nn, the following are true:

  1. (1)

    The map 𝒬n(η):𝒬n(Q)Conjn(As(Q))\mathcal{Q}_{n}(\eta):\mathcal{Q}_{n}(Q)\to\operatorname{Conj}_{n}(\operatorname{As}(Q)) defined as xexx\mapsto e_{x} is a quandle homomorphism

  2. (2)

    the map 𝒬n(σ):𝒬n(Q)Conjn(Inn(Q))\mathcal{Q}_{n}(\sigma):\mathcal{Q}_{n}(Q)\to\operatorname{Conj}_{n}(\operatorname{Inn}(Q)) defined as xRxx\mapsto R_{x} is a quandle homomorphism.

Proof.

The result follows from Proposition 3.4. ∎

Proposition 3.6.

Let QQ be a quandle. Then for each nn, the following holds:

  1. (1)

    the map ψ:As(𝒬n(Q))As(Q)\psi:\operatorname{As}(\mathcal{Q}_{n}(Q))\to\operatorname{As}(Q) defined on the generators as ψ(eq)=eqn\psi(e_{q})=e_{q}^{n} is a group homomorphism.

  2. (2)

    the map ηn:𝒬n(Q)Conj(As(Q))\eta_{n}:\mathcal{Q}_{n}(Q)\to\operatorname{Conj}(\operatorname{As}(Q)) defined as xexnx\mapsto e_{x}^{n} is a quandle homomorphism.

  3. (3)

    the map σn:𝒬n(Q)Conj(Inn(Q))\sigma_{n}:\mathcal{Q}_{n}(Q)\to\operatorname{Conj}(\operatorname{Inn}(Q)) defined as xRxnx\mapsto R_{x}^{n} is a quandle homomorphism.

Proof.

Let p,q𝒬n(Q)p,q\in\mathcal{Q}_{n}(Q). Then the map ψ:As(𝒬n(Q))As(Q)\psi:\operatorname{As}(\mathcal{Q}_{n}(Q))\to\operatorname{As}(Q) defined on the generators as ψ(eq)=eqn\psi(e_{q})=e_{q}^{n} is a group homomorphism if and only if

ψ(epnq)\displaystyle\psi(e_{p*_{n}q}) =ψ(eq1epeq)\displaystyle=\psi(e_{q}^{-1}e_{p}e_{q})
\displaystyle\iff epnqn\displaystyle e_{p*^{n}q}^{n} =eqnepneqn\displaystyle=e_{q}^{-n}e_{p}^{n}e_{q}^{n}
\displaystyle\iff (eqnepeqn)n\displaystyle(e_{q}^{-n}e_{p}e_{q}^{n})^{n} =eqnepneqn,\displaystyle=e_{q}^{-n}e_{p}^{n}e_{q}^{n},

which is true. Thus ψ:As(𝒬n(Q))As(Q)\psi:\operatorname{As}(\mathcal{Q}_{n}(Q))\to\operatorname{As}(Q) is a group homomorphism. The second assertion follows from the fact that the natural map η:𝒬n(Q)Conj(As(𝒬n(Q)))\eta:\mathcal{Q}_{n}(Q)\to\operatorname{Conj}(\operatorname{As}(\mathcal{Q}_{n}(Q))) is a quandle homomorphism. The last assertion follows from the fact that there is a natural surjective group homomorphism As(Q)Inn(Q)\operatorname{As}(Q)\to\operatorname{Inn}(Q) given by exRxe_{x}\mapsto R_{x}. ∎

4. Connected components of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C}))

Consider the group GL(2,)\operatorname{GL}(2,\mathbb{C}) of invertible 22-by-22 matrices over the field of complex numbers \mathbb{C}. The projective general linear group PGL(2,)\operatorname{PGL}(2,\mathbb{C}) of degree 22 over \mathbb{C} is the quotient of GL(2,)\operatorname{GL}(2,\mathbb{C}) by its center Z(GL(2,))\operatorname{Z}(\operatorname{GL}(2,\mathbb{C}))\cong\mathbb{C}^{*}. It is known that each connected component of a quandle is a subquandle. In this section we will prove that each non-trivial connected components of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) are 22-connected.

It is well known that there are exactly three families of conjugacy classes in GL(2,)\operatorname{GL}(2,\mathbb{C}), namely:

  1. (1)

    for each λ1λ2\lambda_{1}\neq\lambda_{2}\in\mathbb{C}^{*}, the conjugacy class of (λ100λ2)\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix},

  2. (2)

    for each λ\lambda\in\mathbb{C}^{*}, the conjugacy class of (λ10λ)\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix},

  3. (3)

    for each λ\lambda\in\mathbb{C}^{*}, the conjugacy class of (λ00λ)\begin{pmatrix}\lambda&0\\ 0&\lambda\end{pmatrix}, which is a singleton set.

Remark 4.1.

We disregard (3)(3) above because each of its quandles corresponds to the trivial quandle with one element.

Notation.
  1. (1)

    For λ,λ1,λ2\lambda,\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}, we denote the set of conjugacy class of (λ100λ2)\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} and (λ10λ)\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix} by Mλ1,λ2M_{\lambda_{1},\lambda_{2}} and MλM_{\lambda}, respectively. Note that Mλ1,λ2=Mλ2,λ1M_{\lambda_{1},\lambda_{2}}=M_{\lambda_{2},\lambda_{1}}

  2. (2)

    We denote the matrix (λ100λ2)\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} by D(λ1,λ2)D(\lambda_{1},\lambda_{2}).

  3. (3)

    For any XGL(2,)X\in\operatorname{GL}(2,\mathbb{C}), the trace of XX and determinant of XX are denoted by trX\operatorname{tr}X and detX\det X, respectively.

Lemma 4.1.

Each connected component of the quandle Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) having more than one element is a non-trivial (infinite) subquandle.

Proof.

Let λ1,λ2,α\lambda_{1},\lambda_{2},\alpha\in\mathbb{C}^{*} and λ1λ2\lambda_{1}\neq\lambda_{2}. Then

(λ100λ2)1(λ1α0λ2)(λ100λ2)\displaystyle\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}^{-1}\begin{pmatrix}\lambda_{1}&\alpha\\ 0&\lambda_{2}\end{pmatrix}\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} =1λ1λ2(λ200λ1)(λ1α0λ2)(λ100λ2)\displaystyle=\frac{1}{\lambda_{1}\lambda_{2}}\begin{pmatrix}\lambda_{2}&0\\ 0&\lambda_{1}\end{pmatrix}\begin{pmatrix}\lambda_{1}&\alpha\\ 0&\lambda_{2}\end{pmatrix}\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}
=(λ1αλ2/λ10λ2)\displaystyle=\begin{pmatrix}\lambda_{1}&\alpha\lambda_{2}/\lambda_{1}\\ 0&\lambda_{2}\end{pmatrix}
(λ1α0λ2).\displaystyle\neq\begin{pmatrix}\lambda_{1}&\alpha\\ 0&\lambda_{2}\end{pmatrix}.

Thus Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is a non-trivial subquandle; it is also an infinite quandle since α\alpha\in\mathbb{C}^{*}.

Let λ,α\lambda,\alpha\in\mathbb{C}^{*}. Then

(λα0λ)(λ0αλ)\displaystyle\begin{pmatrix}\lambda&\alpha\\ 0&\lambda\end{pmatrix}*\begin{pmatrix}\lambda&0\\ \alpha&\lambda\end{pmatrix} =1λ2(λ0αλ)(λα0λ)(λ0αλ)\displaystyle=\frac{1}{\lambda^{2}}\begin{pmatrix}\lambda&0\\ -\alpha&\lambda\end{pmatrix}\begin{pmatrix}\lambda&\alpha\\ 0&\lambda\end{pmatrix}\begin{pmatrix}\lambda&0\\ \alpha&\lambda\end{pmatrix}
=1λ2(λ3+α2λαλ2α3α2λ+λ3)\displaystyle=\frac{1}{\lambda^{2}}\begin{pmatrix}\lambda^{3}+\alpha^{2}\lambda&\alpha\lambda^{2}\\ -\alpha^{3}&-\alpha^{2}\lambda+\lambda^{3}\end{pmatrix}
(λ10λ)\displaystyle\neq\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}

Thus MλM_{\lambda} is a non-trivial subquandle; it is also an infinite quandle since α\alpha\in\mathbb{C}^{*}. ∎

4.1. The proof that Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is 22-connected, where λ1±λ2\lambda_{1}\neq\pm\lambda_{2}

In this subsection we will prove that for λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} where λ1±λ2\lambda_{1}\neq\pm\lambda_{2}, the quandle Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is 22-connected (see Theorem 4.4).

Lemma 4.2.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} and λ1λ2\lambda_{1}\neq\lambda_{2}. Then there exists XMλ1,λ2X\in\operatorname{M}_{\lambda_{1},\lambda_{2}} such that D(λ2,λ1)X=D(λ1,λ2)D(\lambda_{2},\lambda_{1})*X=D(\lambda_{1},\lambda_{2}) if and only if λ1=λ2\lambda_{1}=-\lambda_{2}.

Proof.

For X=(x1,1x1,2x2,1x2,2)X=\begin{pmatrix}x_{1,1}&x_{1,2}\\ x_{2,1}&x_{2,2}\end{pmatrix} in GL(2,)\operatorname{GL}(2,\mathbb{C}), we have:

(x1,1x1,2x2,1x2,2)(λ100λ2)=(λ200λ1)(x1,1x1,2x2,1x2,2)\displaystyle\begin{pmatrix}x_{1,1}&x_{1,2}\\ x_{2,1}&x_{2,2}\end{pmatrix}\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}=\begin{pmatrix}\lambda_{2}&0\\ 0&\lambda_{1}\end{pmatrix}\begin{pmatrix}x_{1,1}&x_{1,2}\\ x_{2,1}&x_{2,2}\end{pmatrix} (1)
{x1,1(λ1λ2)=0x2,2(λ1λ2)=0\displaystyle\Longleftrightarrow~{}\begin{cases}x_{1,1}(\lambda_{1}-\lambda_{2})=0\\ x_{2,2}(\lambda_{1}-\lambda_{2})=0\end{cases}
{x1,1=0x2,2=0(sinceλ1λ2).\displaystyle\Longleftrightarrow~{}\begin{cases}x_{1,1}=0\\ x_{2,2}=0\end{cases}(~{}\textrm{since}~{}\lambda_{1}\neq\lambda_{2}). (2)

Thus for XX to be in Mλ1,λ2\operatorname{M}_{\lambda_{1},\lambda_{2}}, the trace of XX must be zero. This proves the necessity of λ1=λ2.\lambda_{1}=-\lambda_{2}.

Now let λ1=λ2=λ\lambda_{1}=-\lambda_{2}=\lambda, and take X=(01λ20.)X=\begin{pmatrix}0&1\\ \lambda^{2}&0.\end{pmatrix} Then clearly XX is in Mλ1,λ2\operatorname{M}_{\lambda_{1},\lambda_{2}} and satisfies conditions in (2) which further implies that it satisfies (1)\eqref{eq1}. Thus D(λ2,λ1)X=D(λ1,λ2)D(\lambda_{2},\lambda_{1})*X=D(\lambda_{1},\lambda_{2}). ∎

Lemma 4.3.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} such that λ1±λ2\lambda_{1}\neq\pm\lambda_{2}, and A=(abcd)Mλ1,λ2A=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{\lambda_{1},\lambda_{2}} such that either b0b\neq 0 or c0c\neq 0. Then the following holds:

  1. (i)

    there exists X,YMλ1,λ2X,Y\in M_{\lambda_{1},\lambda_{2}} such that D(λ1,λ2)=AXD(\lambda_{1},\lambda_{2})=A*X and D(λ2,λ1)=AY.D(\lambda_{2},\lambda_{1})=A*Y.

  2. (ii)

    there exists X,YM1/λ1,1/λ2X,Y\in M_{1/\lambda_{1},1/\lambda_{2}} such that D(λ1,λ2)X1=AD(\lambda_{1},\lambda_{2})*X^{-1}=A and D(λ2,λ1)Y1=A.D(\lambda_{2},\lambda_{1})*Y^{-1}=A.

Proof.

We will prove part (i). The proof of part (ii) is done along the same lines.

