Masters PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 82

KNOTS, TANGLES AND BRAID ACTIONS

by
LIAM THOMAS WATSON
B.Sc. The University of British Columbia
A THESIS SUBMITTED IN PARTIAL FULFILLMENT OF
THE REQUIREMENTS FOR THE DEGREE OF

MASTER OF SCIENCE
in
THE FACULTY OF GRADUATE STUDIES
Department of Mathematics
We accept this thesis as conforming
to the required standard

.....................................

.....................................

THE UNIVERSITY OF BRITISH COLUMBIA


October 2004
c Liam Thomas Watson, 2004

In presenting this thesis in partial fulfillment of the requirements for an advanced
degree at the University of British Columbia, I agree that the Library shall make
it freely available for reference and study. I further agree that permission for ex-
tensive copying of this thesis for scholarly purposes may be granted by the head of
my department or by his or her representatives. It is understood that copying or
publication of this thesis for financial gain shall not be allowed without my written
permission.

(Signature)

Department of Mathematics
The University of British Columbia
Vancouver, Canada

Date
Abstract
Recent work of Eliahou, Kauffmann and Thistlethwaite suggests the use of
braid actions to alter a link diagram without changing the Jones polynomial.
This technique produces non-trivial links (of two or more components) having
the same Jones polynomial as the unlink. In this paper, examples of distinct
knots that can not be distinguished by the Jones polynomial are constructed
by way of braid actions. Moreover, it is shown in general that pairs of knots
obtained in this way are not Conway mutants, hence this technique provides
new perspective on the Jones polynomial, with a view to an important (and
unanswered) question: Does the Jones polynomial detect the unknot?

ii
Table of Contents
Abstract ii

Table of Contents iii

List of Figures v

Acknowledgement vii

Chapter 1. Introduction 1

Chapter 2. Knots, Links and Braids 3


2.1 Knots and Links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 Braids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

Chapter 3. Polynomials 8
3.1 The Jones Polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2 The Alexander Polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
3.3 The HOMFLY Polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Chapter 4. Tangles and Linear Maps 21


4.1 Conway Tangles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
4.2 Conway Mutation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4.3 The Skein Module . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4 Linear Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.5 Braid Actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

Chapter 5. Kanenobu Knots 37


5.1 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.2 Basic Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.3 Main Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.5 Generalisation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.6 More Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.7 Observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

iii
Table of Contents iv

Chapter 6. Thistlethwaite Links 62


6.1 Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.2 Some 2-component examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.3 A 3-component example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.4 Closing Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

Bibliography 72

iv
List of Figures
2.1 Diagrams of the Trefoil Knot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2 The Hopf Link . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.3 The braid generator σi and its inverse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.4 The link β̄ formed from the closure of β. . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

4.1 Some diagrams of Conway tangles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21


4.2 The generator ei . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.3 The tangle T β ∈ S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

5.1 The Kanenobu knot K(T, U ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37


5.2 Distinct knots that are not Conway mutants . . . . . . . . . . . . . . . . . . . . . . 43
5.3 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.5 Another diagram of the Kanenobu knot K(T, U ) . . . . . . . . . . . . . . . . . . 50
5.6 The link |β| . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.7 The closure of a Kanenobu braid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.8 The braid σ13 σ2−1 σ1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.9 The Kanenobu
3 −1 braid (σ13 σ2−1 σ1 )(σ −3 −1
5 σ4 σ5 ) . . . . . . . . . . . . . . . . . . . . . . . 53
5.10 The knot (σ1 σ2 σ1 )(σ5 σ4 σ5 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
−3 −1
54
5.11 The knot K(0, 0) ∼ 52 # 52? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.12 The knot K(σ2 , σ2−1 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.13 The braid σ12 σ2−3 σ1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.14 The Kanenobu
2 −3 braid (σ12 σ2−3 σ1 )(σ −2 3 −1
5 σ4 σ5 ) . . . . . . . . . . . . . . . . . . . . . . . 56
5.15 The knot (σ1 σ2 σ1 )(σ5 σ4 σ5 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
−2 3 −1
57
5.16 The knot K(0, 0) ∼ 61 # 61? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.17 The knot K(σ2 , σ2−1 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.18 The 2-bridge link obtained from β ∈ B 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
5.19 The square knot 31 # 3?1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.20 The knot (σ1 σ2−1 σ1 )(σ5−1 σ4 σ5−1 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

6.1 The Thislethwaite link H(T, U ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63


6.2 The link H(0, 0) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
6.3 The braid ω ∈ B3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
−1
6.4 The link H(T ω , U ω ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
6.5 The result of ω 2 acting on H(T, U ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.6 The result of ω 4 acting on H(T, U ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

v
List of Figures vi

6.7 A non-trivial, 2-component link . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69


6.8 A non-trivial, 2-component link . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.9 A non-trivial, 3-component link . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

vi
Acknowledgement
It’s a strange process, writing down the details to something you’ve been work-
ing on for some time. And while it is a particularly solitary experience, the
final product is not accomplished on you’re own. To this end, there are many
people that should be recognised as part of this thesis.
First, I would like to thank my supervisor Dale Rolfsen for ongoing patience,
guidance and instruction. It has been a privilege and a pleasure to learn from
Dale throughout my undergraduate and masters degrees, and I am grateful for
his generous support, both intellectual and financial. I have learned so much.
Of course, I am indebted to all of my teachers as this work comprises much of
what I have learned so far. However, I would like to single out Bill Casselman,
as the figures required for this thesis could not have been produced without
his guidance.
To my parents, Peter and Katherine Watson, family and friends, I am so
fortunate to have a support network that allows me to pursue mathematics.
In particular, I would like to thank Erin Despard for continued encouragement,
support and perspective, everyday.

vii
Chapter 1
Introduction

The study of knots begins with a straight-forward question: Can we distinguish


between two closed loops, embedded in three dimensions? This leads naturally
to a more general question of links, that is, the ability to distinguish between
two systems of embedded closed loops. Early work by Alexander [1, 2], Artin [3,
4], Markov [24] and Reidemeister [29] made inroads into the subject, developing
the first knot and link invariants, as well as the combinatorial and algebraic
languages with which to approach the subject. The subtle relationship between
the combinatorial and algebraic descriptions continue to set the stage for the
study of knots and links.
With the discovery of the Jones polynomial [15] in 1985, along with a two
variable generalization [12] shortly thereafter, the study of knots was given
new focus. These new polynomial invariants could be viewed as combinatorial
objects, derived directly from a diagram of the knot, or as algebraic objects,
resulting from representations of the braid group. However, although the new
polynomials were able to distinguish between knots that had previously caused
difficulties, they led to new questions in the study of knots that have yet to be
answered.
In particular, we are led to the phenomenon of distinct knots having the same
Jones polynomial. There are many examples of families of knots that share
common Jones polynomials. Such examples have given way to a range of tools
to describe this occurrence [30, 31]. In particular, it is unknown if there is a
non-trivial knot that has trivial Jones polynomial. This question motivates the
understanding of knots that cannot be distinguished by the Jones polynomial,
as well as the development of examples of such along with tools to explain
the phenomenon. The prototypical method for producing two knots having

1
Chapter 1. Introduction 2

the same Jones polynomial is known as Conway mutation. However, it is well


known that this method will not alter an unknot to produce a non-trivial knot.
Recent work of Eliahou, Kauffman and Thistlethwaite [9] suggests the use
of braid group actions in the study links having the same Jones polynomial.
Revisiting earlier work of Kanenobu [18], new families of knots are described in
this work. Once again, there is a subtle relationship between the combinatorics
and the algebra associated with such examples. As a result, the study of knots
obtained through braid actions can be restated in terms of fixed points of an
associated group action.
The study of this braid action certainly merits attention, as the work of Eli-
ahou, Kauffman and Thistlethwaite [9] explores Thistletwaite’s discovery [33]
of links having the trivial Jones polynomial, settling the question for links
having more than one component. As a result, only the case of knots is left
unanswered as of this writing.
This thesis is a study of families of knots sharing a common Jones polynomial.
In chapter 1 the classical definitions and results of knot theory are briefly
reviewed, developing the necessary background for the definitions of the Jones,
Alexander and HOMFLY polynomials in chapter 2. Then, in chapter 3, the
linear theory of tangles (due to Conway [8]) is carefully reviewed. Making use
of this linear structure, we define a new form of mutation by way of an action
of the braid group on the set of tangles.
The main results of this work are contained in chapter 5. We produce exam-
ples of distinct knots that share a common Jones polynomial, and develop a
generalization of knots due to Kanenobu [18]. Moreover, it is shown (theorem
5.3) that knots constructed in this way are not related by Conway mutation.
We conclude by restating the results of Eliahou, Kauffman and Thistlethwaite
[9] in light of this action of the braid group, giving examples of non-trivial
links having trivial Jones polynomial in chapter 6.
Chapter 2
Knots, Links and Braids

2.1 Knots and Links


A knot K is a smooth or piecewise linear embedding of a closed curve in a
3-dimensional manifold. Usually, the manifold of choice is either R 3 or S3 , so
that the knot K may be denoted

S1 ,→ R3 ⊂ S3 .

While it is important to remember that we are dealing with curves in 3-


dimensions, it is difficult to work with such objects. As a result, we deal
primarily with a projection of a knot to a 2-dimensional plane called a knot
diagram. In this way a knot may be represented on the page as in figure 2.1.

Figure 2.1: Diagrams of the Trefoil Knot

In such a diagram the indicates that the one section of the knot (the
broken line) has passed behind another (the solid line) to form a crossing.
In general there will not be any distinction made between the knot K and a
diagram representing it. That is, we allow a given diagram to represent a knot

3
Chapter 2. Knots, Links and Braids 4

and denote the diagram by K also. It should be pointed out, however, that
there are many diagrams for any given knot. Indeed, K and K 0 are equivalent
knots (denoted K ∼ K 0 ) if they are related by isotopy in S3 . Therefore, the
diagrams for K and K 0 may be very different.
An n-component link is a collection of knots. That is, a link is a disjoint union
of embedded circles
an
S1i ,→ R3 ⊂ S3
i=1

where each
S1i ,→ R3 ⊂ S3
is a knot. Of course, a 1-component link is simply a knot, and a non-trivial
link can have individual components that are unknotted.

Figure 2.2: The Hopf Link

To study links by way of diagrams, it is crucial to be able to alter a link


diagram in a way that reflects changes in the link resulting from isotopy in S 3 .
To this end, we introduce the Reidemeister Moves defined in [28, 29].

∼ ∼ (R1 )

∼ ∼ (R2 )

∼ (R3 )

In each of the three moves, it is understood that the diagram is unchanged


outside a small disk inside which the move occurs.

Theorem (Reidemeister). Two link diagrams represent the same link iff the
diagrams are related by planar isotopy, and the Reidemeister moves.
Chapter 2. Knots, Links and Braids 5

Assigning an orientation to each component of a link L gives rise to the oriented


~
link L.

Definition 2.1. Let C be the set of crossings of a diagram L. The writhe of


~ is obtained taking a sum over all crossings C
an orientation L
X
~ =
w(L) w(c)
c∈C

where w(c) = ±1 is determined by a right hand rule as in


   
w =1 and w = −1.

While writhe is not a link invariant, it does give rise to the following definition.

Definition 2.2. For components L1 and L2 of L let C 0 ⊂ C be the set of


crossings of L formed by the interaction of L 1 and L2 . The linking number of
L1 and L2 is given by
X w(c)
lk(L1 , L2 ) = .
2
0 c∈C

The linking number is a link invariant. Note that, for the Hopf link of figure
2.2, there are two distinct orientations. One orientation has linking number 1,
the other linking number −1 and hence there are two distinct oriented Hopf
links.

2.2 Braids
There are many equivalent definitions of braids (see [6], [10], [27]). In this
setting it is natural to start from a geometric point of view.
Let E ⊂ R3 denote the yz-plane and let E 0 denote its image shifted by 1 in
the x direction. Consider the the collection of points

P = {1, . . . , n} = {(0, 0, 1), . . . , (0, 0, n)} ⊂ E

and denote by
P 0 = {(1, 0, 1), . . . , (1, 0, n)}
the image of P in E 0 .
Chapter 2. Knots, Links and Braids 6

Definition 2.3. A (n-strand) braid is a collection of embedded arcs (or strands)

αi : [0, 1] ,→ [0, 1] × E ⊂ R3

such that

(a) αi (0) = i ∈ P

(b) αi (1) ∈ P 0

(c) αi ∩ αj = ∅ as embedded arcs for i 6= j.

(d) αi is monotone increasing in the x direction.

As with knots, it will be convenient to consider the diagram of a braid by


projecting to the xy-plane. Also, we may consider equivalence of braids via
isotopy (through braids), although we will confuse the notion of a braid and
its equivalence class.
In [3, 4] Artin showed that there is a well defined group structure for braids.
The identity braid is represented by setting each arc to a constant map α i (x) =
(x, 0, i) so that each strand is a straight line. Multiplication of braids is defined
by concatenation, so that inverses are constructed by reflecting in the xz-plane.
The n-strand braid group has presentation
* +
σσ =σ σ |i − j| > 1
i j j i
Bn = σ1 . . . σn−1
σi σj σi = σj σi σj |i − j| = 1

where the generators correspond to a crossing formed between the i and i+1
strand as in figure 2.3.
Just as group elements are formed by words in the generators, a braid diagram
for a given element can be constructed by concatenation of braids of the form
shown in figure 2.3.
If E and E 0 of an n-strand braid are identified so that P = P 0 pointwise, the
result is a collection of embeddings of S 1 in R3 and we obtain a link.
Given any braid β we can form a link β̄ by taking the closure in this way. It is
a theorem of Alexander [1] that every link arises as the closure of some braid.
Given a link diagram L, it is always possible to construct a braid β such that
β̄ = L. Two such constructions (there are many) are due to Morton [26] and
Vogel [34].
Chapter 2. Knots, Links and Braids 7




 

 
 
 
 
 


 

Figure 2.3: The braid generator σi and its inverse

 

 



Figure 2.4: The link β̄ formed from the closure of β.

Now it should be noted that the group operation

σi σi−1 = 1 = σi−1 σi

corresponds exactly to the Reidemeister move R 2 , while the group relation

σi σj σi = σ j σi σj

corresponds to the Reidemeister move R 3 . This suggests the possibility of


studying equivalence of links through braid representatives. To this end we
define the Markov moves. Suppose β ∈ B n and write β = (β, n). Then

(β1 β2 , n) ∼M (β2 β1 , n) (M1 )

(β, n) ∼M (βσn±1 , n+1) (M2 )

where ∼M denotes Markov equivalence. The following theorem, due to Markov


[24], is proved in detail in [6].

Theorem (Markov). Two links β¯1 and β¯2 are equivalent iff β1 ∼M β2 .
Chapter 3
Polynomials

3.1 The Jones Polynomial


Define the Kauffman bracket hLi of a link diagram L recursively by the axioms
 
=1 (3.1)
     
−1
=a +a (3.2)
   
Lt =δ L (3.3)

where a is a formal variable and


δ = −a−2 − a2
so that hLi is an element of the (Laurent) polynomial ring Z[a, a −1 ]. In some
cases a is specified as a non-zero complex number, in which case hLi ∈ C.
The Kauffman bracket is invariant under the Reidemeister moves R 2 and R3 .
To get invariance under R1 , we recall definition 2.1 for the writhe of an orien-
~ of the diagram L. The writhe of a crossing is ±1 and is determined
tation L
by a right hand rule. That is
   
w =1 and w = −1

~ ∈ Z. Now
so that w(L)
~
−a−3w(L) hLi ∈ Z[a, a−1 ]
is invariant under R1 and gives rise to an invariant of oriented links.

8
Chapter 3. Polynomials 9

Definition 3.1. The Jones Polynomial [15, 19] is given by



~
VL~ (t) = −a−3w(L) hLi − 1
a=t 4

where t is a commuting variable.

Note that it will often be convenient to work with t = a −4 , and the polynomial
obtained through this substitution will be referred to as the Jones Polynomial
also.
As we shall see, there are many examples of distinct links having the same
Jones Polynomial. However, the following is still unknown:

~ (t) = 1 imply that K ∼


Question 3.2. For a knot K, does VK ?

3.2 The Alexander Polynomial


For any knot K, let F be an orientable surface such that ∂F = K. Such a
surface always exists [32], and is called a Seifert surface for the link K. The
homology of such a surface is given by
M
H1 (F, Z) = Z
2g

where g is the genus of the surface F . Let {a i } be a set of generators for


H1 (F, Z) where i ∈ {1, . . . , 2g} .
Let
D 2 = {z ∈ C : |z| < 1}
and consider a tubular neighborhood N (K) ∼ = K × D 2 of the link K. That is,
an embedding
S1 × D 2 ,→ S3
such that K is the restriction to S1 × {0}.
Now consider the surface F in the complement X = S 3 r N (K). Here F is
being confused with its image in the compliment X, by abuse. For a regular
neighborhood
F × [1, 1] ⊂ S3
Chapter 3. Polynomials 10

there are natural inclusions


i± : F ,→ F × {±1}
where F = F × {0} is the Seifert surface in X. Therefore a cycle x ∈ H 1 (F, Z)
gives rise to a cycle x± = i±
? (x) ∈ H1 (X, Z).

Definition 3.3. The Seifert Form is the bilinear form


v : H1 (F, Z) × H1 (F, Z) → Z
(x, y) 7→ lk(x, y + )
and it is represented by the Seifert Matrix
 
V = lk(ai , a+
j )

where y + = i+
? (y).

e the infinite cyclic cover [25, 32] of the knot com-


The aim is to construct X,
3
plement X = S r N (K). To do this, start with a countable collection {X i }i∈Z
of
Xi = X r (F × (−, ))
for some small  ∈ (0, 1). The boundary of this space contains two identical
copies of F denoted by
F ± = F × {±},
and the infinite cyclic cover of X is defined
S ,
e= i∈Z X i
X
Fi+ ∼ Fi+1

by identifying Fi+ ⊂ ∂Xi with Fi+1



⊂ ∂Xi+1 for each i ∈ Z.
The space obtained corresponds to the short exact sequence

1 // π X
1
e // π1 X // H1 (X, Z) // 0

 
α ~
// lk α, K

so that the infinite cyclic group H1 (X, Z) = hti gives the covering translations
of Xe & X. Now H1 (X̃, Z), although typically not finitely generated as an
abelian group, is finitely generated as a Z[t, t −1 ]-module by the {ai }. The
variable t corresponds to the hti-action taking X i to Xi+1 .
Chapter 3. Polynomials 11

Definition 3.4. H1 (X̃, Z) is called the Alexander module and has module
presentation V − tV > , where V is the Seifert matrix. This is a knot invariant.

This gives rise to another polynomial invariant due to Alexander [2].