We will show that there exists XMλ1,λ2X\in M_{\lambda_{1},\lambda_{2}} such that D(λ1,λ2)=AX=X1AXD(\lambda_{1},\lambda_{2})=A*X=X^{-1}AX, that is, we will prove that XX is a diagonalizable matrix in Mλ1,λ2M_{\lambda_{1},\lambda_{2}}. We will prove the existence of XX when c0c\neq 0, the other case where b0b\neq 0 is similar. The system of equations for the eigenvectors of AA is (i{1,2}i\in\{1,2\})

(00)=(aλibcdλi)(xy)\begin{pmatrix}0\\ 0\end{pmatrix}=\begin{pmatrix}a-\lambda_{i}&b\\ c&d-\lambda_{i}\end{pmatrix}\begin{pmatrix}x\\ y\end{pmatrix}

so that

(aλib|0cdλi|0)\displaystyle\begin{pmatrix}a-\lambda_{i}&b&\bigm{|}&0\\ c&d-\lambda_{i}&\bigm{|}&0\end{pmatrix} \displaystyle\longrightarrow (aλib|01(dλi)/c|0)\displaystyle\begin{pmatrix}a-\lambda_{i}&b&\bigm{|}&0\\ 1&(d-\lambda_{i})/c&\bigm{|}&0\end{pmatrix}
(0b(aλi)(dλi)/c|01(dλi)/c|0)\displaystyle\longrightarrow\begin{pmatrix}0&b-(a-\lambda_{i})(d-\lambda_{i})/c&\bigm{|}&0\\ 1&(d-\lambda_{i})/c&\bigm{|}&0\end{pmatrix} \displaystyle\longrightarrow (1(dλi)/c|000|0),\displaystyle\begin{pmatrix}1&(d-\lambda_{i})/c&\bigm{|}&0\\ 0&0&\bigm{|}&0\end{pmatrix},

where in the last passage we used the characteristic equation. The eigenvectors are then:

(u(λ1d)/cu) and (v(λ2d)/cv)\begin{pmatrix}u(\lambda_{1}-d)/c\\ u\end{pmatrix}\qquad\text{ and }\qquad\begin{pmatrix}v(\lambda_{2}-d)/c\\ v\end{pmatrix}

where we keep the scalars u(0)u~{}(\neq 0) and v(0)v~{}(\neq 0) for later to choose them appropriately so that the diagonalizing matrix,

X=(u(λ1d)/cv(λ2d)/cuv),X=\begin{pmatrix}u(\lambda_{1}-d)/c&v(\lambda_{2}-d)/c\\ u&v\end{pmatrix},

is in Mλ1,λ2\operatorname{M}_{\lambda_{1},\lambda_{2}}. Thus, we need

{λ1+λ2=trX=v+u(λ1d)/cλ1λ2=detX=uv(λ1λ2)/c\displaystyle\begin{cases}\lambda_{1}+\lambda_{2}=\operatorname{tr}X=v+u(\lambda_{1}-d)/c\\ \lambda_{1}\lambda_{2}=\det X=uv(\lambda_{1}-\lambda_{2})/c\end{cases} (3)
\displaystyle\iff {v=λ1+λ2u(λ1d)/cλ1λ2=u[λ1+λ2u(λ1d)/c](λ1λ2)/c\displaystyle\begin{cases}v=\lambda_{1}+\lambda_{2}-u(\lambda_{1}-d)/c\\ \lambda_{1}\lambda_{2}=u\big{[}\lambda_{1}+\lambda_{2}-u(\lambda_{1}-d)/c\big{]}(\lambda_{1}-\lambda_{2})/c\end{cases}
\displaystyle\Longleftrightarrow {0=u2(λ1d)(λ1λ2)/c2u(λ1+λ2)(λ1λ2)/c+λ1λ2v=λ1+λ2u(λ1d)/c\displaystyle\begin{cases}0=u^{2}(\lambda_{1}-d)(\lambda_{1}-\lambda_{2})/c^{2}-u(\lambda_{1}+\lambda_{2})(\lambda_{1}-\lambda_{2})/c+\lambda_{1}\lambda_{2}\\ v=\lambda_{1}+\lambda_{2}-u(\lambda_{1}-d)/c\end{cases} (4)

Since λ1λ20\lambda_{1}\lambda_{2}\neq 0, the quadratic equation in uu cannot have solutions u=0u=0 and u=c(λ1+λ2)/(λ1d)u=c(\lambda_{1}+\lambda_{2})/(\lambda_{1}-d) (note that λ1λ2\lambda_{1}\neq\lambda_{2}), which further implies that v=0v=0 is not a solution of the above equations. Thus, choosing a solution of the quadratic equation for uu and obtaining vv in terms of this uu provides X=X(u,v)Mλ1,λ2X=X(u,v)\in\operatorname{M}_{\lambda_{1},\lambda_{2}}, such that AX=D(λ1,λ2)A*X=D(\lambda_{1},\lambda_{2}).

Similarly, one can solve uu^{\prime} and vv^{\prime} for a diagonalizing matrix

Y=(u(λ2d)/cv(λ1d)/cuv)Y=\begin{pmatrix}u^{\prime}(\lambda_{2}-d)/c&v^{\prime}(\lambda_{1}-d)/c\\ u^{\prime}&v^{\prime}\end{pmatrix}

to be in Mλ1,λ2\operatorname{M}_{\lambda_{1},\lambda_{2}}. Thus there exists YMλ1,λ2Y\in M_{\lambda_{1},\lambda_{2}} such that AY=D(λ2,λ1)A*Y=D(\lambda_{2},\lambda_{1}). ∎

Theorem 4.4.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}, and λ1λ2\lambda_{1}\neq-\lambda_{2}. Then Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is a connected quandle. If λ1±λ2\lambda_{1}\neq\pm\lambda_{2}, then Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is a 22-connected quandle.

Proof.

If λ1=λ2=λ\lambda_{1}=\lambda_{2}=\lambda, then being a trivial quandle with one element, Mλ,λM_{\lambda,\lambda} is connected.

Now suppose λ1±λ2\lambda_{1}\neq\pm\lambda_{2}. By Q1-axiom D(λ1,λ2)=D(λ1,λ2)D(λ1,λ2)D(\lambda_{1},\lambda_{2})=D(\lambda_{1},\lambda_{2})*D(\lambda_{1},\lambda_{2}). By Lemma 4.2, there does not exist any YMλ1,λ2Y\in M_{\lambda_{1},\lambda_{2}} such that D(λ1,λ2)=D(λ2,λ1)YD(\lambda_{1},\lambda_{2})=D(\lambda_{2},\lambda_{1})*Y. Thus Mλ1,λ2M_{\lambda_{1},\lambda_{2}} cannot be 11-connected. Furthermore, by Lemma 4.3, for any A=(abcd)Mλ1,λ2A=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{\lambda_{1},\lambda_{2}} such that either c0c\neq 0 or b0b\neq 0, there exist X1,Y1,X2,Y2Mλ1,λ2X_{1},Y_{1},X_{2},Y_{2}\in M_{\lambda_{1},\lambda_{2}} such that AX1=D(λ1,λ2)A*X_{1}=D(\lambda_{1},\lambda_{2}), AY1=D(λ2,λ1)A*Y_{1}=D(\lambda_{2},\lambda_{1}), A=D(λ1,λ2)X2A=D(\lambda_{1},\lambda_{2})*X_{2}, and A=D(λ2,λ1)Y2.A=D(\lambda_{2},\lambda_{1})*Y_{2}. Thus for any AMλ1,λ2A\in M_{\lambda_{1},\lambda_{2}}, either there exists XMλ1,λ2X\in M_{\lambda_{1},\lambda_{2}} such that D(λ1,λ2)X=AD(\lambda_{1},\lambda_{2})*X=A or there exist Y,ZMλ1,λ2Y,Z\in M_{\lambda_{1},\lambda_{2}} such that D(λ1,λ2)YZ=AD(\lambda_{1},\lambda_{2})*Y*Z=A. Thus by Theorem 2.3, Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is 22-connected. ∎

4.2. The proof that Mλ,λM_{\lambda,-\lambda} is 22-connected

In this subsection, we will prove that for λ\lambda\in\mathbb{C}^{*}, the quandle Mλ,λM_{\lambda,-\lambda} is 22-connected (see Theorem 4.11).

Lemma 4.5.

Let A=(abcd)M1,1A=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{1,-1} such that a{1,1}a\notin\{1,-1\}. Then there exist X,YM1,1X,Y\in M_{1,-1} such that D(1,1)X=AD(1,-1)*X=A and D(1,1)Y=AD(-1,1)*Y=A.

Proof.

Since AM1,1A\in M_{1,-1} and a{1,1}a\notin\{1,-1\}, it implies that d{1,1}d\notin\{1,-1\}, and bb and cc are non-zero. Thus if we substitute these values in Equation (4), the value of vv cannot be zero unless uu is zero. Thus, we can use the same type of proof as done in Lemma 4.3. ∎

Lemma 4.6.

Let A=(1001)A=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix} and B=(10α1)B=\begin{pmatrix}-1&0\\ \alpha&1\end{pmatrix}, where α\alpha\in\mathbb{C}^{*}. Then there does not exist any XM1,1X\in M_{1,-1} such that AX=BA*X=B.

Proof.

Assume on the contrary and suppose that there exists XM1,1X\in M_{1,-1} with X=(abcd)X=\begin{pmatrix}a&b\\ c&d\end{pmatrix}, a,b,c,da,b,c,d\in\mathbb{C}, a+d=0a+d=0, adbc=1ad-bc=-1 and

(1001)(abcd)=(abcd)(10α1).\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}-1&0\\ \alpha&1\end{pmatrix}.

This implies that d=a=b=0d=a=b=0, which contradicts adbc=1ad-bc=-1. This completes the proof. ∎

Lemma 4.7.

Let γ=±1\gamma=\pm 1 and α\alpha\in\mathbb{C}. Then the following holds:

  1. (1)

    there exists XM1,1X\in M_{1,-1} such that D(γ,γ)X=(γα0γ)D(\gamma,-\gamma)*X=\begin{pmatrix}\gamma&\alpha\\ 0&-\gamma\end{pmatrix}.

  2. (2)

    there exists XM1,1X\in M_{1,-1} such that D(γ,γ)X=(γ0αγ)D(\gamma,-\gamma)*X=\begin{pmatrix}\gamma&0\\ \alpha&-\gamma\end{pmatrix}.

Proof.

We will prove (1) part. The proof of (2) is along similar lines.

Suppose A=(γα0γ)A=\begin{pmatrix}\gamma&\alpha\\ 0&-\gamma\end{pmatrix} and Xβ=(γβ0γ)X_{\beta}=\begin{pmatrix}\gamma&\beta\\ 0&-\gamma\end{pmatrix}. Then,

(γβ0γ)1(γ00γ)(γβ0γ)=(γ2β0γ)\begin{pmatrix}\gamma&\beta\\ 0&-\gamma\end{pmatrix}^{-1}\begin{pmatrix}\gamma&0\\ 0&-\gamma\end{pmatrix}\begin{pmatrix}\gamma&\beta\\ 0&-\gamma\end{pmatrix}=\begin{pmatrix}\gamma&2\beta\\ 0&-\gamma\end{pmatrix}

which is an upper triangular matrix in M1,1\operatorname{M}_{1,-1}. Thus taking β=α/2\beta=\alpha/2 implies that D(γ,γ)Xβ=A.D(\gamma,-\gamma)*X_{\beta}=A.

Theorem 4.8.

The quandle M1,1M_{1,-1} is a 22-connected quandle.

Proof.

The quandle M1,1M_{1,-1} is non-trivial (see Lemma 4.1). By Lemma 4.6, we know that M1,1M_{1,-1} is not 11-connected. Noting that if XM1,1X\in M_{1,-1}, then X1M1,1X^{-1}\in M_{1,-1}. By Theorem 2.3, we know that to prove M1,1M_{1,-1} is 22-connected, it is sufficient to prove that for any AM1,1A\in M_{1,-1}, either there exists XM1,1X\in M_{1,-1} such that D(1,1)X=AD(1,-1)*X=A or there exists Y,ZM1,1Y,Z\in M_{1,-1} such that D(1,1)YZ=AD(1,-1)*Y*Z=A. In the following points we have covered all the possibilities for AM1,1A\in M_{1,-1}.

  1. (a)

    If A=D(1,1)A=D(-1,1), then by Lemma 4.2 there exists XM1,1X\in M_{1,-1} such that D(1,1)X=D(1,1)=AD(1,-1)*X=D(-1,1)=A.

  2. (b)

    If A{(1α01),(10α1)}A\in\left\{\begin{pmatrix}1&\alpha\\ 0&-1\end{pmatrix},\begin{pmatrix}1&0\\ \alpha&-1\end{pmatrix}\right\}, where α\alpha\in\mathbb{C}, then by Lemma 4.7, there exists XM1,1X\in M_{1,-1} such that D(1,1)X=AD(1,-1)*X=A.

  3. (c)

    If A{(1α01),(10α1)}A\in\left\{\begin{pmatrix}-1&\alpha\\ 0&1\end{pmatrix},\begin{pmatrix}-1&0\\ \alpha&1\end{pmatrix}\right\}, where α\alpha\in\mathbb{C}, then by Lemma 4.7 there exists ZM1,1Z\in M_{1,-1} such that D(1,1)Z=AD(-1,1)*Z=A. Now by Case (a) we see that D(1,1)YZ=AD(1,-1)*Y*Z=A for some YM1,1Y\in M_{1,-1}.

  4. (d)

    Let A=(abcd)A=\begin{pmatrix}a&b\\ c&d\end{pmatrix}, where a{1,1}.a\notin\{1,-1\}. Then by Lemma 4.5, there exists XM1,1X\in M_{1,-1} such that D(1,1)X=AD(1,-1)*X=A.

This completes the proof. ∎

Theorem 4.9.

Let λ,λ1,λ2,k\lambda,\lambda_{1},\lambda_{2},k\in\mathbb{C}^{*}. Then, Mλ1,λ2\operatorname{M}_{\lambda_{1},\lambda_{2}} and Mkλ1,kλ2\operatorname{M}_{k\lambda_{1},k\lambda_{2}} are isomorphic quandles.

Proof.

Set

fk:Mλ1,λ2\displaystyle f_{k}\,:\,\operatorname{M}_{\lambda_{1},\lambda_{2}} Mkλ1,kλ2\displaystyle\longrightarrow\operatorname{M}_{k\lambda_{1},k\lambda_{2}}
A\displaystyle A kA\displaystyle\longmapsto kA

We leave the proof of injectivity and surjectivity to the reader and we prove that fkf_{k} is a quandle homomorphism:

fk(AB)\displaystyle f_{k}(A*B) =fk(B1AB)\displaystyle=f_{k}(B^{-1}AB)
=k(B1AB)\displaystyle=k(B^{-1}AB)
=B1k1kAkB\displaystyle=B^{-1}k^{-1}kAkB
=(kB)1(kA)(kB)\displaystyle=(kB)^{-1}(kA)(kB)
=(f(B))1f(A)f(B)\displaystyle=\big{(}f(B)\big{)}^{-1}f(A)f(B)
=f(A)f(B).\displaystyle=f(A)*f(B).

Remark 4.2.

The proof of Theorem 4.9 may suggest that the same argument is valid with kk replaced by KK, a matrix in the centralizer of quandle Mλ1,λ2M_{\lambda_{1},\lambda_{2}} (viewed as a subset of GL(2,)\operatorname{GL}(2,\mathbb{C})). We will next show that, although true, this does not contribute any new information. For λ\lambda\in\mathbb{C}^{*}, Mλ,λM_{\lambda,\lambda} has only one element. Thus, the Mλ,λM_{\lambda,\lambda}’s are isomorphic among themselves and non-isomorphic with MλM_{\lambda} and Mλ1,λ2M_{\lambda_{1},\lambda_{2}}, where λ1λ2\lambda_{1}\neq\lambda_{2}\in\mathbb{C}^{*}, since the latter are infinite quandles. Given λ1λ2\lambda_{1}\neq\lambda_{2}\in\mathbb{C}^{*} and KGL(2,)K\in\operatorname{GL}(2,\mathbb{C}), for KK to commute with D(λ1,λ2)D(\lambda_{1},\lambda_{2}) then KK has to be a diagonal matrix. To further commute with other matrices, KK also has to be a scalar matrix. But this amounts to multiplication by a scalar. This concludes this remark.

Corollary 4.10.

Let λ,k\lambda,k\in\mathbb{C}^{*}. Then MλM_{\lambda} and MkλM_{k\lambda} are isomorphic quandles.

Proof.

Analogous to the proof of Theorem 4.9. ∎

Remark 4.3.

Analogous to Remark 4.2 we note down that in Corollary 4.10 it is of no use to replace kk by a matrix KK which is in the centralizer of Mλ.M_{\lambda}.

Theorem 4.11.

The quandle Mλ,λM_{\lambda,-\lambda} is a 22-connected quandle.

Proof.

The result follows from Theorem 4.8 and Corollary 4.9. ∎

4.3. The proof that MλM_{\lambda} is 22-connected

In this subsection, we will prove that for λ\lambda\in\mathbb{C}^{*}, the quandle MλM_{\lambda} is 22-connected (see Theorem 4.15).

Lemma 4.12.

The quandle M1M_{1} is not 11-connected.

Proof.

Let (abcd)GL(2,)\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in\operatorname{GL}(2,\mathbb{C}), then

(1201)(abcd)\displaystyle\begin{pmatrix}1&2\\ 0&1\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix} =(abcd)(1101)\displaystyle=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}1&1\\ 0&1\end{pmatrix}
(a+2cb+2dcd)\displaystyle\iff\begin{pmatrix}a+2c&b+2d\\ c&d\end{pmatrix} =(aa+bcc+d)\displaystyle=\begin{pmatrix}a&a+b\\ c&c+d\end{pmatrix}

implies that d=a/2d=a/2 and c=0c=0. Now for XX to be in M1M_{1}, the trace criteria implies that a=4/3a=4/3, but the determinant criteria implies that a=±2.a=\pm\sqrt{2}. Thus XM1X\notin M_{1}, and hence M1M_{1} is not 11-connected. ∎

Lemma 4.13.

Let A=(abcd)M1A=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{1}.

  1. (1)

    If b0b\neq 0, then there exists XM1X\in M_{1} such that X1(1011)X=AX^{-1}\begin{pmatrix}1&0\\ 1&1\end{pmatrix}X=A.

  2. (2)

    If c0c\neq 0, then there exists XM1X\in M_{1} such that X1(1101)X=AX^{-1}\begin{pmatrix}1&1\\ 0&1\end{pmatrix}X=A.

Proof.

We will prove the first one, the other is similar to it.

Let y0y\neq 0, and (xyzw)M1\begin{pmatrix}x&y\\ z&w\end{pmatrix}\in M_{1}. Then

(xyzw)(αβγδ)=(1011)(xyzw)\displaystyle\begin{pmatrix}x&y\\ z&w\end{pmatrix}\begin{pmatrix}\alpha&\beta\\ \gamma&\delta\end{pmatrix}=\begin{pmatrix}1&0\\ 1&1\end{pmatrix}\begin{pmatrix}x&y\\ z&w\end{pmatrix}
\displaystyle\implies {xα+yγ=xxβ+yδ=yzα+wγ=x+zzβ+wδ=y+w\displaystyle\begin{cases}x\alpha+y\gamma=x\\ x\beta+y\delta=y\\ z\alpha+w\gamma=x+z\\ z\beta+w\delta=y+w\end{cases}

(x0y00x0yz0w00z0w)(αβγδ)=(xyx+zy+w)\begin{pmatrix}x&0&y&0\\ 0&x&0&y\\ z&0&w&0\\ 0&z&0&w\end{pmatrix}\begin{pmatrix}\alpha\\ \beta\\ \gamma\\ \delta\end{pmatrix}=\begin{pmatrix}x\\ y\\ x+z\\ y+w\end{pmatrix}

One can check that the matrix 𝒳=(x0y00x0yz0w00z0w)\mathcal{X}=\begin{pmatrix}x&0&y&0\\ 0&x&0&y\\ z&0&w&0\\ 0&z&0&w\end{pmatrix} is invertible (det𝒳=1\det\mathcal{X}=1), thus we have a unique solution for α,β,γ\alpha,\beta,\gamma and δ\delta in terms of x,y,z,wx,y,z,w. Thus

(αβγδ)=(w0y00w0yz0x00z0x)(xyx+zy+w)\begin{pmatrix}\alpha\\ \beta\\ \gamma\\ \delta\end{pmatrix}=\begin{pmatrix}w&0&-y&0\\ 0&w&0&-y\\ -z&0&x&0\\ 0&-z&0&x\end{pmatrix}\begin{pmatrix}x\\ y\\ x+z\\ y+w\end{pmatrix}

which implies

{α=1xyβ=y2γ=x2δ=1+xy\begin{cases}\alpha=1-xy\\ \beta=-y^{2}\\ \gamma=x^{2}\\ \delta=1+xy\end{cases}

Thus for any A=(abcd)M1A=\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{1} such that b0b\neq 0, there exists (xyzw)M1\begin{pmatrix}x&y\\ z&w\end{pmatrix}\in M_{1}, where either x=c,y=ιb,z=2cc1ιb,w=2cx=\sqrt{c},~{}y=\iota\sqrt{b},~{}z=\dfrac{2\sqrt{c}-c-1}{\iota\sqrt{b}}~{},w=2-\sqrt{c} or x=c,y=ιb,z=2cc1ιb,w=2cx=\sqrt{c},~{}y=-\iota\sqrt{b},~{}z=\dfrac{2\sqrt{c}-c-1}{-\iota\sqrt{b}}~{},w=2-\sqrt{c}, such that X1(1011)X=AX^{-1}\begin{pmatrix}1&0\\ 1&1\end{pmatrix}X=A. ∎

Theorem 4.14.

The quandle M1M_{1} is 22-connected.

Proof.

By Lemma 4.12, we know M1M_{1} is not 11-connected. By Lemma 4.13, we know that for any AM1A\in M_{1}, either there exists XM1X\in M_{1} such that (1011)X=A\begin{pmatrix}1&0\\ 1&1\end{pmatrix}*X=A or there exist Y,ZM1Y,Z\in M_{1} such that (1011)YZ=A\begin{pmatrix}1&0\\ 1&1\end{pmatrix}*Y*Z=A. Thus by Theorem 2.3, M1M_{1} is 22-connected. ∎

Theorem 4.15.

Let λ\lambda\in\mathbb{C}^{*}. Then MλM_{\lambda} is 22-connected.

Proof.

The result follows from Theorem 4.13 and Corollary 4.10. ∎

5. Study of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C}))

In this section we will prove the following theorems.

Theorem.

Let n2n\geq 2. Then Conj(GL(2,))Conjn(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C}))\ncong\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})).

Theorem.

For any m,n+m,n\in\mathbb{Z}^{+} with mnm\neq n, the quandle Conjm(PGL(2,))\operatorname{Conj}_{m}(\operatorname{PGL}(2,\mathbb{C})) is not isomorphic to Conjn(PGL(2,))\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C})).

Proposition 5.1.

Let Cl(A)={P1AP|PGL(2,)}\operatorname{Cl}(A)=\{P^{-1}AP~{}|~{}P\in\operatorname{GL}(2,\mathbb{C})\} and Cl(A,n)={PnAPn|PGL(2,)}\operatorname{Cl}(A,n)=\{P^{-n}AP^{n}~{}|~{}P\in\operatorname{GL}(2,\mathbb{C})\}, where n+n\in\mathbb{Z}^{+} and AGL(2,)A\in\operatorname{GL}(2,\mathbb{C}). Then Cl(A)=Cl(A,n)\operatorname{Cl}(A)=\operatorname{Cl}(A,n).

Proof.

One way is obvious. Let P1APCl(A)P^{-1}AP\in\operatorname{Cl}(A). Since PP is invertible, we can find a matrix PGL(2,)P^{\prime}\in\operatorname{GL}(2,\mathbb{C}) such that Pn=PP^{\prime n}=P (see [14, Theorem 4.1]). Thus we have P1AP=PnAPnP^{-1}AP=P^{\prime-n}AP^{\prime n}. ∎

Proposition 5.2.