Definition 3.5. The Laurent polynomial


.
∆K (t) = det(V − tV > ) ∈ Z[t, t−1 ]

is an invariant of the knot K called the Alexander polynomial. It is defined


.
up to multiplication by a unit ±t±n (indicated by =).

This knot invariant is of particular interest due to this topological construction.

Question 3.6. Is there a similar topological interpretation for the Jones poly-
nomial?

On the other hand, it is easy to generate knots K such that ∆ K (t) = 1 (see
for example, [32]).
.
Theorem 3.7. For any knot K, ∆K (t) = ∆K (t−1 ).

proof. Given the n × n Seifert matrix V,


.
∆K = det(V − tV > )
= det(V > − tV )
= (−t)n det(V − t−1 V > )
.
= ∆K (t−1 ).

Theorem 3.8. For any knot K, ∆K (1) = ±1.

proof. Setting t = 1 and using the standard (symplectic) basis for H 1 (F, Z)
Chapter 3. Polynomials 12

gives
 
V − V > = lk(ai , a+
j ) − lk(a j , a +
i )
 
= lk(ai , a+
j ) − lk(a −
j , a i )
 
= lk(ai , a+
j ) − lk(a i , a −
j )
 
0 1
−1 0 
 
 0 1 
 
 −1 0 
= 
 .. 
 . 
 
 0 1
−1 0
therefore
∆K (1) = ± det(V − V > ) = ±1.

Corollary 3.9. For any knot K


.
∆K (t) = c0 + c1 (t−1 + t1 ) + c2 (t−1 + t1 )2 + · · ·
where ci ∈ Z.
.
proof. The symmetry given by ∆K (t) = ∆K (t−1 ) gives rise to the form
m
. X k
∆K (t) = bk t
k=0

where cm−r = ±cr , and the same choice of sign is made for each r. Now m
must be even, since m odd gives rise to ∆ K (1) even, contradicting theorem
3.8. Further, if cm−r = −cr then cm/2 = 0 and
m
X
∆K (1) = bk = 0,
k=0

again contradicting theorem 3.8. Therefore c m−r = cr , and


.
∆K (t) = c0 + c1 (t−1 + t1 ) + c2 (t−1 + t1 )2 + · · ·
as required.
Chapter 3. Polynomials 13

We will see the form given in corollary 3.9 in the next section. Together with
the normalization ∆K (1) = 1, it is sometimes referred to as the Alexander-
Conway polynomial as it has a recursive definition, originally noticed by Alexan-
der [2] and later exploited by Conway [8].
It should be noted that there are generalizations of this construction to invari-
ants of oriented links that have been omitted. Nevertheless, we shall see that
the recursive definition of ∆K (t) is defined for all oriented links.

3.3 The HOMFLY Polynomial


A two variable polynomial [12, 16] that restricts to each of the polynomials
introduced may be defined, albeit by very different means.
The n-strand braid group Bn generates a group algebra Hn over Z[q, q −1 ] which
has relations
(i) σ i σj = σ j σi for |i − j| > 1
(ii) σ i σj σi = σ j σi σj for |i − j| = 1
(iii) σi2 = (q − 1)σi + q ∀ i ∈ {1, . . . , n−1}
called the Hecke algebra. By allowing q to take values in C, H n can be seen
as a quotient of the group algebra CB n . Just as
{1} < B2 < B3 < B4 < · · ·
we have that
Z[q, q −1 ] ⊂ H2 ⊂ H3 ⊂ H4 ⊂ · · · .
Note that for q = 1, the relation (iii) reduces to σ i2 = 1 and we obtain the
relations for the symmetric group S n .
Definition 3.10. Sets of positive permutation braids may be defined recur-
sively via
Σ0 = {1}
Σi = {1} ∪ σi Σi−1 for i > 0.
A monomial m ∈ Hn is called normal if it has the form
m = m1 m2 . . . mn−1
where mi ∈ Σi .
Chapter 3. Polynomials 14

The normal monomials form a basis for H n , and it follows that

dimZ[q,q−1 ] (Hn ) = n!.

Moreover, this basis allows us to present any element of H n+1 in the form

x1 + x 2 σn x3

for xi ∈ Hn . The relation (i) implies that

xσn = σn x

whenever x ∈ Hn−1 , giving rise to the decomposition



Hn+1 ∼= Hn ⊕ Hn ⊗Hn−1 Hn .

Now we define a linear trace function

tr : Hn −→ Z[q ±1 , z]
σi 7−→ z

that is normalized so that tr(1) = 1.


Theorem 3.11. tr(x1 x2 ) = tr(x2 x1 ) for xi ∈ Hn .

proof. By linearity, it suffices to show that tr(m 1 m2 ) = tr(m2 m1 ) for nor-


mal monomials mi ∈ Hn . Since the theorem is clearly true for the normal
monomials of H2 , we proceed by induction.
Suppose first that m1 = m01 σn m001 where m01 , m001 ∈ Hn and m2 ∈ Hn (that is,
m2 contains no σn ). Then

tr(m1 m2 ) = tr(m01 σn m001 m2 )


= z tr(m01 m001 m2 )
= z tr(m2 m01 m001 ) by induction
= tr(m2 m01 σn m001 )
= tr(m2 m1 ).

Now more generally write

m1 = m01 σn m001 and m2 = m02 σn m002

where m0i , m00i ∈ Hn . In this case we will make use of the following:
Chapter 3. Polynomials 15

(1) tr(µ1 σn µ2 σn ) = tr(σn µ1 σn µ2 )


(2) tr(µ1 σn µ2 σn µ3 ) = tr(µ3 µ1 σn µ2 σn )
where µi ∈ Hn are in normal form so that µi = µ0i σn−1 µ00i with µ0i , µ00i ∈ Hn−1 .
(1)
tr(µ1 σn µ2 σn ) = tr(µ1 σn µ02 σn−1 µ002 σn )
= tr(µ1 µ02 σn σn−1 σn µ002 ) using (i)
= tr(µ1 µ02 σn−1 σn σn−1 µ002 ) using (ii)
2
= z tr(µ1 µ02 σn−1 µ002 )
= z tr(µ1 µ02 [(q − 1)σn−1 + q]µ002 ) using (iii)
= z(q − 1) tr(µ1 µ02 σn−1 µ002 ) + zq tr(µ1 µ02 µ002 )
= z(q − 1) tr(µ1 µ2 ) + zq tr(µ01 σn−1 µ001 µ02 µ002 )
= z(q − 1) tr(µ1 µ2 ) + zq tr(µ01 µ001 µ02 σn−1 µ002 ) induction
= z(q − 1) tr(µ1 µ2 ) + zq tr(µ01 µ001 µ2 )
= z tr(µ0i [(q − 1)σn−1 + q]µ001 µ2 )
2
= z tr(µ01 σn−1 µ001 µ2 ) using (iii)
= tr(µ01 σn−1 σn σn−1 µ001 µ2 )
= tr(µ01 σn σn−1 σn µ001 µ2 ) using (ii)
= tr(σn µ01 σn−1 µ001 σn µ2 ) using (i)
= tr(σn µ1 σn µ2 )

(2)
tr(µ1 σn µ2 σn µ3 )
= tr(µ1 µ2 σn µ3 σn ) applying (1)
= tr(µ1 µ2 σn µ03 σn−1 µ003 σn )
2
= z tr(µ1 µ2 µ03 σn−1 µ003 ) as above
= z(q − 1) tr(µ1 µ2 µ03 σn−1 µ003 ) + zq tr(µ1 µ2 µ03 µ003 ) using (iii)
= z(q − 1) tr(µ03 σn−1 µ003 µ1 µ2 ) + zq tr(µ03 µ003 µ1 µ2 ) induction
2
= z tr(µ03 σn−1 µ003 µ1 µ2 ) using (iii)
= tr(σn µ03 σn−1 µ003 σn µ1 µ2 ) as above
= tr(σn µ3 σn µ1 µ2 )
= tr(µ3 µ1 σn µ2 σn ) applying (1)
Chapter 3. Polynomials 16

Now the proof is complete, since

tr(m1 m2 ) = tr(m01 σn m001 m02 σn m002 )


= tr(m002 m01 σn m001 m02 σn ) by (2)
= tr(σn m002 m01 σn m001 m02 ) by (1)
= tr(m02 σn m002 m01 σn m001 ) by (2)
= tr(m2 m1 ).

Now since elements of Hn+1 are of the form x1 + x2 σn x3 where xi ∈ Hn , the


trace function may be extended from H n to Hn+1 by

tr(x1 + x2 σn x3 ) = tr(x1 ) + z tr(x2 x3 ).

The aim is to use the trace function to define a link invariant. In particular
we would like to make use of this trace on braids in the composite

Bn −→ Hn −→ Z[q ±1 , z].
w
To do this, we introduce a change of variables λ = qz where

1−q λq(1 − q)
z=− and w=−
1 − λq 1 − λq
so that
1−q+z
λ= .
zq
Definition 3.12. The HOMFLY polynomial is given by
 n−1  
1 − λq √ e
Xβ̄ (q, λ) = − √ λ tr(β)
λ(1 − q)

where β ∈ Bn is a monomial in Hn and e = e(β) is the exponent sum of β


(equivalently, the abelianization B n → Z).

Note that the closure of the identity braid in B n gives the n component unlink

···
| {z }
n
Chapter 3. Polynomials 17

and the HOMFLY polynomial for this link is given by


 n−1
1 − λq
−√ . (3.4)
λ(1 − q)
 √ ±1 
±1
Theorem 3.13. Let β̄ = L then XL (q, λ) ∈ Z q , λ is a link invari-
ant.

proof. By Markov’s theorem, we need only check that X L (q, λ) is invariant


under M1 and M2 . The fact that tr(β1 β2 ) = tr(β2 β1 ) from Theorem 3.11 gives
invariance under M1 , so it remains to check invariance under M 2 . Suppose
then that β ∈ Bn . With the above substitution we have
1−q
tr(σn ) = −
1 − λq
so that
 n  
1 − λq √ e+1
Xβσn (q, λ) = −√ λ tr(βσi )
λ(1 − q)
  
√ 1 − λq 1−q
= λ −√ − Xβ̄ (q, λ)
λ(1 − q) 1 − λq
= Xβ̄ (q, λ).

Further, from (iii) we can derive

σi2 = (q − 1)σi + q
σi = (q − 1) + qσi−1
qσi−1 = σi + 1 − q
σi−1 = q −1 σi + q −1 − 1
Chapter 3. Polynomials 18

hence

tr(σi−1 ) = tr(q −1 σi + q −1 − 1)
= q −1 tr(σi ) + q −1 − 1
 
−1 1−q
=q − + q −1 − 1
1 − λq
 
−1 1−q
=q − +1 −1
1 − λq
 
−1 −1 + q + 1 − λq
=q −1
1 − λq
1 − λ − 1 + λq
=
1 − λq
1−q
= −λ .
1 − λq
Thus
 n  
1 − λq √ e−1
X −1 (q, λ) = −√ λ tr(βσ1−1 )
βσn λ(1 − q)
  
1 1 − λq 1−q
=√ −√ −λ Xβ̄ (q, λ)
λ λ(1 − q) 1 − λq
= Xβ̄ (q, λ)

and Xβ̄ (q, λ) is a link invariant.

Both the single variable polynomials may be retrieved from the HOMFLY
polynomial via the substitutions

VL (t) = XL (t, t)
. 
∆L (t) = XL t, t−1 .

Another definition of the HOMFLY polynomial is possible. For β ∈ B n suppose


that L = β̄, oriented so that the generator σi is a positive crossing (that is,
w(σi ) = 1). Suppose that β contains some σ i 1 and write

β = γ 1 σi γ2
1
a similar construction is possible for σi−1
Chapter 3. Polynomials 19

for γi ∈ Bn . By applying M1 we can define

L ∼ L0 = γσi

where γ = γ1 γ2 . Let
L+ = γσi2 and L− = γ̄.
The relation (iii) gives
γσi2 = (q − 1)γσi + qγ
so that
tr(γσi2 ) − q tr(γ) = (q − 1) tr(γσi ).
Let e = e(γ) be the exponent sum of γ. Then
1 p p
p p Xγσ2 − q λXγ̄
q λ i
 n−1  
1 1 − λq √ e+2
=p p −√ λ tr(γσi2 )
q λ λ(1 − q)
 n−1  
p p 1 − λq √ 2
− q λ −√ λ tr(γ)
λ(1 − q)
√ e+1
λ  n−1
1 − λq 
= √ −√ tr(γσi2 ) − q tr(γ)
q λ(1 − q)
√ e+1
λ  n−1
1 − λq
= √ −√ ((q − 1) tr(γσi ))
q λ(1 − q)
  n−1  
√ 1 1 − λq √ e+1
= q−√ − √ λ tr(γσ1 )
q λ(1 − q)
 
√ 1
= q − √ Xγσ1
q

and this shows that XL (q, λ) satisfies the skein relation


 
1 p p √ 1
p p XL+ (q, λ) − q λXL− (q, λ) = q − √ XL0 (q, λ).
q λ q

By introducing the substitution


p p √ 1
t= q λ and x = q− √
q
Chapter 3. Polynomials 20

we can define
PL (t, x) = XL (q, λ)
where PL (t, x) ∈ Z[t±1 , x±1 ] is computed recursively from the axioms

P (t, x) = 1 (3.5)

t−1 PL+ (t, x) − tPL− (t, x) = xPL0 (t, x). (3.6)

In this setting, L+ , L− and L0 are diagrams that are identical except for in a
small region where they differ as in

L+ L− L0

By a simple application of (3.6), the polynomial of the n component unlink is


 n−1
t−1 − t
(3.7)
x

in the skein definition of the HOMFLY polynomial. This agrees with (3.4)
under the substitutions
p p √ 1
t= q λ and x = q− √ .
q

As indicated earlier, both the Jones polynomial and the Alexander polynomial
may be computed recursively as they each satisfy a skein relation by specifying
 
√ 1
VL (t) = PL t, t − √
t
 
. √ 1
∆L (t) = PL 1, t − √ .
t
Chapter 4
Tangles and Linear Maps

4.1 Conway Tangles


In the recursive computation of the Kauffman bracket of a link, the order in
which the crossings are reduced is immaterial. In many cases it will be conve-
nient to group crossings together in the course of computation. From Conway’s
point of view [8], such groupings or tangles form the building blocks of knots
and links. In addition, this point of view will allow us to take advantage of
the well-developed tools of linear algebra.

Definition 4.1. Given a link L in S3 consider a 3-ball B 3 ⊂ S3 such that


∂B 3 intersects L in exactly 4 points. The intersection B 3 ∩ L is called a
Conway tangle (or simply, a tangle) denoted by T . The exterior of the tangle
S3 r B 3 ∩ L is called an external wiring, denoted by L r T .

Note that, as S3 r B 3 is a ball, the external wiring L r T is a tangle also.

Figure 4.1: Some diagrams of Conway tangles

A tangle, as a subset of a link, may be considered up to equivalence under


isotopy. When a diagram of the link L is considered, a tangle may be repre-
sented by a disk in the projection plane, with boundary intersecting the link

21
Chapter 4. Tangles and Linear Maps 22

in 4 points. Equivalence of tangle diagrams then, is given by the Reidemeister


moves, where the four boundary points are fixed.
Further, the Kauffman bracket of a tangle T may be computed by way of the
axioms (3.2) and (3.3). Thus the the Kauffman bracket of any tangle may be
written in terms of tangles having no crossings or closed loops. There are only
two such tangles, and they are denoted by

0= and ∞ = .

These tangles are fundamental in the sense that they form a basis for presenting
the bracket of a given tangle T . That is
     
T = x0 + x∞

where x0 , x∞ ∈ Z[a, a−1 ].


Definition 4.2. Let T be a Conway tangle and
 
 
  
T = x0 x∞  
 

where x0 , x∞ ∈ Z[a, a−1 ]. The bracket vector of T is denoted


 
br(T ) = x0 x∞ .

In this way, the Kauffman bracket divides Conway tangles into equivalence
classes completely determined by br(T ). For example,
     
−1
=a +a

and    
br = a a−1 .

We can define a product for tangles that is similar to multiplication in the


braid group. Given Conway tangles T and U the product T U is a Conway
tangle obtained by concatenation:
  
TU = = 

 

Notice that when T ∈ B2 this is exactly braid multiplication.


Chapter 4. Tangles and Linear Maps 23

Definition 4.3. The Kauffman bracket skein module S br is the Z[a, a−1 ]-
module generated by isotopy classes of Conway tangles, modulo equivalence
given by axioms (3.2) and (3.3) defining the Kauffman bracket.

The tangles {0, ∞} provide a module basis for S br so that the elements T ∈ S br
may be represented by br(T ).
Suppose the tangle T is contained in some link L. Then writing L = L(T ) and
considering T ∈ S br gives rise to a Z[a, a−1 ]-linear map
f : S br −→ Z[a, a−1 ] (4.1)
 
hL(0)i
T 7−→ br(T )
hL(∞)i
where    
L(0) = L and L(∞) = L .
This map is simply an evaluation map computing the bracket of L(T ) since
 
hL(0)i
hL(T )i = br(T ) = f (T ).
hL(∞)i

Given a tangle T , one may form a link in a number of ways by choosing an


external wiring. As with the previous construction, there are only two such
external wirings which do not produce any new crossings.
Definition 4.4. For any Conway tangle T we may form the numerator closure

TN =

and the denominator closure


TD =  .

Now returning to the link L(T ), recall that the external wiring L r T is itself
a tangle. Again, all crossings and closed loops may be eliminated using the
bracket axioms so that
 


 
hL r T i = br(L r T ) 
 



 
= hT N i hT D i br(L r T )>
Chapter 4. Tangles and Linear Maps 24

This gives rise to another Z[a, a−1 ]-linear evaluation map

S br −→ Z[a, a−1 ] (4.2)


 
L r T 7−→ hT N i hT D i br(L r T )>

In fact, combining the linear maps (4.1) and (4.2) forms a bilinear map

F : S br × S br −→ Z[a, a−1 ] (4.3)


(T, L r T ) 7−→ hL(T )i

where
 
hL(T )i = hT N i hT D i br(L r T )>
  N
  N
 
 x0 + x∞ 

=      br(L r T )>
D D 
x0 + x∞
 N
 
D

 
= br(T ) 


N
 
D
 br(L r T )>

 
δ 1
= br(T ) br(L r T )>
1 δ

so that given T, U ∈ S br we have


 
δ 1
F (T, U ) = br(T ) br(U )> .
1 δ

Definition 4.5. For Conway tangles T and U define the link J(T, U ) = (T U ) N
and call this the join of T and U .

We have that
hJ(T, U )i = F (T, U )
by definition, and further

L(T ) ∼ J(T, L r T ).

Definition 4.6. For Conway tangles T and U define the connected sum
T D #U D = (T U )D .
Chapter 4. Tangles and Linear Maps 25

Since any link L may be written as T D for some tangle T this definition gives
rise to a connected sum for links. It follows that
hL1 #L2 i = hL1 ihL2 i,
and
VL1 #L2 = VL1 VL2
provided orientations agree. A similar argument gives such an equality for the
HOMFLY polynomial, and hence the Alexander polynomial as well.