For any n+n\in\mathbb{Z}^{+}, there are exactly three families of connected components in Conjn(GL(2,))\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})), namely:

  1. (1)

    for each λ1λ2\lambda_{1}\neq\lambda_{2}\in\mathbb{C}^{*}, the conjugacy class of (λ100λ2)\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} in GL(2,)\operatorname{GL}(2,\mathbb{C}),

  2. (2)

    for each λ\lambda\in\mathbb{C}^{*}, the conjugacy class of (λ10λ)\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix} in GL(2,)\operatorname{GL}(2,\mathbb{C}),

  3. (3)

    for each λ\lambda\in\mathbb{C}^{*}, the singleton set {(λ00λ)}\left\{\begin{pmatrix}\lambda&0\\ 0&\lambda\end{pmatrix}\right\}.

Proof.

Proof follows from Proposition 5.2 and considerations written down in Section 4. ∎

Lemma 5.3.

Let λ\lambda\in\mathbb{C}^{*} and n+n\in\mathbb{Z}^{+}. Then Type(Mλ)=\operatorname{Type}(M_{\lambda})=\infty, where MλM_{\lambda} is a subquandle of Conjn(GL(2,)).\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})).

Proof.

Let P=(λ01λ)P=\begin{pmatrix}\lambda&0\\ 1&\lambda\end{pmatrix}. By induction one can prove that Pn=(λn0nλn1λn)P^{n}=\begin{pmatrix}\lambda^{n}&0\\ n\lambda^{n-1}&\lambda^{n}\end{pmatrix} for all n+n\in\mathbb{Z}^{+}. Now,

(λ10λ)nPnnPmtimes\displaystyle\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}\underbrace{*_{n}P*_{n}\cdots*_{n}P}_{m~{}\textrm{times}} =Pnm(λ10λ)Pnm\displaystyle=P^{-nm}\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}P^{nm}
=(λ+(nm)/λ1(nm/λ)2λ(nm)/λ)\displaystyle=\begin{pmatrix}\lambda+(nm)/\lambda&1\\ -(nm/\lambda)^{2}&\lambda-(nm)/\lambda\end{pmatrix}
(λ10λ).\displaystyle\neq\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}.

Thus, TypeMλ=\operatorname{Type}{M_{\lambda}}=\infty. ∎

Corollary 5.4.

For each n+n\in\mathbb{Z}^{+}, Type(Conjn(GL(2,))=\operatorname{Type}(\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C}))=\infty.

Definition 5.1.

An element zz\in\mathbb{C} is said to be a root of unity if there exists n+n\in\mathbb{Z}^{+} such that zn=1z^{n}=1. In this case, we also say that zz is an nn-th root of unity.

Definition 5.2.

For p+p\in\mathbb{Z}^{+}, an element zz\in\mathbb{C} is said to be a pp-th primitive root of unity if zp=1z^{p}=1 and zk1z^{k}\neq 1 for 1kp11\leq k\leq p-1.

Lemma 5.5.

Let ω1\omega_{1} and ω2\omega_{2} be nn-th roots of unity, and λ\lambda\in\mathbb{C}^{*}. Then Mλω1,λω2M_{\lambda\omega_{1},\lambda\omega_{2}} is an nn-subquandle of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})).

Proof.

Let XMλω1,λω2X\in M_{\lambda\omega_{1},\lambda\omega_{2}}. Then there exists PGL(2,)P\in\operatorname{GL}(2,\mathbb{C}) such that

P1(λω100λω2)P=X.P^{-1}\begin{pmatrix}\lambda\omega_{1}&0\\ 0&\lambda\omega_{2}\end{pmatrix}P=X.

Since Xn=λn𝕀X^{n}=\lambda^{n}\mathbb{I}, then AXXntimes=XnAXn=AA\underbrace{*X*\cdots*X}_{n-times}=X^{-n}AX^{n}=A, for all A,XMλω1,λω2A,X\in M_{\lambda\omega_{1},\lambda\omega_{2}}. ∎

Corollary 5.6.

Let ω1\omega_{1} and ω2\omega_{2} be nn-th roots of unity, and λ\lambda\in\mathbb{C}^{*}. Then Mλω1,λω2M_{\lambda\omega_{1},\lambda\omega_{2}} is a trivial subquandle of Conjn(GL(2,))\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})).

Lemma 5.7.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}, and n+n\in\mathbb{Z}^{+}. Then Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is an nn-subquandle of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) if and only if λ1/λ2\lambda_{1}/\lambda_{2} is an nn-th root of unity.

Proof.

If λ1=λ2\lambda_{1}=\lambda_{2}, then Mλ1,λ2M_{\lambda_{1},\lambda_{2}} being a trivial subquandle implies that it is an nn-subquandle.

If λ1λ2\lambda_{1}\neq\lambda_{2} and λ1/λ2\lambda_{1}/\lambda_{2} is an nn-th root of unity, then by Lemma 5.5, Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is an nn-subquandle of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})).

Now we will prove that for λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} where λ1λ2\lambda_{1}\neq\lambda_{2}, if Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is an nn-subquandle of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})), then λ1/λ2\lambda_{1}/\lambda_{2} is an nn-th root of unity. Let X=(λ1α0λ2)X=\begin{pmatrix}\lambda_{1}&\alpha\\ 0&\lambda_{2}\end{pmatrix} and P=(λ100λ2)P=\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}, where α\alpha\in\mathbb{C}^{*}. Clearly, XX and PP are in Mλ1,λ2M_{\lambda_{1},\lambda_{2}}. By induction we have Pn=(λ1n00λ2n)P^{n}=\begin{pmatrix}\lambda_{1}^{n}&0\\ 0&\lambda_{2}^{n}\end{pmatrix}. Now

XPPntimes\displaystyle X\underbrace{*P*\cdots*P}_{n-\textrm{times}} =PnXPn\displaystyle=P^{-n}XP^{n}
=(1/(λ1λ2)n)(λ2n00λ1n)(λ1α0λ2)(λ1n00λ2n)\displaystyle=(1/(\lambda_{1}\lambda_{2})^{n})\begin{pmatrix}\lambda_{2}^{n}&0\\ 0&\lambda_{1}^{n}\end{pmatrix}\begin{pmatrix}\lambda_{1}&\alpha\\ 0&\lambda_{2}\end{pmatrix}\begin{pmatrix}\lambda_{1}^{n}&0\\ 0&\lambda_{2}^{n}\end{pmatrix}
=(λ1α(λ2/λ1)n0λ2).\displaystyle=\begin{pmatrix}\lambda_{1}&\alpha(\lambda_{2}/\lambda_{1})^{n}\\ 0&\lambda_{2}\end{pmatrix}.

For Mλ1,λ2M_{\lambda_{1},\lambda_{2}} to be nn-subquandle of Conj(GL(2,)\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C}), the equality α(λ2/λ1)n=α\alpha(\lambda_{2}/\lambda_{1})^{n}=\alpha must hold for all α\alpha\in\mathbb{C}^{*}, which implies that (λ1/λ2)n=1(\lambda_{1}/\lambda_{2})^{n}=1.

Corollary 5.8.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}, and m+m\in\mathbb{Z}^{+}. Then the type of Mλ1,λ2M_{\lambda_{1},\lambda_{2}} as a subquandle of Conjm(GL(2,))\operatorname{Conj}_{m}(\operatorname{GL}(2,\mathbb{C})) is finite if and only if λ1/λ2\lambda_{1}/\lambda_{2} is a kk-th root of unity for some k+k\in\mathbb{Z}^{+}.

Now we prove the first main theorem.

Theorem 5.9.

The quandle Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) is not isomorphic to Conjn(GL(2,))\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})) for any n2n\geq 2.

Proof.

Let us suppose there exists an isomorphism

f:Conj(GL(2,)Conjn(GL(2,)).f:\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})\to\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})).

Then the image of each connected component under ff must be a connected component and the type of each connected component must be preserved under the map ff. This implies that each connected component of Conjn(GL(2,))\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})), n2n\geq 2 having more than one element is a non-trivial subquandle (see Lemma 4.1), which is not true by Corollary 5.6 (see Proposition 5.2). Thus Conj(GL(2,))Conjn(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C}))\ncong\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C})) for any n2n\geq 2. ∎

For AGL(2,)A\in\operatorname{GL}(2,\mathbb{C}), we denote its image in PGL(2,)\operatorname{PGL}(2,\mathbb{C}) by [A][A], that is, [A]={kA|k}[A]=\{kA~{}|~{}k\in\mathbb{C}^{*}\}.

Remark 5.1.

Let [A],[B]PGL(2,)[A],[B]\in\operatorname{PGL}(2,\mathbb{C}). Then [A][A] and [B][B] are in the same conjugacy class if and only if AA is conjugate to kBkB in GL(2,)\operatorname{GL}(2,\mathbb{C}) for some kk\in\mathbb{C}^{*}. This implies that following are the only families of distinct conjugacy classes in PGL(2,)\operatorname{PGL}(2,\mathbb{C}):

  1. (1)

    The set M[λ1,λ2]={[A]|AMkλ1,kλ2 for some k}M_{[\lambda_{1},\lambda_{2}]}=\{[A]~{}|~{}A\in M_{k\lambda_{1},k\lambda_{2}}~{}\textrm{ for some }~{}k\in\mathbb{C}^{*}\}, where λ1\lambda_{1} and λ2\lambda_{2} are distinct elements in \mathbb{C}^{*}.

  2. (2)

    M[λ]={[A]|AMkλ for some k}M_{[\lambda]}=\{[A]~{}|~{}A\in M_{k\lambda}~{}\textrm{ for some }~{}k\in\mathbb{C}^{*}\}, where λ\lambda\in\mathbb{C}^{*}.

  3. (3)

    the singleton set containing the identity element of PGL(2,)\operatorname{PGL}(2,\mathbb{C}).

We note down the following results for PGL(2,)\operatorname{PGL}(2,\mathbb{C}) without writing their proofs as they can be proved along similar lines as done in the case of GL(2,)\operatorname{GL}(2,\mathbb{C}), see above in this section.

Proposition 5.10.

For any n+n\in\mathbb{Z}^{+}, the only connected components in Conjn(PGL(2,))\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C})) are M[λ1,λ2]M_{[\lambda_{1},\lambda_{2}]}, M[λ]M_{[\lambda]} and the identity element, where λ1,λ2,λ\lambda_{1},\lambda_{2},\lambda\in\mathbb{C}^{*} and λ1λ2\lambda_{1}\neq\lambda_{2}.

Lemma 5.11.

Let λ\lambda\in\mathbb{C}^{*} and n+n\in\mathbb{Z}^{+}. Then Type(M[λ])=\operatorname{Type}(M_{[\lambda]})=\infty, where M[λ]M_{[\lambda]} is a subquandle of Conjn(PGL(2,))\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C})).

Corollary 5.12.

For each n+n\in\mathbb{Z}^{+}, Type(Conjn(PGL(2,))=\operatorname{Type}(\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C}))=\infty.

Lemma 5.13.

Each connected component of Conj(PGL(2,))\operatorname{Conj}(\operatorname{PGL}(2,\mathbb{C})) having more than one element is a non-trivial subquandle.

Lemma 5.14.

Let λ1\lambda_{1} and λ2\lambda_{2} be in \mathbb{C}^{*}, and n+n\in\mathbb{Z}^{+}. Then M[λ1,λ2]M_{[\lambda_{1},\lambda_{2}]} is an nn-subquandle of Conj(PGL(2,))\operatorname{Conj}(\operatorname{PGL}(2,\mathbb{C})) if and only if λ1/λ2\lambda_{1}/\lambda_{2} is an nn-th root of unity.

Lemma 5.15.

The quandle Conjn(PGL(2,))\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C})) has exactly nn connected components which are trivial subquandles.

Proof.

The connected components in Conjn(PGL(2,))\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C})) which are trivial are exactly those connected components in Conj(PGL(2,))\operatorname{Conj}(\operatorname{PGL}(2,\mathbb{C})) which are nn-subquandles. The number of ways one can choose two nn-th roots of unity such that one of them is identity is nn. Thus, by Lemma 5.14 there are only nn connected components in Conj(PGL(2,))\operatorname{Conj}(\operatorname{PGL}(2,\mathbb{C})) which are nn-subquandles. ∎

Thus we have the second main theorem.

Theorem 5.16.

For any m,n+m,n\in\mathbb{Z}^{+} with mnm\neq n, the quandle Conjm(PGL(2,))\operatorname{Conj}_{m}(\operatorname{PGL}(2,\mathbb{C})) is not isomorphic to Conjn(PGL(2,))\operatorname{Conj}_{n}(\operatorname{PGL}(2,\mathbb{C})).

Proof.

The result follows from Lemma 5.15. ∎

Question 5.17.

Let m,n+m,n\in\mathbb{Z}^{+} such that mnm\neq n. Is it true that Conjm(GL(2,))Conjn(GL(2,))\operatorname{Conj}_{m}(\operatorname{GL}(2,\mathbb{C}))\ncong\operatorname{Conj}_{n}(\operatorname{GL}(2,\mathbb{C}))?

6. Classification of connected components of Conj(GL(2,)\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})

In Section 4, we proved that the non-trivial connected components of Conj(GL(2,))\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})) are 22-connected (see Theorem 4.4, 4.11, 4.15), and hence they are connected subquandles. In this section, we partially classify these connected components.

Theorem 6.1.

Let λ,λ1,λ2\lambda,\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}. Then MλM_{\lambda} is not isomorphic to Mλ1,λ2M_{\lambda_{1},\lambda_{2}}.

Proof.

Assume on the contrary and suppose that f:MλMλ1,λ2f:M_{\lambda}\to M_{\lambda_{1},\lambda_{2}} is a quandle isomorphism. Since (λ10λ)\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}, (λ20λ)\begin{pmatrix}\lambda&2\\ 0&\lambda\end{pmatrix} and (λ30λ)\begin{pmatrix}\lambda&3\\ 0&\lambda\end{pmatrix} commute with one another, so do their images under the map ff (as ff is a quandle homomorphism). Thus there exists SGL(2,)S\in\operatorname{GL}(2,\mathbb{C}) which simultaneously diagonalize f((λ10λ))f(\begin{pmatrix}\lambda&1\\ 0&\lambda\end{pmatrix}), f((λ20λ))f(\begin{pmatrix}\lambda&2\\ 0&\lambda\end{pmatrix}), and f((λ30λ))f(\begin{pmatrix}\lambda&3\\ 0&\lambda\end{pmatrix}) (see [28, Page 569]). But since there are only two diagonal matrices in Mλ1,λ2M_{\lambda_{1},\lambda_{2}} and ff is a bijection, this is absurd. The proof is complete. ∎

Theorem 6.2.

Let λ1,λ2,λ1,λ2\lambda_{1},\lambda_{2},\lambda_{1}^{\prime},\lambda_{2}^{\prime}\in\mathbb{C} and p,q+p,q\in\mathbb{Z}^{+} where pqp\neq q. Then the following holds:

  1. (1)

    If λ2/λ1\lambda_{2}/\lambda_{1} is a pp-th primitive root of unity and λ2/λ1\lambda_{2}^{\prime}/\lambda_{1}^{\prime} is not a root of unity, then Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is not isomorphic to Mλ1,λ2M_{\lambda_{1}^{\prime},\lambda_{2}^{\prime}}.

  2. (2)

    If λ2/λ1\lambda_{2}/\lambda_{1} is pp-th primitive root of unity and λ2/λ1\lambda_{2}^{\prime}/\lambda_{1}^{\prime} is qq-th primitive root of unity, then Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is not isomorphic to Mλ1,λ2M_{\lambda_{1}^{\prime},\lambda_{2}^{\prime}}.

Proof.

The proof of (1) follows from Lemma 5.7.

Now for the proof of (2). Without loss of generality we assume q<pq<p. By Lemma 5.7 and Theorem 4.9, it is sufficient to prove that M1,ωM_{1,\omega} is not isomorphic to M1,ωM_{1,\omega^{\prime}}, where ω\omega and ω\omega^{\prime} are pp-th and qq-th primitive roots of unity, respectively. Note that M1,ωM_{1,\omega} is pp-quandle and M1,ωM_{1,\omega^{\prime}} is qq-quandle. Now let X=(100w)X=\begin{pmatrix}1&0\\ 0&w\end{pmatrix} which is an element in M1,ωM_{1,\omega}. Then

(110ω)qX\displaystyle\begin{pmatrix}1&1\\ 0&\omega\end{pmatrix}*^{q}X =Xq(110ω)Xq\displaystyle=X^{-q}\begin{pmatrix}1&1\\ 0&\omega\end{pmatrix}X^{q}
=(1wq0w)\displaystyle=\begin{pmatrix}1&w^{q}\\ 0&w\end{pmatrix}
(110ω),\displaystyle\neq\begin{pmatrix}1&1\\ 0&\omega\end{pmatrix},

as q<pq<p and ω\omega is a pp-th primitive root of unity. Thus M1,ωM_{1,\omega} is not a qq-quandle and hence M1,ωM_{1,\omega} is not isomorphic to M1,ωM_{1,\omega^{\prime}}. ∎

Lemma 6.3.

Let λ,λ\lambda,\lambda^{\prime}\in\mathbb{C} and M1,λM_{1,\lambda} be quandle isomorphic to M1,λM_{1,\lambda^{\prime}}. If λ\lambda is not a root of unity, then M1,λnM_{1,\lambda^{n}} is isomorphic to M1,λnM_{1,\lambda^{\prime n}} for all n+n\in\mathbb{Z}^{+}.

Proof.

Let h1:M1,λM1,λh_{1}:M_{1,\lambda}\to M_{1,\lambda^{\prime}} be a quandle isomorphism. Now for n+n\in\mathbb{Z}^{+}, define

hn:M1,λn\displaystyle h_{n}:M_{1,\lambda^{n}} M1,λn as\displaystyle\to M_{1,\lambda^{\prime n}}~{}\textrm{ as }
B\displaystyle B h1(A)n,\displaystyle\mapsto h_{1}(A)^{n},

where BM1,λnB\in M_{1,\lambda^{n}}, B=AnB=A^{n}, AM1,λA\in M_{1,\lambda}. Note that such AA always exists, see [14].

Now we will prove that hnsh_{n}^{\prime}s are well-defined and are bijective. Suppose A1,A2M1,λA_{1},A_{2}\in M_{1,\lambda} such that A1n=A2nA_{1}^{n}=A_{2}^{n}. Then there exist P1,P2GL(2,)P_{1},P_{2}\in\operatorname{GL}(2,\mathbb{C}) such that A1=P11D(1,λ)P1A_{1}=P_{1}^{-1}D(1,\lambda)P_{1} and A2=P21D(1,λ)P2A_{2}=P_{2}^{-1}D(1,\lambda)P_{2}. Then A1n=A2nA_{1}^{n}=A_{2}^{n} implies that

(P11D(1,λ)P1)n\displaystyle(P_{1}^{-1}D(1,\lambda)P_{1})^{n} =(P21D(1,λ)P2)n\displaystyle=(P_{2}^{-1}D(1,\lambda)P_{2})^{n}
P11D(1,λn)P1\displaystyle P_{1}^{-1}D(1,\lambda^{n})P_{1} =P21D(1,λn)P2\displaystyle=P_{2}^{-1}D(1,\lambda^{n})P_{2}
(P1P21)1D(1,λn)(P1P21)\displaystyle(P_{1}P_{2}^{-1})^{-1}D(1,\lambda^{n})(P_{1}P_{2}^{-1}) =D(1,λn)\displaystyle=D(1,\lambda^{n})
((P1P21)1D(1,λ)(P1P21))n\displaystyle\big{(}(P_{1}P_{2}^{-1})^{-1}D(1,\lambda)(P_{1}P_{2}^{-1})\big{)}^{n} =D(1,λn),\displaystyle=D(1,\lambda^{n}),

which further implies that (P1P21)1D(1,λ)(P1P21)=D(ω,ωλ)(P_{1}P_{2}^{-1})^{-1}D(1,\lambda)(P_{1}P_{2}^{-1})=D(\omega,\omega^{\prime}\lambda) (see Theorem [14, Theorem 4.1] and noting that any 22-by-22 matrix with distinct non-zero eigen values have four square roots) where ω\omega and ω\omega^{\prime} are nn-th roots of unity. Noting that under similar transformation, eigen values are preserved and λ\lambda is not a root of unity, thus (P1P21)1D(1,λ)(P1P21)=D(1,λ)(P_{1}P_{2}^{-1})^{-1}D(1,\lambda)(P_{1}P_{2}^{-1})=D(1,\lambda) and hence A1=A2A_{1}=A_{2}. Thus each matrix in M1,λnM_{1,\lambda^{n}} has a unique root in M1,λM_{1,\lambda}, and as a consequence the maps hn:M1,λnM1,λnh_{n}:M_{1,\lambda^{n}}\to M_{1,\lambda^{\prime n}} are well-defined for all n+n\in\mathbb{Z}^{+}. Furthermore, since h1h_{1} is a quandle isomorphism, thus λ\lambda^{\prime} is not a root of unity, and therefore by the preceding kind of reasoning, the hnh_{n} maps are injective. It is trivial to see that the hnh_{n} maps are surjective.

Now we are left with the proof that these hnh_{n} maps are quandle isomorphisms. Let P,QM1,λnP,Q\in M_{1,\lambda^{n}}. Then there exist unique A,BM1,λA,B\in M_{1,\lambda} such that An=PA^{n}=P and Bn=QB^{n}=Q. Now

hn(PQ)\displaystyle h_{n}(P*Q) =hn(AnBn)\displaystyle=h_{n}(A^{n}*B^{n})
=hn(BnAnBn)\displaystyle=h_{n}(B^{-n}A^{n}B^{n})
=hn((BnABn)n)\displaystyle=h_{n}\big{(}(B^{-n}AB^{n})^{n}\big{)}
=(h1(BnABn))n\displaystyle=\big{(}h_{1}(B^{-n}AB^{n})\big{)}^{n}
=(h1(AnB))n\displaystyle=\big{(}h_{1}(A*^{n}B)\big{)}^{n}
=(h1(A)nh1(B))n\displaystyle=\big{(}h_{1}(A)*^{n}h_{1}(B)\big{)}^{n}
=(h1(B)nh1(A)h1(B)n)n\displaystyle=\big{(}h_{1}(B)^{-n}h_{1}(A)h_{1}(B)^{n}\big{)}^{n}
=h1(B)nh1(A)nh1(B)n\displaystyle=h_{1}(B)^{-n}h_{1}(A)^{n}h_{1}(B)^{n}
=hn(Bn)1hn(An)hn(Bn)\displaystyle=h_{n}(B^{n})^{-1}h_{n}(A^{n})h_{n}(B^{n})
=hn(An)hn(Bn)\displaystyle=h_{n}(A^{n})*h_{n}(B^{n})
=hn(P)hn(Q).\displaystyle=h_{n}(P)*h_{n}(Q).

This completes the proof. ∎

Theorem 6.4.

Let the quandle Mλ1,λ2M_{\lambda_{1},\lambda_{2}} be isomorphic to Mλ1,λ2M_{\lambda_{1}^{\prime},\lambda_{2}^{\prime}}, where λ2/λ1\lambda_{2}/\lambda_{1} is not a root of unity. Then Mλ1n,λ2nM_{\lambda_{1}^{n},\lambda_{2}^{n}} is isomorphic to Mλ1n,λ2nM_{\lambda_{1}^{\prime n},\lambda_{2}^{\prime n}}, for all n+n\in\mathbb{Z}^{+}.

Proof.

The proof follows from Theorem 4.9 and Lemma 6.3. ∎

Corollary 6.5.

Let λ2/λ1\lambda_{2}/\lambda_{1} and λ2/λ1\lambda_{2}^{\prime}/\lambda_{1}^{\prime} not be roots of unity, and Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is not isomorphic to Mλ1,λ2M_{\lambda_{1}^{\prime},\lambda_{2}^{\prime}}. Then Mλ11/n,λ21/nM_{\lambda_{1}^{1/n},\lambda_{2}^{1/n}} is not isomorphic to Mλ11/n,λ21/n,M_{\lambda_{1}^{\prime 1/n},\lambda_{2}^{\prime 1/n}}, for any n+n\in\mathbb{Z}^{+}.