4.2 Conway Mutation


Consider a link diagram containing some Conway tangle T . We can choose
the coordinate system so that T is contained in the unit disk, for convenience
of notation. Further, we can arrange that the 4 points of intersection between
the link and the boundary of the disk are
 
1 1
±√ , ±√ , 0 .
2 2
Let ρ be a 180 degree rotation of the unit disk about any of the three coor-
dinate axis. Note that ρ leaves the external wiring unchanged and, for such a
projection, ρ fixes the boundary points as a set.
Definition 4.7. Given a link L(T ) where T ∈ S br define the Conway mutant
denoted by L(ρT ).

Notice that
ρ =

ρ =

ρ =

so that
 
hL(0)i
hL(T )i = br(T )
hL(∞)i
 
hL(0)i
= br(ρT )
hL(∞)i
= hL(ρT )i.
Chapter 4. Tangles and Linear Maps 26

Moreover, with orientation dictated by the external wiring


w(T ) = w(ρT )
so we have the following theorem.
Theorem 4.8. VL(T ) = VL(ρT ) .

While it may be that L(T )  L(ρT ), it is certain that this method does not
provide an answer to question 3.2: It can be shown that a Conway mutant of
the unknot is always unknotted [30]. Theorem 4.8 is in fact a corollary of a
stronger statement.
Theorem 4.9. PL(T ) = PL(ρT ) .

proof. Using the skein relation (3.6) defining the HOMFLY polynomial, it is
possible to decompose any tangle T into a linear combination of the form

T = a1 + a2 + a3

where ai ∈ Z[t±1 , x±1 ]. Therefore, these tangles provide a basis for presenting
the HOMFLY polynomial of a tangle T . Thus, we can define a Z[t ±1 , x±1 ]-
module S P generated by isotopy classes of tangles up to equivalence under
the skein relation. Moreover, if L = L(T ) then we have a Z[t ±1 , x±1 ]-linear
evaluation map
S P −→ Z[t±1 , x±1 ]
T 7−→ PL(T )
or, more generally, the bi-linear evaluation map
S P × S P −→ Z[t±1 , x±1 ]
(T, U ) 7−→ PJ(T,U ) .
Since the basis n o
, ,
is ρ-invariant, it follows that PT and PρT are equal hence

S P × S PR
RRR
RRR
R((
ρ×id Z[t±1 , x±1 ]
ll66
lll
 lll
SP × SP
Chapter 4. Tangles and Linear Maps 27

commutes and
PL(T ) = PL(ρT ) .

4.3 The Skein Module


Everything that has been said regarding tangles to this point can be stated in
a more general setting [30, 31].
Definition 4.10. Given a link L in S3 consider a 3-ball B 3 ⊂ S3 such that
∂B 3 intersects L in exactly 2n points. The intersection B 3 ∩ L is called an
n-tangle denoted by T . As before, the exterior of the n-tangle S 3 r B 3 ∩ L is
another n-tangle L r T called an external wiring.

In this setting, Conway tangles arise for n = 2 as 2-tangles.


Let Mn be the (infinitely generated) free Z[a, a −1 ]-module generated by the
set of equivalence classes of n-tangles. The axioms (3.2) and (3.3) defining the
bracket give rise to an ideal In ⊂ Mn generated by
     
−1
−a −a (4.4)
   
Tt −δ T (4.5)

where δ = −a−2 − a2 and the h i indicate that the rest of the tangle is left
unchanged.
Definition 4.11. The Z[a, a−1 ]-module

Sn = Mn In

is called the (Kauffman bracket) skein module. Note that S 2 = S br .

Due to the form of In it is possible to choose representatives for each equiv-


alence class in Sn that have neither crossings nor closed loops. These tangles
form a basis for Sn . We have seen, for example, that S2 is 2-dimensional as a
module, with basis given by the fundamental Conway tangles

and .
Chapter 4. Tangles and Linear Maps 28

Theorem 4.12. Sn has dimension


(2n)!
Cn =
n!(n + 1)!
as a module.

proof. Simply put, we need to determine how many n-tangles there are that
have no crossings or closed loops. That is, given a disk in the plane with 2n
marked points on the boundary, how many ways can the points be connected
by non-intersecting arcs (up to isotopy)?
Clearly, C1 = 1, and as discussed earlier C2 = 2.
Now suppose n > 2 and consider a disk with 2n points on the boundary.
Starting at some chosen boundary point and numbering clockwise, the point
labeled 1 must connect to an even labeled point, say 2k. This arc divides the
disk in two: One disk having 2(k − 1) marked points, the other with 2(n − k).
Therefore
n
X
Cn = Ck−1 Cn−k
k=1
= C0 Cn−1 + C1 Cn−2 + · · · + Cn−1 C0
where C0 = 1 by convention. Now consider the generating function

X
f (x) = Ci xi
i=0

and notice !

X i
X
2
(f (x)) = Ck−1 Cn−k xi
i=0 k=1
so that
x(f (x))2 = f (x) − 1
and √
1 − 4x
1−
f (x) = .
2x

To deduce the coefficients of f (x), first consider the expansion of z about 1.

X

z= di (z − 1)i
i=0
Chapter 4. Tangles and Linear Maps 29

where (
1 i=0
di = (−1)i−1 (2i−3)!! .
2i i!
i>0
Therefore with z = 1 − 4x

X

1 − 4x = di (−4x)i
i=0

X
=1+ (−1)i 2i 2i di xi
i=1
and

1 X
f (x) = − (−1)i 2i 2i di xi
2x
i=1

X
= (−1)i−1 2i 2i−1 di xi−1
i=1
so that
Ci = (−1)i 2i+1 2i di+1
(−1)i (2i − 1)!!
= (−1)i − 2i+1 2i
2i+1 (i + 1)!
2i (2i − 1)!!
=
(i + 1)!
for i > 1. Finally, the fact that
2n (2n − 1)!! (2n)!
=
(n + 1)! n!(n + 1)!
follows by induction since
2n+1 (2(n + 1) − 1)!! 2(2n + 1) 2n (2n − 1)!!
=
(n + 2)! n+2 (n + 1)!
2(n + 1)(2n + 1)
= Cn
(n + 2)(n + 1)
(2n + 2)(2n + 1) (2n)!
=
(n + 2)(n + 1) n!(n + 1)!
(2(n + 1))!
= .
(n + 1)!(n + 2)!
Chapter 4. Tangles and Linear Maps 30

As an example of theorem 4.12, the 5-dimensional module S 3 has basis given


by n o
, , , , . (4.6)

Let a given n-tangle diagram be contained in the unit disk so that, of the
2n boundary points, n have positive x-coordinate while the remaining n have
negative x-coordinate. With this special position, multiplication of tangles
by concatenation, as introduced for Conway tangles, extends to all n-tangles.
When two n-tangles are in fact n-braids, we are reduced to multiplication in
Bn . With this multiplication, Sn has an algebra structure called the Temperly-
Lieb algebra [21, 22].
The n-dimensional Temperly-Lieb algebra TL n over Z[a, a−1 ] has generators
e1 , e2 , . . . , en−1 and relations

e2i = δei (4.7)


ei ej = e j ei for |i − j| > 1 (4.8)
ei ej ei = e i for |i − j| = 1. (4.9)

The multiplicative identity for this algebra is exactly the identity in B n , and
the generators are tangles of the form shown in figure 4.2.









Figure 4.2: The generator ei

For example, TL3 is generated by {e1 , e2 }, while the basis for the module S3
is the set of elements {1, e1 , e2 , e1 e2 , e2 e1 } as in (4.6).
Notice that there is a representation of the braid group via the Kauffman
Chapter 4. Tangles and Linear Maps 31

bracket given by

Bn −→ TLn
σi 7−→ a + a−1 ei
σi−1 7−→ aei + a−1 .

4.4 Linear Maps


Let L = L(T1 , . . . , Tk ) be a link where {Ti } is a collection of subtangles Ti ⊂ L.
If Ti0 is a tangle such that Ti = Ti0 as elements of Sn then, in the most general
setting, L0 = L(T10 , . . . , Tk0 ) is a mutant of L (relative to the Kauffman bracket).
Therefore when w(L) = w(L0 ), we have that

V L = V L0 .

Of course, it may be that L  L0 and this approach has been used in attempts
to answer question 3.2 [30, 31].
Let’s first revisit Conway mutation in this context. We have, given the bilinear
evaluation map F and a 180 degree rotation ρ, the commutative diagram

S2 × S2P
PPP F
PPP
P((
ρ×id Z[a, a−1 ]
nn66
nnnn
 n F
n
S2 × S 2

since F = F ◦ (ρ × id). We saw that the linear transformation ρ was in fact


the identity transformation on S2 , and as a result the link L = J(T, U ) and
the mutant L0 = J(ρT, U ) have the same Kauffman bracket.
A possible generalization arises naturally at this stage. As was pointed out
earlier, it is possible to construct a link from two tangles in many different and
complicated ways. Starting with T ⊂ B T3 ⊂ S3 and U ⊂ BU3 ⊂ S3 , the link
L(T, U ) is constructed by choosing an external wiring of S 3 r (BT3 ∪ BU3 ). In
Chapter 4. Tangles and Linear Maps 32

this setting we have


 
hL(0, U )i
hL(T, U )i = br(T )
hL(∞, U )i
 
hL(0, 0)i hL(0, ∞)i
= br(T ) br(U )>
hL(∞, 0)i hL(∞, ∞)i
= br(T ) L br(U )>

which gives rise to the bilinear map

G : S2 × S2 −→ Z[a, a−1 ]
(T, U ) 7−→ br(T ) L br(U )> .