Proposition 6.6.

For λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} and λ1λ2\lambda_{1}\neq\lambda_{2}, the maximal trivial subquandle in Mλ1,λ2M_{\lambda_{1},\lambda_{2}} is of cardinality 22.

Proof.

Clearly the set {(λ100λ2),(λ200λ1)}\left\{\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix},\begin{pmatrix}\lambda_{2}&0\\ 0&\lambda_{1}\end{pmatrix}\right\} is a trivial subquandle of Mλ1,λ2M_{\lambda_{1},\lambda_{2}} of cardinality 22. Now suppose TT is a trivial subquandle of Mλ1,λ2M_{\lambda_{1},\lambda_{2}}, having more than two elements. Since TT is trivial, thus the elements of TT commutes pair wise which further implies that they are simultaneous diagonalizable (see [28, Page 569]), which is a contradiction. ∎

Since the set of upper triangular matrices in MλM_{\lambda} is a commutative set, we have the following result.

Proposition 6.7.

The maximal cardinality of a trivial quandle in MλM_{\lambda} is equal to the cardinality of \mathbb{R}.

For λ1λ2\lambda_{1}\neq\lambda_{2}\in\mathbb{C}^{*}, D(λ1,λ2)D(λ2,λ1)D(λ1,λ2)=D(λ1,λ2)D(\lambda_{1},\lambda_{2})*D(\lambda_{2},\lambda_{1})*D(\lambda_{1},\lambda_{2})=D(\lambda_{1},\lambda_{2}) in Mλ1,λ2M_{\lambda_{1},\lambda_{2}}.

Lemma 6.8.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}, and λ1λ2\lambda_{1}\neq\lambda_{2}, λ2±1\lambda_{2}\neq\pm 1, λ1λ2\lambda_{1}\neq-\lambda_{2}, λ1λ22\lambda_{1}\neq\lambda_{2}^{2}. Then there exist A,BMλ1,λ2A,B\in M_{\lambda_{1},\lambda_{2}} such that ABBAAB\neq BA and D(λ1,λ2)AB=D(λ1,λ2)D(\lambda_{1},\lambda_{2})*A*B=D(\lambda_{1},\lambda_{2}).

Proof.

Let A=(abcd),B=(efgh)Mλ1,λ2A=\begin{pmatrix}a&b\\ c&d\end{pmatrix},B=\begin{pmatrix}e&f\\ g&h\end{pmatrix}\in M_{\lambda_{1},\lambda_{2}}, such that D(λ1,λ2)AB=D(λ1,λ2)D(\lambda_{1},\lambda_{2})*A*B=D(\lambda_{1},\lambda_{2}). As λ1λ2\lambda_{1}\neq\lambda_{2}, thus ABAB is a diagonal matrix. Now for λμ\lambda\neq\mu\in\mathbb{C}^{*}, BA=D(λ,μ)BA=D(\lambda,\mu)

(efgh)(abcd)=(λ00μ)\displaystyle\iff\begin{pmatrix}e&f\\ g&h\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}\lambda&0\\ 0&\mu\end{pmatrix}
(abcd)=(λ00μ)(hfge)1λ1λ2\displaystyle\iff\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}\lambda&0\\ 0&\mu\end{pmatrix}\begin{pmatrix}h&-f\\ -g&e\end{pmatrix}\frac{1}{\lambda_{1}\lambda_{2}}
{a=hλλ1λ2b=fλλ1λ2c=gμλ1λ2d=eμλ1λ2\displaystyle\iff\begin{dcases}a=\frac{h\lambda}{\lambda_{1}\lambda_{2}}\\ b=\frac{-f\lambda}{\lambda_{1}\lambda_{2}}\\ c=\frac{-g\mu}{\lambda_{1}\lambda_{2}}\\ d=\frac{e\mu}{\lambda_{1}\lambda_{2}}\end{dcases} (5)

Now onwards, assume BA=D(λ,μ)BA=D(\lambda,\mu). Since A,BMλ1,λ2A,B\in M_{\lambda_{1},\lambda_{2}}, BA=D(λ,μ)BA=D(\lambda,\mu) the determinant argument implies

λ1λ2=adbc=μλheμλfg(λ1λ2)2\displaystyle\lambda_{1}\lambda_{2}=ad-bc=\frac{\mu\lambda he-\mu\lambda fg}{(\lambda_{1}\lambda_{2})^{2}}
\displaystyle\implies μλ=(λ1λ2)2\displaystyle\mu\lambda=(\lambda_{1}\lambda_{2})^{2} (6)

and the trace argument implies

λ1+λ2=a+d=λh+μeλ1λ2\displaystyle\lambda_{1}+\lambda_{2}=a+d=\frac{\lambda h+\mu e}{\lambda_{1}\lambda_{2}}
\displaystyle\implies λh+μe=λ1λ2(λ1+λ2)\displaystyle\lambda h+\mu e=\lambda_{1}\lambda_{2}(\lambda_{1}+\lambda_{2})
\displaystyle\implies λh+μ(λ1+λ2h)=λ1λ2(λ1+λ2)\displaystyle\lambda h+\mu(\lambda_{1}+\lambda_{2}-h)=\lambda_{1}\lambda_{2}(\lambda_{1}+\lambda_{2})
\displaystyle\implies {h=(λ1λ2μ)(λ1+λ2)λμ (here we used λμ)e=(λ1+λ2)(λ1λ2λ)μλfg=(λ1+λ2)2(λ1λ2μ)(λ1λ2λ)+λ1λ2(μλ)2(μλ)2\displaystyle\begin{dcases}h=\frac{(\lambda_{1}\lambda_{2}-\mu)(\lambda_{1}+\lambda_{2})}{\lambda-\mu}~{}{\textrm{ (here we used }\lambda\neq\mu)}\\ e=\frac{(\lambda_{1}+\lambda_{2})(\lambda_{1}\lambda_{2}-\lambda)}{\mu-\lambda}\\ fg=-\frac{(\lambda_{1}+\lambda_{2})^{2}(\lambda_{1}\lambda_{2}-\mu)(\lambda_{1}\lambda_{2}-\lambda)+\lambda_{1}\lambda_{2}(\mu-\lambda)^{2}}{(\mu-\lambda)^{2}}\end{dcases} (7)

Now onwards take μ=λ1\mu=\lambda_{1} and λ=λ1λ22\lambda=\lambda_{1}\lambda_{2}^{2} ( λ2±1\lambda_{2}\neq\pm 1 is used here, as μλ\mu\neq\lambda) . Thus

{e=(λ1+λ2)λ21+λ2h=λ1+λ21+λ2a=eb=fλ2c=gλ2d=h\displaystyle\begin{dcases}e=\frac{(\lambda_{1}+\lambda_{2})\lambda_{2}}{1+\lambda_{2}}\\ h=\frac{\lambda_{1}+\lambda_{2}}{1+\lambda_{2}}\\ a=e\\ b=-f\lambda_{2}\\ c=-\frac{g}{\lambda_{2}}\\ d=h\end{dcases} (8)

and

fg=λ2(λ1+λ2)2+λ1(1+λ2)2(1+λ2)2\displaystyle fg=-\lambda_{2}\frac{-(\lambda_{1}+\lambda_{2})^{2}+\lambda_{1}(1+\lambda_{2})^{2}}{(1+\lambda_{2})^{2}}~{}~{} (9)

Thus for suitable values of ff and gg, there exist A,BMλ1,λ2A,B\in M_{\lambda_{1},\lambda_{2}} such that D(λ1,λ2)AB=D(λ1,λ2)D(\lambda_{1},\lambda_{2})*A*B=D(\lambda_{1},\lambda_{2}). We ar now left to prove that ABBAAB\neq BA. Now

AB=BA\displaystyle AB=BA
\displaystyle\iff (abcd)(afgd)=(afgd)(abcd)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}a&f\\ g&d\end{pmatrix}=\begin{pmatrix}a&f\\ g&d\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}
\displaystyle\iff {bg=fcaf+bd=ab+fdac+dg=ag+cd\displaystyle\begin{dcases}bg=fc\\ af+bd=ab+fd\\ ac+dg=ag+cd\end{dcases}

Since λ1λ2\lambda_{1}\neq-\lambda_{2} and λ20,±1\lambda_{2}\neq 0,\pm 1, the above equations implies that fg=0fg=0. Now

fg=0\displaystyle fg=0
\displaystyle\iff (λ1+λ2)2+λ1(1+λ2)2=0(using Equation(9))\displaystyle-(\lambda_{1}+\lambda_{2})^{2}+\lambda_{1}(1+\lambda_{2})^{2}=0{~{}~{}(\textrm{using Equation\eqref{6.9})}}
\displaystyle\iff either λ1=1 or λ1=λ22.\displaystyle\textrm{ either }{\lambda_{1}=1}\textrm{ or }\lambda_{1}=\lambda_{2}^{2}.

Since it is given that neither λ1=1\lambda_{1}=1 nor λ1=λ22\lambda_{1}=\lambda_{2}^{2}, we have fg0fg\neq 0. This further implies that for the values in Equation (8), we have ABBAAB\neq BA and D(λ1,λ2)AB=D(λ1,λ2).D(\lambda_{1},\lambda_{2})*A*B=D(\lambda_{1},\lambda_{2}).

Theorem 6.9.

Let λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} such that λ1±λ2\lambda_{1}\neq\pm\lambda_{2}. Then there exist A,BMλ1,λ2A,B\in M_{\lambda_{1},\lambda_{2}} such that ABBAAB\neq BA and D(λ1,λ2)AB=D(λ1,λ2)D(\lambda_{1},\lambda_{2})*A*B=D(\lambda_{1},\lambda_{2}).

Proof.

By Lemma 6.8, we need to establish the statement only for the following pairs of eigen values:

  1. (1)

    for (±1,λ2)(\pm 1,\lambda_{2}), where λ2±1\lambda_{2}\neq\pm 1,

  2. (2)

    for (λ1,±1)(\lambda_{1},\pm 1), where λ1±1\lambda_{1}\neq\pm 1,

  3. (3)

    for λ1=λ22\lambda_{1}=\lambda_{2}^{2}, where λ2±1\lambda_{2}\neq\pm 1.

Here we are proving the statement for the above cases:

  • 1)

    For (±1,λ2)(\pm 1,\lambda_{2}), where λ2±1:\lambda_{2}\neq\pm 1: Pick k{±1,1/λ22}k\in\mathbb{C}^{*}\setminus\{\pm 1,1/\lambda_{2}^{2}\}. Then by Lemma 6.8, the statement holds for M±k,kλ2M_{\pm k,k\lambda_{2}}, thus by Theorem 4.9, the statement holds for M±1,λ2M_{\pm 1,\lambda_{2}}.

  • 2)

    For (λ1,±1)(\lambda_{1},\pm 1), where λ1±1:\lambda_{1}\neq\pm 1: Pick k{±1,1/λ1,λ1}k\in\mathbb{C}^{*}\setminus\{\pm 1,1/\lambda_{1},\lambda_{1}\}. Then by Lemma 6.8, the statement holds for M±kλ1,±kM_{\pm k\lambda_{1},\pm k}, thus by Theorem 4.9, the statement holds for Mλ1,±1M_{\lambda_{1},\pm 1}.

  • 3)

    For λ1=λ22\lambda_{1}=\lambda_{2}^{2}, where λ2±1\lambda_{2}\neq\pm 1: Pick k{±1,1/(λ2)2}k\in\mathbb{C}^{*}\setminus\{\pm 1,1/(\lambda_{2})^{2}\}. Then by Lemma 6.8, the statement holds for Mkλ22,kλ2M_{k\lambda_{2}^{2},k\lambda_{2}}, thus by Theorem 4.9, the statement holds for Mλ22,λ2M_{\lambda_{2}^{2},\lambda_{2}}.

Theorem 6.10.

Let R3\operatorname{R}_{3} be the dihedral quandle of order 33. Then

  1. (1)

    R3\operatorname{R}_{3} is not a subquandle of Mλ1,λ2M_{\lambda_{1},\lambda_{2}} for λ1±λ2\lambda_{1}\neq\pm\lambda_{2}\in\mathbb{C}^{*};

  2. (2)

    R3\operatorname{R}_{3} is not a subquandle of MλM_{\lambda} for λ\lambda\in\mathbb{C}^{*}.