If additionally there is a linear transformation

τ : S2 × S2 −→ S2 × S2
(T, U ) 7−→ (τ1 T, τ2 U )

which acts as the identity as a linear transformation of modules, then we have


the commutative diagram

S2 × S2P
PPPG
PPP
P((
τ =τ1 ×τ2 Z[a, a−1 ]
nn66
nnn
 nnn G
S2 × S 2

and finally, if w(L(T, U )) = w(L(τ1 T, τ2 U )) we can conclude that

VL(T,U ) = VL(τ1 T,τ2 U ) .

4.5 Braid Actions


The three strand braid group has presentation

B3 = hσ1 , σ2 |σ1 σ2 σ1 = σ2 σ1 σ2 i

where
σ1 = and σ2 = .
Chapter 4. Tangles and Linear Maps 33

Figure 4.3: The tangle T β ∈ S2

Given T ∈ S2 and a β ∈ B3 we can define a new 2-tangle denoted T β as in


figure 4.3.

Proposition 4.13. The map

S2 × B3 −→ S2
(T, β) 7−→ T β

is a well defined group action.

proof. Let idB3 be the identity braid. Then for any tangle T we have

T idB3 = T

since


∼ 

For β, β 0 ∈ B3 , the product ββ 0 is defined by concatenation so that


0 0
(T β )β = T ββ

by planar isotopy of the diagram

  
Chapter 4. Tangles and Linear Maps 34

Proposition 4.14. For T ∈ S2 and β ∈ B3


 −3 
−a 0
br(T σ1 ) = br(T ) (4.10)
a−1 a
 
σ2 a a−1
br(T ) = br(T ) . (4.11)
0 −a−3

proof. Applying the action of σ1 to an arbitrary tangle T , we may decompose


T σ1 into the tangles

and

in S2 . Relaxing these diagrams gives rise to the following computation in S 2 :


 
 
br(T σ1 ) = br(T ) 

 

    
a + a −1 δ
 
= br(T ) 

   

a +a −1

 −3 
−a 0
= br(T ) −1
a a
Chapter 4. Tangles and Linear Maps 35

Similarly, for the action of σ2 :


 
 
br(T σ2 ) = br(T ) 

 

    
+a −1
a 
= br(T ) 

   

a −1
+a δ
 
a a−1
= br(T )
0 −a−3

This gives rise to a group homomorphism

Φ : B3 −→ GL2 (Z[a, a−1 ])


 −3 
−a 0
σ1 7−→
a−1 a
 
a a−1
σ2 7−→
0 −a−3
since
  
−a−2 −a−4 −a−3 0
Φ(σ1 σ2 σ1 ) =
1 0 a−1 a
 
0 −a−3
= −3
−a 0
   −2 
a a−1 −a −a−4
=
0 −a−3 1 0
= Φ(σ2 σ1 σ2 ).

Question 4.15. Is this representation of B 3 faithful?

With the B3 -action on S2 , consider the linear transformation given by

β : S2 × S2 −→ S2 × S2
−1
(T, U ) 7−→ (T β , U β ).
Chapter 4. Tangles and Linear Maps 36

For a link of the form L(T, U ), this leads to the definition of a new link
−1
L(T β , U β ).
Denote the evaluation matrix of L(T, U ) by
 
hL(0, 0)i hL(0, ∞)i
L=
hL(∞, 0)i hL(∞, ∞)i

and suppose that L ∈ GL2 (Z[a, a−1 ]), that is, det(L) 6= 0. Note that

hL(T, U )i = br(T ) L br(U )>


−1
hL(T β , U β )i = br(T )Φ(β) L (Φ(β −1 ))> br(U )> .

So, defining a second B3 -action

B3 × GL2 (Z[a, a−1 ]) −→ GL2 (Z[a, a−1 ])


(β, L) 7−→ Φ(β) L (Φ(β −1 ))> ,

we are led to an algebraic question. When a non-trivial β ∈ B 3 gives rise to


a fixed point under this action, the linear transformation given by β is the
identity. Thus G = G ◦ β and we have the commutative diagram

S2 × S2P
PPPG
PPP
P((
β Z[a, a−1 ]
nn66
nnn
 nnn G
S2 × S 2

where L ∈ Fix(β), so that


−1
hL(T, U )i = hL(T β , U β )i.

In particular, we would like to study the case where


−1
L(T, U )  L(T β , U β ).

Question 4.16. For a given link L(T, U ) with evaluation matrix L ∈ GL 2 (Z[a, a−1 ]),
what are the elements β ∈ B3 such that L ∈ Fix(β)?

This question is the main focus of the following chapters.


Chapter 5
Kanenobu Knots

5.1 Construction
Shortly after the discovery of the HOMFLY polynomial, Kanenobu introduced
families of distinct knots having the same HOMFLY polynomial and hence the
same Jones and Alexander polynomials as well [18]. It turns out that these
knots are members of a much larger class of knots which we will denote by
K(T, U ) for tangles T, U ∈ S2 .

Figure 5.1: The Kanenobu knot K(T, U )

Proposition 5.1. Suppose x is a non-trivial polynomial in Z[a, a −1 ] so that


 
x δ
X = ∈ GL(Z[a, a−1 ])
δ δ2
where δ = −a−2 − a2 . Then Φ(σ2 ) X Φ(σ2−1 )> = X and X ∈ Fix(σ2 ) under
the B3 -action on GL(Z[a, a−1 ]).

proof. Since
   
a a−1 a−1 a
Φ(σ2 ) = and Φ(σ2−1 ) =
0 −a−3 0 −a3

37
Chapter 5. Kanenobu Knots 38

we have
      
a a−1 x δ a−1 0 ax + a−1 δ aδ + a−1 δ 2 a−1 0
=
0 −a−3 δ δ 2 a −a3 −a−3 δ −a−3 δ 2 a −a3
 
x + a−2 δ + a2 δ + δ 2 −a4 δ − a2 δ 2
=
−a−4 δ − a−2 δ 2 δ2
 
x + δ(a−2 + a2 + δ) δ(−a4 − a2 δ)
=
δ(−a−4 − a−2 δ) δ2
 
x δ(−a4 + 1 + a4 )
=
δ(−a−4 + a−4 + 1) δ2
 
x δ
=
δ δ2
with δ = −a−2 − a2 .

To compute the bracket of the Kanenobu knot K(T, U ), we need the evaluation
matrix
 
* + * +
 
 
 
 
 
 
K= 
.

* + * + 
 
 
 
 
 

Since K(0, 0) is the connected sum of two figure eight knots (the figure eight
is denoted 41 as in [32]), we can compute
D E D E
41 = σ2 σ1−1 σ2 σ1−1

using the Temperly-Lieb algebra TL3 . We have


σ2 σ1−1 = 1 + a2 e1 + a−2 e2 + e2 e1
as an element of TL3 so that
2
σ2 σ1−1 = (1 + a2 e1 + a−2 e2 + e2 e1 )2
= 1 + (3a2 +a4 δ)e1 + (3a−2 +a−4 δ)e2 + e1 e2 + (4+a2 δ+a−2 δ)e2 e1 .
Chapter 5. Kanenobu Knots 39

Then the Kauffman bracket is given by


 

−1 2
σ2 σ1 = δ 2 + (3a2 + a4 δ)δ + (3a−2 + a−4 δ)δ + 1 + (4 + a2 δ + a−2 δ)(1)
= δ(−a−6 + 2a−4 + 2a4 − a6 ) + 5
= a−8 − a−4 + 1 − a4 + a8

and
2
h41 #41 i = a−8 − a−4 + 1 − a4 + a8
= a−16 − 2a−12 + 3a−8 − 4a−4 + 5 − 4a4 + 3a8 − 2a12 + a16 .

In addition, it can be seen from the braid closure σ 2 σ1−1 σ2 σ1−1 ∼ 41 that
w(41 ) = 0 and hence

w (K(0, 0)) = w(41 #41 ) = 0.

Thus the Jones polynomial of K(0, 0) is given by

VK(0,0) = a−16 − 2a−12 + 3a−8 − 4a−4 + 5 − 4a4 + 3a8 − 2a12 + a16 .

Now the evaluation matrix for K(T, U ) is given by


 −8 
(a − a−4 + 1 − a4 + a8 )2 δ
K=
δ δ2

since the three entries for K(0, ∞), K(∞, 0) and K(∞, ∞) are all equivalent to
unlinks with no crossings via applications of the Reidemeister move R 2 (recall
that R2 leaves the Kauffman bracket unchanged).
For any tangle diagram T , denote by T ? the tangle diagram obtained by switch-
ing each crossing of T . That is, for any choice of orientation

w(T ? ) = −w(T ).

This can be extended to knot diagrams K, where K ? is the diagram such that

w(K ? ) = −w(K)

so that K ? is the mirror image of K.


When U = T ? ,
w(K(T, U )) = 0
Chapter 5. Kanenobu Knots 40

the bilinear evaluation map for the bracket

G(T, U ) = br(T ) K br(U )>

computes the Jones polynomial

VK(T,U ) = br(T ) K br(U )> .

Since K is of the form given in proposition 5.1, the bilinear map defined by
the braid σ2n ∈ B3

σ2n : S2 × S2 → S2 × S2
n −n
(T, U ) 7→ (T σ2 , U σ2 )

is the identity transformation for every n ∈ Z. Moreover, when U = T ?


 n −n

w K(T σ2 , U σ2 ) = 0

so we have the following theorem.

Theorem 5.2. When U = T ? , the family of knots given by


n −n
K(T σ2 , U σ2 )

for n ∈ Z are indistinguishable by the Jones polynomial.

Of course, that these are in fact distinct knots remains to be seen.


Kanenobu’s original knot families [18] can be recovered from
 2n 2m

Kn,m = K T σ2 , T σ2

where n, m ∈ Z and T is the 0-tangle.

Theorem (Kanenobu). Kn,m and Kn0 ,m0 have the same HOMFLY polyno-
mial when |n−m| = |n0 −m0 |. Moreover, when (n, m) and (n0 , m0 ) are pairwise
distinct, these knots are distinct.

The knots of Kanenobu’s theorem are distinguished by their Alexander module


structure [18].
Chapter 5. Kanenobu Knots 41

5.2 Basic Examples


Consider the Kanenobu knots
 
K0 = K ,
 
K1 = K ,
 
K2 = K ,

and notice that by applying the action of σ 2 we have


σ2 σ2
K0 // K1 // K2

so that by construction, these knots have the same Jones polynomial.


First recall that the HOMFLY polynomial of the n-component unlink

···
| {z }
n

is given by
 n−1
t−1 − t
.
x
As this polynomial will be used often, we define
t−1 − t
P0 = .
x
Now, using the skein relation (3.6) defining the HOMFLY polynomial P (t, x),
we can compute

−t = −t−1 +x

= t−2 − t−1 x

and
t−1 =t +x

= t2 + tx .
Chapter 5. Kanenobu Knots 42

Combining these tangles pairwise, we have


     
−1
, = , +t x ,
   
2
− tx , −x ,

giving rise to equality among HOMFLY polynomials

PK2 = PK0 + (t−1 x − tx)P0 − x2 P02


 2
t−1 − t t−1 − t
= PK0 + (t−1 x − tx) − x2
x x
= P K0 .

This common polynomial1 is

PK0 (t, x) = (t−4 − 2t−2 + 3 − 2t2 + t4 ) + (−2t−2 + 2 − 2t2 )x2 + x4 .

Notice that this is in agreement with Kanenobu’s theorem.


On the other hand K1 has HOMFLY polynomial

PK1 (t, x) = (2t−2 − 3 + 2t2 ) + (3t−2 − 8 + 3t2 )x2 + (t−2 − 5 + t2 )x4 − x6

and we can conclude that K0 and K1 are distinct knots despite having the
same Jones polynomial:

a−16 − 2a−12 + 3a−8 − 4a−4 + 5 − 4a4 + 3a8 − 2a12 + a16

Further, applying theorem 4.9, these knots cannot be Conway mutants as they
have different HOMFLY polynomials.

5.3 Main Theorem


Theorem 5.3. For each 2-tangle T there exists a pair of external wirings for
T that produce distinct links that have the same Jones polynomial. Moreover,
the links obtained are not Conway mutants.
1
computed using KNOTSCAPE
Chapter 5. Kanenobu Knots 43

Figure 5.2: Distinct knots that are not Conway mutants

proof. Take U = T ? (that is, the tangle such that w(U ) = −w(T )) and define
Kanenobu knots for the pair (T, U )

K = K(T, U ) and K σ2 = K(T σ2 , U σ2 ).

Then, by construction, we have that

V K = V K σ2 .

It remains to show that these are in fact distinct knots. To see this, we compute
the HOMFLY polynomials PK and PK σ2 .
Now with the requirement that the tangle U = T ? , there are two choices of
orientations for the tangles that are compatible with an orientation of the knot
(or possibly link, in which case a choice of orientation is made) K(T, U ). They
are  
Type 1 , 

 
Type 2 

, 

so we proceed in two cases.


Type 1. Using the skein relation we can decompose


= aT + bT


= aU + bU
Chapter 5. Kanenobu Knots 44

where aT , bT , aU , bU ∈ Z[t± , x± ]. Combining pairwise we obtain


     
,  = a T aU , + a T bU ,
   
+ b T aU , + b T bU ,

so that

PK = aT aU PK1 + (aT bU + bT aU )P0 + bT bU P02


 −1   −1 2
t −t t −t
= aT aU PK1 + (aT bU + bT aU ) + b T bU
x x
= a T a U P K1 + R
 −1   −1 2
where R = (aT bU + bT aU ) t x−t + bT bU t x−t . Now applying the action
of σ2 we have

 

−1
T σ2 = and U σ2 =

therefore
     
σ2 σ2−1
T ,U = a T aU , + a T bU ,
   
+ b T aU , + b T bU ,
   
= a T aU , + a T bU ,
   
+ b T aU , + b T bU ,

so that

PK σ2 = aT aU PK2 + (aT bU + bT aU )P0 + bT bU P02


= P K0 + R
Chapter 5. Kanenobu Knots 45

since PK2 = PK0 . However, as PK1 6= PK0 we have that


PK 6= PK σ2
giving rise to distinct knots.
Type 2. As before, the tangles are decomposed via the HOMFLY skein
relation
= aT + bT


= aU + bU

for some other aT , bT , aU , bU ∈ Z[t± , x± ]. Combining pairwise


     


,  = a T aU , + a T bU ,
   
+ b T aU , + b T bU ,

and
PK = aT aU PK0 + (aT bU + bT aU )P0 + bT bU P02
 −1   −1 2
t −t t −t
= aT aU PK0 + (aT bU + bT aU ) + b T bU
x x
= a T a U P K0 + R
with R ∈ Z[t± , x± ] as before. Again, applying the action of σ 2 we have
     
σ2 σ2−1
T ,U = a T aU , + a T bU ,
   
+ b T aU , + b T bU ,
   
= a T aU , + a T bU ,
   
+ b T aU , + b T bU ,

and the HOMFLY polynomial


PK σ2 = aT aU PK1 + (aT bU + bT aU )P0 + bT bU P02
= aT aU PK1 + R.
Chapter 5. Kanenobu Knots 46

Once again, as PK1 6= PK0 we have that

PK 6= PK σ2

giving rise to distinct knots.


Finally, as K and K σ2 have distinct HOMFLY polynomials in both cases, it
follows from theorem 4.9 that these knots cannot be Conway mutants.

5.4 Examples
1. Consider the Kanenobu knot K(T, U ) where

T = and U =

as in figure 5.3.

Definition 5.4. A tangle is called rational if it is of the form T β where β ∈ B3


and the tangle T is either the 0-tangle or the ∞-tangle. This is equivalent to
Conway’s definition for rational tangles [8].

As a result, the computation of the bracket for rational tangles is straightfor-


ward. In this example we have
σ1−3 σ13
T = and U =

so that
 
  −a−3 0 −3
br(T ) = 0 1
a−1 a
 3  3  3 
  −a 0 −a 0 −a 0
= 0 1
a a−1 a a−1 a a−1
 
  −a9 0
= 0 1
a−1 − a3 + a7 a−3
 −1 
= a − a3 + a7 a−3
Chapter 5. Kanenobu Knots 47

and
 
  −a−3 0 3
br(U ) = 0 1
a−1 a
   
  −a−3 0 −a−3 0 −a−3 0
= 0 1
a−1 a a−1 a a−1 a
 −7 
= a − a−1 + a a3 .

Now

hK(T, U )i = br(T ) K br(U )>


 −8 
(a − a−4 + 1 − a4 + a8 )2 −a−2 − a2
= br(T ) br(U )>
−a−2 − a2 a−4 + 2 + a4
= a−24 − 4a−20 + 10a−16 − 19a−12 + 27a−8 − 33a−4
+ 37 − 33a4 + 27a8 − 19a12 + 10a16 − 4a20 + a24

hence

VK(T,U ) =a−24 − 4a−20 + 10a−16 − 19a−12 + 27a−8 − 33a−4


+ 37 − 33a4 + 27a8 − 19a12 + 10a16 − 4a20 + a24 .

Figure 5.3: Example 1

This gives a collection of knots having the same Jones polynomial

VK(T,U ) = V n −n
K(T σ2 ,U σ2 )
Chapter 5. Kanenobu Knots 48

n −n
since w(K(T σ2 , U σ2 )) = 0 for all n ∈ Z.
−1
The Knots K(T, U ) and K(T σ2 , U σ2 ) are distinct as they have different HOM-
FLY polynomials2 :

XK(T,U ) = − 1 + (6t−2 − 12 + 6t2 )x2 + (9t−2 − 24 + 9t2 )x4


+ (5t−2 − 19 + 5t2 )x6 + (t−2 − 7 + t2 )x8 − x10

X −1 =(t−4 − 4t−2 + 7 − 4t2 + t4 )


K(T σ2 ,U σ2 )
+ (2t−4 − 7t−2 + 10 − 7t2 + 2t4 )x2
+ (t−4 − 6t−2 + 8 − 6t2 + t4 )x4
+ (−2t−2 + 4 − 2t2 )x6 + x8

In particular, these knots cannot be Conway mutants in view of theorem 4.9.


2. Now consider the case when T, U are not rational tangles. For this example
take

T = and U =

in K(T, U ). We can compute


 
br(T ) = a−5 − 2a−1 + a3 − a7 −a−11 + 2a−7 − 2a−3 + a
 
br(U ) = −a−7 + a−3 − 2a + a5 a−1 − 2a3 + 2a7 − a11

so that

hK(T, U )i = −a−28 + 5a−24 − 15a−20 + 31a−16 − 52a−12 + 73a−8 − 88a−4


+ 95 − 88a4 + 73a8 − 52a12 + 31a16 − 15a20 + 5a24 − a28 .