Proof.

We will first prove (1). Given λ1±λ2\lambda_{1}\neq\pm\lambda_{2}\in\mathbb{C}^{*}, we will consider, without loss of generality, an element (abcd)Mλ1,λ2\begin{pmatrix}a&b\\ c&d\end{pmatrix}\in M_{\lambda_{1},\lambda_{2}}, and together with (λ100λ2)\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} and (abcd)(λ100λ2)\begin{pmatrix}a&b\\ c&d\end{pmatrix}*\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}, we will force these three elements to constitute the underlying set of an isomorph of R3\operatorname{R}_{3}. We note that, for R3\operatorname{R}_{3}, when two distinct elements are operated the result is the other element. Note that

(abcd)(λ100λ2)=(abλ2/λ1cλ1/λ2d)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}*\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}=\begin{pmatrix}a&b\lambda_{2}/\lambda_{1}\\ c\lambda_{1}/\lambda_{2}&d\end{pmatrix}

We now force the equality:

(abcd)(abλ2/λ1cλ1/λ2d)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}*\begin{pmatrix}a&b\lambda_{2}/\lambda_{1}\\ c\lambda_{1}/\lambda_{2}&d\end{pmatrix} =(λ100λ2)\displaystyle=\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} (10)
\displaystyle\iff (abcd)(abλ2/λ1cλ1/λ2d)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}\begin{pmatrix}a&b\lambda_{2}/\lambda_{1}\\ c\lambda_{1}/\lambda_{2}&d\end{pmatrix} =(abλ2/λ1cλ1/λ2d)(λ100λ2)\displaystyle=\begin{pmatrix}a&b\lambda_{2}/\lambda_{1}\\ c\lambda_{1}/\lambda_{2}&d\end{pmatrix}\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}
\displaystyle\iff (a2+bcλ1/λ2b(aλ2/λ1+d)c(a+dλ1/λ2)bcλ2/λ1+d2)\displaystyle\begin{pmatrix}a^{2}+bc\lambda_{1}/\lambda_{2}&b(a\lambda_{2}/\lambda_{1}+d)\\ c(a+d\lambda_{1}/\lambda_{2})&bc\lambda_{2}/\lambda_{1}+d^{2}\end{pmatrix} =(aλ1bλ22/λ1cλ12/λ2dλ2)\displaystyle=\begin{pmatrix}a\lambda_{1}&b\lambda_{2}^{2}/\lambda_{1}\\ c\lambda_{1}^{2}/\lambda_{2}&d\lambda_{2}\end{pmatrix}
\displaystyle\iff {a2+bcλ1/λ2=aλ1b(aλ2/λ1+d)=bλ22/λ1c(a+dλ1/λ2)=cλ12/λ2bcλ2/λ1+d2=dλ2\displaystyle\begin{cases}a^{2}+bc\lambda_{1}/\lambda_{2}=a\lambda_{1}\\ b(a\lambda_{2}/\lambda_{1}+d)=b\lambda_{2}^{2}/\lambda_{1}\\ c(a+d\lambda_{1}/\lambda_{2})=c\lambda_{1}^{2}/\lambda_{2}\\ bc\lambda_{2}/\lambda_{1}+d^{2}=d\lambda_{2}\end{cases}

if and only if

a2+bcλ1/λ2=aλ1\displaystyle a^{2}+bc\lambda_{1}/\lambda_{2}=a\lambda_{1} (11)
b=0aλ2/λ1+d=λ22/λ1\displaystyle b=0\,\lor\,a\lambda_{2}/\lambda_{1}+d=\lambda_{2}^{2}/\lambda_{1} (12)
c=0a+dλ1/λ2=λ12/λ2\displaystyle c=0\,\lor\,a+d\lambda_{1}/\lambda_{2}=\lambda_{1}^{2}/\lambda_{2} (13)
bcλ2/λ1+d2=dλ2\displaystyle bc\lambda_{2}/\lambda_{1}+d^{2}=d\lambda_{2} (14)

We now explore the consequences of b=0b=0 or c=0c=0. In Equations (11) and (14), we now have:

{a2=aλ1d2=dλ2{a=0a=λ1d=0d=λ2\displaystyle\begin{cases}a^{2}=a\lambda_{1}\\ d^{2}=d\lambda_{2}\end{cases}\iff\begin{cases}a=0\,\lor\,a=\lambda_{1}\\ d=0\,\lor\,d=\lambda_{2}\end{cases}
\displaystyle\iff [a=0d=0][a=0d=λ2][a=λ1d=0][a=λ1d=λ2]\displaystyle[a=0\,\land\,d=0]\,\lor\,[a=0\,\land\,d=\lambda_{2}]\,\lor\,[a=\lambda_{1}\,\land\,d=0]\,\lor\,[a=\lambda_{1}\,\land\,d=\lambda_{2}]

The first three instances, along with bc=0bc=0, imply

λ1λ2=det(abcd)=0\lambda_{1}\lambda_{2}=\det\begin{pmatrix}a&b\\ c&d\end{pmatrix}=0

which is impossible since λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}. The instance a=λ1d=λ2a=\lambda_{1}\,\land\,d=\lambda_{2} gives rise to the following three instances:

  1. 1.

    (abcd)=(λ100λ2)\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix} which implies that the prospective isomorph to R3\operatorname{R}_{3} has at most two elements and therefore cannot be an isomorph to R3\operatorname{R}_{3}.

  2. 2.

    (abcd)=(λ1b0λ2)\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}\lambda_{1}&b\\ 0&\lambda_{2}\end{pmatrix} with b0b\neq 0 but by performing the operations with the other two elements we do not obtain a quandle structure with these three elements. Namely, forcing

    (λ1bλ2/λ10λ2)=(λ100λ2)(λ1b0λ2)\displaystyle\begin{pmatrix}\lambda_{1}&b\lambda_{2}/\lambda_{1}\\ 0&\lambda_{2}\end{pmatrix}=\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}*\begin{pmatrix}\lambda_{1}&b\\ 0&\lambda_{2}\end{pmatrix}

    implies

    λ2=λ1λ2λ1=2λ2.\lambda_{2}=\lambda_{1}-\lambda_{2}\qquad\iff\lambda_{1}=2\lambda_{2}.

    Forcing

    (λ1b0λ2)=(λ100λ2)(λ1bλ2/λ10λ2)\displaystyle\begin{pmatrix}\lambda_{1}&b\\ 0&\lambda_{2}\end{pmatrix}=\begin{pmatrix}\lambda_{1}&0\\ 0&\lambda_{2}\end{pmatrix}*\begin{pmatrix}\lambda_{1}&b\lambda_{2}/\lambda_{1}\\ 0&\lambda_{2}\end{pmatrix}

    implies

    λ12+λ22=λ1λ2\lambda_{1}^{2}+\lambda_{2}^{2}=\lambda_{1}\lambda_{2}. Now using λ1=2λ2\lambda_{1}=2\lambda_{2}, we get 4=14=1. which is absurd.

  3. 3.

    (abcd)=(λ10cλ2)\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}\lambda_{1}&0\\ c&\lambda_{2}\end{pmatrix} with c0c\neq 0 with the same remark as for the previous item, arguing along the same lines.

We resume the study of Equations (11), (12), (13) and (14), now certain that b0b\neq 0 and c0c\neq 0:

(){a2+bcλ1/λ2=aλ1aλ2/λ1+d=λ22/λ1a+dλ1/λ2=λ12/λ2bcλ2/λ1+d2=dλ2{aλ2+dλ1=λ22aλ2+dλ1=λ12λ22=λ12λ2=±λ1\displaystyle(\dagger\dagger)\begin{cases}a^{2}+bc\lambda_{1}/\lambda_{2}=a\lambda_{1}\\ a\lambda_{2}/\lambda_{1}+d=\lambda_{2}^{2}/\lambda_{1}\\ a+d\lambda_{1}/\lambda_{2}=\lambda_{1}^{2}/\lambda_{2}\\ bc\lambda_{2}/\lambda_{1}+d^{2}=d\lambda_{2}\end{cases}\Longrightarrow\begin{cases}a\lambda_{2}+d\lambda_{1}=\lambda_{2}^{2}\\ a\lambda_{2}+d\lambda_{1}=\lambda_{1}^{2}\\ \end{cases}\Longrightarrow\,\lambda_{2}^{2}=\lambda_{1}^{2}\,\Longrightarrow\,\lambda_{2}=\pm\lambda_{1}

but λ2=±λ1\lambda_{2}=\pm\lambda_{1} is impossible by hypothesis. This concludes the proof of (1).

We now prove (2). Since M1M_{1} is isomorphic to MλM_{\lambda} for any λ\lambda\in\mathbb{C}^{*} (see Corollary 4.10), we will work with M1M_{1}. Given (abcd)\begin{pmatrix}a&b\\ c&d\end{pmatrix}, without loss of generality, we will impose an R3\operatorname{R}_{3} quandle structure on the elements

(1101),(abcd),(abcd)(1101)=(acac+bdcc+d)\begin{pmatrix}1&1\\ 0&1\end{pmatrix},\begin{pmatrix}a&b\\ c&d\end{pmatrix},\begin{pmatrix}a&b\\ c&d\end{pmatrix}*\begin{pmatrix}1&1\\ 0&1\end{pmatrix}=\begin{pmatrix}a-c&a-c+b-d\\ c&c+d\end{pmatrix}

So, we impose

(acac+bdcc+d)(1101)=(abcd)\displaystyle\begin{pmatrix}a-c&a-c+b-d\\ c&c+d\end{pmatrix}*\begin{pmatrix}1&1\\ 0&1\end{pmatrix}=\begin{pmatrix}a&b\\ c&d\end{pmatrix}
\displaystyle\iff (acac+bdcc+d)(1101)=(1101)(abcd)\displaystyle\begin{pmatrix}a-c&a-c+b-d\\ c&c+d\end{pmatrix}\begin{pmatrix}1&1\\ 0&1\end{pmatrix}=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}\begin{pmatrix}a&b\\ c&d\end{pmatrix}
\displaystyle\iff (ac2(ac)+bdc2c+d)=(a+cb+dcd) so c=0\displaystyle\begin{pmatrix}a-c&2(a-c)+b-d\\ c&2c+d\end{pmatrix}=\begin{pmatrix}a+c&b+d\\ c&d\end{pmatrix}\text{ so $c=0$ }
\displaystyle\iff (a2a+bd0d)=(ab+d0d) so a=d\displaystyle\begin{pmatrix}a&2a+b-d\\ 0&d\end{pmatrix}=\begin{pmatrix}a&b+d\\ 0&d\end{pmatrix}\text{ so $a=d$ }

and since the trace of the matrices equals 22, then a=1=da=1=d so

(abcd)=(1b01)=(acac+bdcc+d)\displaystyle\begin{pmatrix}a&b\\ c&d\end{pmatrix}=\begin{pmatrix}1&b\\ 0&1\end{pmatrix}=\begin{pmatrix}a-c&a-c+b-d\\ c&c+d\end{pmatrix}

so altogether we have less than three elements and therefore we do not have an isomorph with R3\operatorname{R}_{3}. This completes the proof.

Theorem 6.11.

The dihedral quandle of order 33, R3\operatorname{R}_{3}, is a subquandle of Mλ,λM_{\lambda,-\lambda}, for any λ\lambda\in\mathbb{C}^{*}.

Proof.

Since Mλ,λM_{\lambda,-\lambda} is isomorphic with M1,1M_{1,-1}, we will use the latter for our proof. Without loss of generality, we will prove that

{(abcd),(1001),(abcd)(1001)}={(abcd),(1001),(abcd)}M1,1,\left\{\,\begin{pmatrix}a&b\\ c&d\end{pmatrix},\begin{pmatrix}1&0\\ 0&-1\end{pmatrix},\begin{pmatrix}a&b\\ c&d\end{pmatrix}\ast\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\,\right\}=\left\{\,\begin{pmatrix}a&b\\ c&d\end{pmatrix},\begin{pmatrix}1&0\\ 0&-1\end{pmatrix},\begin{pmatrix}a&-b\\ -c&d\end{pmatrix}\,\right\}\,\subset\,M_{1,-1},

is the underlying set of R3\operatorname{R}_{3} in M1,1M_{1,-1}, find conditions for a,b,c,da,b,c,d and prove that the subquandle so generated is in fact R3\operatorname{R}_{3}. We thus resume the study of the system ()(\dagger\dagger) from the proof of Proposition 6.10, with λ1=1=λ2\lambda_{1}=1=-\lambda_{2}:

()\displaystyle(\dagger\dagger\dagger) {a2bc=aa+d=1ad=1bc+d2=d{da=1d2+dbc=0a2abc=0{da=1d±=1±1+4bc2a±=+1±1+4bc2\displaystyle\begin{cases}a^{2}-bc=a\\ -a+d=1\\ a-d=-1\\ -bc+d^{2}=-d\end{cases}\Longleftrightarrow\begin{cases}d-a=1\\ d^{2}+d-bc=0\\ a^{2}-a-bc=0\end{cases}\Longrightarrow\begin{cases}d-a=1\\ d_{\pm}={\displaystyle\frac{-1\pm\sqrt{1+4bc}}{2}}\\ a_{\pm}={\displaystyle\frac{+1\pm\sqrt{1+4bc}}{2}}\end{cases}\Longrightarrow
{d+a=1d+=1+1+4bc2a=+11+4bc2 (along with bc=3/4)0=a+d+(trace equation)1=ad+bc(determinant equation){a=1/2d+=1/2bc=3/4\displaystyle\Longrightarrow\begin{cases}d_{+}-a_{-}=1\\ d_{+}={\displaystyle\frac{-1+\sqrt{1+4bc}}{2}}\\ a_{-}={\displaystyle\frac{+1-\sqrt{1+4bc}}{2}}\quad\text{ (along with $bc=3/4$)}\\ 0=a_{-}+d_{+}\qquad(\text{trace equation})\\ -1=a_{-}d_{+}-bc\quad(\text{determinant equation})\end{cases}\Longrightarrow\begin{cases}a_{-}=-1/2\\ d_{+}=1/2\\ bc=3/4\end{cases}

so the three elements of the prospective isomorph with R3\operatorname{R}_{3} are:

(1001),(1/2bc1/2),(1/2bc1/2)withbc=3/4.\begin{pmatrix}1&0\\ 0&-1\end{pmatrix},\begin{pmatrix}-1/2&b\\ c&1/2\end{pmatrix},\begin{pmatrix}-1/2&-b\\ -c&1/2\end{pmatrix}\qquad\text{with}\qquad bc=3/4.

We will now check that whenever two distinct elements from these three are operated, the result is the other element.

(1001)(1/2±b±c1/2)=(1/2±b±c1/2)(1001)(1/2±b±c1/2)=\displaystyle\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\ast\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}=\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}=
=(1/2b±c1/2)(1/2±b±c1/2)=(1/4bcbcbc1/4)=(1/2bc1/2)\displaystyle=\begin{pmatrix}-1/2&\mp b\\ \pm c&-1/2\end{pmatrix}\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}=\begin{pmatrix}1/4-bc&\mp b\\ \mp c&bc-1/4\end{pmatrix}=\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}
(1/2±b±c1/2)(1001)=(1001)(1/2±b±c1/2)(1001)=\displaystyle\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}\ast\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}=
=(1/2±bc1/2)(1001)=(1/2bc1/2)\displaystyle=\begin{pmatrix}-1/2&\pm b\\ \mp c&-1/2\end{pmatrix}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}=\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}
(1/2±b±c1/2)(1/2bc1/2)=(1/2bc1/2)(1/2±b±c1/2)(1/2bc1/2)=\displaystyle\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}\ast\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}=\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}\begin{pmatrix}-1/2&\pm b\\ \pm c&1/2\end{pmatrix}\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}=
=(1/4bcb±c1/4bc)(1/2bc1/2)=(1/2b±c1/2)(1/2bc1/2)=\displaystyle=\begin{pmatrix}1/4-bc&\mp b\\ \pm c&1/4-bc\end{pmatrix}\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}=\begin{pmatrix}-1/2&\mp b\\ \pm c&-1/2\end{pmatrix}\begin{pmatrix}-1/2&\mp b\\ \mp c&1/2\end{pmatrix}=
=(1/4+bc00bc1/4)=(1001)\displaystyle=\begin{pmatrix}1/4+bc&0\\ 0&-bc-1/4\end{pmatrix}=\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}

This completes the proof. ∎

7. Future Work

In this section we collect some questions for future work.

Conjecture 7.1.

Let QQ be an infinite type quandle. Then 𝒬n(Q)\mathcal{Q}_{n}(Q) is not isomorphic to 𝒬m(Q)\mathcal{Q}_{m}(Q), for nmn\neq m and n,m+n,m\in\mathbb{Z}^{+}.

Question 7.2.

Let ω1\omega_{1} and ω2\omega_{2} be distinct nn-th primitive roots of unity. Is there any relation between the quandles M1,ω1M_{1,\omega_{1}} and M1,ω2M_{1,\omega_{2}}?

Question 7.3.

Does there exist a connected quandle QQ such that 𝒬n(Q)\mathcal{Q}_{n}(Q) is connected for all n+n\in\mathbb{Z}^{+}?

Question 7.4.

Classify infinite type Mλ1,λ2M_{\lambda_{1},\lambda_{2}} quandles, where λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{C}^{*}.

Question 7.5.

Can a dihedral quandle R2k+1\operatorname{R}_{2k+1}, where k+k\in\mathbb{Z}^{+}, be subquandle of Mλ1,λ2\operatorname{M}_{\lambda_{1},\lambda_{2}} or Mλ\operatorname{M}_{\lambda}, where λ,λ1,λ2\lambda,\lambda_{1},\lambda_{2}\in\mathbb{C}^{*} and λ1λ2\lambda_{1}\neq\lambda_{2}?

Question 7.6.

What other well known connected quandles are subquandles of Conj(GL(2,)\operatorname{Conj}(\operatorname{GL}(2,\mathbb{C})?

Acknowledgement.

Pedro Lopes is supported by CAMGSD via project UIDB/04459/2020, financed by national funds FCT/MCTES (PIDDAC). Manpreet Singh is supported by UIDP/04459/2020 Post-Doctoral Research Fellowship at CAMGSD, financed by national funds FCT/MCTES (PIDDAC).

References

  • [1] N. Andruskiewitsch and M. Graña. From racks to pointed Hopf algebras. Adv. Math., 178(2):177–243, 2003.
  • [2] V. Bardakov, T. Nasybullov, and M. Singh. Automorphism groups of quandles and related groups. Monatsh. Math., 189(1):1–21, 2019.
  • [3] V. Bardakov and M. Singh. Quandle cohomology, extensions and automorphisms. J. Algebra, 585:558–591, 2021.
  • [4] V. G. Bardakov, P. Dey, and M. Singh. Automorphism groups of quandles arising from groups. Monatsh. Math., 184(4):519–530, 2017.
  • [5] V. G. Bardakov and D. A. Fedoseev. Multiplication of quandle structures. arXiv e-prints, page arXiv:2204.12571, Apr. 2022.
  • [6] V. G. Bardakov, I. B. S. Passi, and M. Singh. Quandle rings. J. Algebra Appl., 18(8):1950157, 23, 2019.
  • [7] V. G. Bardakov, M. Singh, and M. Singh. Free quandles and knot quandles are residually finite. Proc. Amer. Math. Soc., 147(8):3621–3633, 2019.
  • [8] V. G. Bardakov, M. Singh, and M. Singh. Link quandles are residually finite. Monatsh. Math., 191(4):679–690, 2020.
  • [9] G. Bianco and M. Bonatto. On connected quandles of prime power order. Beitr. Algebra Geom., 62(3):555–586, 2021.
  • [10] M. Bonatto and D. Stanovský. Commutator theory for racks and quandles. J. Math. Soc. Japan, 73(1):41–75, 2021.
  • [11] J. S. Carter, M. Elhamdadi, M. A. Nikiforou, and M. Saito. Extensions of quandles and cocycle knot invariants. J. Knot Theory Ramifications, 12(6):725–738, 2003.
  • [12] J. S. Carter, D. Jelsovsky, S. Kamada, L. Langford, and M. Saito. State-sum invariants of knotted curves and surfaces from quandle cohomology. Electron. Res. Announc. Amer. Math. Soc., 5:146–156, 1999.
  • [13] J. S. Carter, D. Jelsovsky, S. Kamada, L. Langford, and M. Saito. Quandle cohomology and state-sum invariants of knotted curves and surfaces. Trans. Amer. Math. Soc., 355(10):3947–3989, 2003.
  • [14] A. Choudhry. Extraction of nnth roots of 2×22\times 2 matrices. Linear Algebra Appl., 387:183–192, 2004.
  • [15] W. E. Clark and M. Saito. Algebraic properties of quandle extensions and values of cocycle knot invariants. J. Knot Theory Ramifications, 25(14):1650080, 17, 2016.
  • [16] M. Eisermann. Homological characterization of the unknot. J. Pure Appl. Algebra, 177(2):131–157, 2003.
  • [17] M. Eisermann. Yang-Baxter deformations of quandles and racks. Algebr. Geom. Topol., 5:537–562, 2005.
  • [18] M. Eisermann. Quandle coverings and their Galois correspondence. Fund. Math., 225(1):103–168, 2014.
  • [19] M. Elhamdadi, N. Fernando, and B. Tsvelikhovskiy. Ring theoretic aspects of quandles. J. Algebra, 526:166–187, 2019.
  • [20] M. Elhamdadi and S. Nelson. Quandles—an introduction to the algebra of knots, volume 74 of Student Mathematical Library. American Mathematical Society, Providence, RI, 2015.
  • [21] G. Ellis. HAP, homological algebra programming, Version 1.47. https://gap-packages.github.io/hap, Aug 2022. Refereed GAP package.
  • [22] P. Etingof, T. Schedler, and A. Soloviev. Set-theoretical solutions to the quantum Yang-Baxter equation. Duke Math. J., 100(2):169–209, 1999.
  • [23] D. Joyce. A classifying invariant of knots, the knot quandle. J. Pure Appl. Algebra, 23(1):37–65, 1982.
  • [24] D. E. Joyce. An Algebraic approach to symmetry with applications to knot theory. ProQuest LLC, Ann Arbor, MI, 1979. Thesis (Ph.D.), University of Pennsylvania.
  • [25] S. Kamada. Surface-knots in 4-space. Springer Monographs in Mathematics. Springer, Singapore, 2017. An introduction.
  • [26] A. Lages and P. Lopes. Quandles of cyclic type with several fixed points. Electron. J. Combin., 26(3):Paper No. 3.42, 28, 2019.
  • [27] A. Lages, P. Lopes, and P. Vojtěchovský. A sufficient condition for a quandle to be Latin. J. Combin. Des., 30(4):251–259, 2022.
  • [28] S. Lang. Algebra, volume 211 of Graduate Texts in Mathematics. Springer-Verlag, New York, third edition, 2002.
  • [29] O. Loos. Reflexion spaces and homogeneous symmetric spaces. Bull. Amer. Math. Soc., 73:250–253, 1967.
  • [30] P. Lopes and D. Roseman. On finite racks and quandles. Comm. Algebra, 34(1):371–406, 2006.
  • [31] S. V. Matveev. Distributive groupoids in knot theory. Mat. Sb. (N.S.), 119(161)(1):78–88, 160, 1982.
  • [32] S. Nelson and C.-Y. Wong. On the orbit decomposition of finite quandles. J. Knot Theory Ramifications, 15(6):761–772, 2006.
  • [33] T. Nosaka. Quandles and topological pairs. SpringerBriefs in Mathematics. Springer, Singapore, 2017. Symmetry, knots, and cohomology.
  • [34] H. Raundal, M. Singh, and M. Singh. Orderability of link quandles. Proc. Edinb. Math. Soc. (2), 64(3):620–649, 2021.
  • [35] S. K. Winker. Quandles, knot invariants, and the n-fold banched cover. ProQuest LLC, Ann Arbor, MI, 1984. Thesis (Ph.D.)–University of Illinois at Chicago.