Again we have a family of knots (distinct from those of example 1) such that

VK(T,U ) = V n −n
K(T σ2 ,U σ2 )

n −n
since w(K(T σ2 , U σ2 )) = 0 for all n ∈ Z.
2
computed using KNOTSCAPE
Chapter 5. Kanenobu Knots 49

Figure 5.4: Example 2

−1
The knots K(T, U ) and K(T σ2 , U σ2 ) of this example are distinct as can be
seen by the HOMFLY polynomials3 :

XK(T,U ) =(−2t−6 + 9t−4 − 22t−2 + 31 − 22t2 + 9t4 − 2t6 )


+ (−3t−6 + 22t−4 − 58t−2 + 82 − 58t2 + 22t4 − 3t6 )x2
+ (−t−6 + 21t−4 − 67t−2 + 94 − 67t2 + 21t4 − t6 )x4
+ (8t−4 − 44t−2 + 62 − 44t2 + 8t4 )x6
+ (t−4 − 15t−2 + 28 − 15t2 + t4 )x8
+ (−2t−2 + 8 − 2t2 )x10 + x12

X −1 =(−4t−4 + 12t−2 − 15 + 12t2 − 4t4 )


K(T σ2 ,U σ2 )
+ (−12t−4 + 56t−2 − 84 + 56t2 − 12t4 )x2
+ (−13t−4 + 99t−2 − 176 + 99t2 − 13t4 )x4
+ (−6t−4 + 87t−2 − 197 + 87t2 − 6t4 )x6
+ (−t−4 + 41t−2 − 130 + 41t2 − t4 )x8
+ (10t−2 − 51 + 10t2 )x10 + (t−2 − 11 + t2 )x12 − x14

Once again, from theorem 4.9 it follows that these knots are not related by
Conway mutation.
3
computed using KNOTSCAPE
Chapter 5. Kanenobu Knots 50

5.5 Generalisation
We saw in proposition 5.1 under the action
B3 × GL2 (Z[a, a−1 ]) −→ GL2 (Z[a, a−1 ])
that   
x δ −1

∈ GL2 Z[a, a ] ⊂ Fix(σ2 ).
δ δ2
As a final task for this chapter, we’ll define a family of links that generate such
evaluation matricies.
Consider a slightly different diagram of the knot K(T, U ), given in figure 5.5.

Figure 5.5: Another diagram of the Kanenobu knot K(T, U )

From this diagram, we are led to define a rather exotic braid closure that will
be of use. That is, for the pair of tangles (T, U ) and an appropriately chosen
braid β ∈ B6 we define a link |β| as in figure 5.6.


 

Figure 5.6: The link |β|

It remains to describe which braids in B 6 give rise to an evaluation matrix of


the appropriate form. For this we will need two braid homomorphisms.
Let N  n and define, for each non-negative m ∈ Z, the inclusion homomor-
phism
im : Bn −→ BN
σk 7−→ σk+m
Chapter 5. Kanenobu Knots 51

for each k ∈ {1, . . . , n − 1}. Note that when m = 0, this is reduces to the
natural inclusion Bn < BN . Now the group B3 ⊕ B3 arises as a subgroup of
B6 by choosing

B3 ⊕ B3 −→ B6
(α, β) 7−→ i0 (α)i3 (β).

Notice that the image i0 (α)i3 (β) contains no occurrence of the generator σ 3
and hence
i0 (α)i3 (β) = i3 (β)i0 (α)
in B6 . Now define the switch homomorphism

s : B3 −→ B3
σ1 7−→ σ2−1
σ2 7−→ σ1−1

and note that, given a 180 degree rotation ρ in the projection plane, ρ(sβ) =
β −1 .

Definition 5.5. For each α ∈ B3 define the Kanenobu braid i0 (α)i3 (sα) ∈ B6 .


 

Figure 5.7: The closure of a Kanenobu braid

Theorem 5.6. Let β be a Kanenobu braid. The evaluation matrix X associ-


ated with the link |β| is an element of Fix(σ 2 ).

proof. We need to compute


 
hK(0, 0)i hK(0, ∞)i
X =
hK(∞, 0)i hK(∞, ∞)i
Chapter 5. Kanenobu Knots 52

where K(T, U ) = |β| for some Kanenobu braid β. Since ρ(sα) = α −1 , the link
K(∞, ∞) is of the form

which is simply the (usual) braid closure αα −1 . Hence


D E D E
K(∞, ∞) = αα−1 = δ 2

since the group operation σi σi−1 coresponds to the Reidemeister move R 2 , so


that the Kauffman bracket is unchanged. Similarly, the link K(∞, 0) (equiva-
lently, K(0, ∞)) is of the form

which reduces, canceling α−1 α, to the link

so that
hK(∞, 0)i = hK(0, ∞)i = δ.
Finally, let the polynomial hK(0, 0)i = x so that
 
x δ
X =
δ δ2

is an element of Fix(σ2 ).
Chapter 5. Kanenobu Knots 53

5.6 More Examples


For the following examples, we introduce the shorthand K(σ 2 , σ2−1 ) referring
to the knot obtained from the action
σ2
K(0, 0) // K(σ , σ −1 ).
2 2

We continue numbering of examples from section 5.4.


3. Taking the braid σ13 σ2−1 σ1 ∈ B3

Figure 5.8: The braid σ13 σ2−1 σ1

we can form the Kanenobu braid (σ13 σ2−1 σ1 )(σ5−3 σ4 σ5−1 ) ∈ B6

Figure 5.9: The Kanenobu braid (σ13 σ2−1 σ1 )(σ5−3 σ4 σ5−1 )

and the generalized Kanenobu knot



K(T, U ) = (σ13 σ2−1 σ1 )(σ5−3 σ4 σ5−1 ) .

K(T, U ) has evaluation matrix


 
x δ
K=
δ δ2

where
x = − a−20 + 2a−16 − 4a−12 + 6a−8 − 7a−4
+ 9 − 7a4 + 6a8 − 4a12 + 2a16 − a20 .
Chapter 5. Kanenobu Knots 54


Figure 5.10: The knot (σ13 σ2−1 σ1 )(σ5−3 σ4 σ5−1 )

Since w(K(T, U )) = 0 when U = T ? , the Jones Polynomial is given by

VK(T,U ) = br(T ) K br(U ).

In particular,
VK(0,0) = x
and we have that
VK(σ2 ,σ−1 ) = x.
2

It fact K(0, 0) ∼ 52 # 52? (following the notation in [32]).


These knots are distinct and not Conway mutants, as can be seen from the
HOMFLY polynomial4 of K(0, 0)

(−4t−2 + 9 − 4t2 ) + (−8t−2 + 20 − 8t2 )x2


+ (−5t−2 + 18 − 5t2 )x4 + (−t−2 + 7 − t2 )x6 + x8

while the HOMFLY polynomial of K(σ2 , σ2−1 ) is

(−t−4 + 3 − t4 )
+ (−t−4 + t−2 + 4 − t2 + t4 − t6 )x2
+ (t−2 + 2 + t2 )x4

4
computed using KNOTSCAPE
Chapter 5. Kanenobu Knots 55

Figure 5.11: The knot K(0, 0) ∼ 52 # 52?

Figure 5.12: The knot K(σ2 , σ2−1 )


Chapter 5. Kanenobu Knots 56

4. Taking the braid σ12 σ2−3 σ1 ∈ B3

Figure 5.13: The braid σ12 σ2−3 σ1

we can form the Kanenobu braid (σ12 σ2−3 σ1 )(σ5−2 σ43 σ5−1 ) ∈ B6

Figure 5.14: The Kanenobu braid (σ12 σ2−3 σ1 )(σ5−2 σ43 σ5−1 )

and the generalized Kanenobu knot



K(T, U ) = (σ12 σ2−3 σ1 )(σ5−2 σ43 σ5−1 ) .

K(T, U ) has evaluation matrix


 
x δ
K=
δ δ2
where
x =a−24 − 2a−20 + 4a−16 − 7a−12 + 9a−8 − 11a−4
+ 13 − 11a4 + 9a8 − 7a12 + 4a16 − 2a20 + a24 .

Note that K(0, 0) ∼ 61 # 61? (following the notation in [32]).


Since w(K(T, U )) = 0 when U = T ? , the Jones Polynomial is given by

VK(T,U ) = br(T ) K br(U ).

In particular,
VK(0,0) = x
and we have that
VK(σ2 ,σ−1 ) = x.
2
Chapter 5. Kanenobu Knots 57


Figure 5.15: The knot (σ12 σ2−3 σ1 )(σ5−2 σ43 σ5−1 )

Once again we obtain distinct knots. The HOMFLY polynomial 5 of K(0, 0) is

(4t−2 − 7 + 4t2 ) + (16t−2 − 36 + 16t2 )x2


+ (17t−2 − 50 + 17t2 )x4 + (7t−2 − 31 + 7t2 )x6
+ (t−2 − 1 + t2 )x8 − x10

While the HOMFLY polynomial of K(σ2 , σ2−1 ) is

(t−6 − t−4 − t−2 + 3 − t2 − t4 + t6 )


+ (−2t−4 − t−2 + 2 − t2 − 2t4 )x2 + (t−2 + 2 + t2 )x4

Once again we have distinct knots that have the same Jones polynomial but
are not related by Conway mutation.

5
computed using KNOTSCAPE
Chapter 5. Kanenobu Knots 58

Figure 5.16: The knot K(0, 0) ∼ 61 # 61?

Figure 5.17: The knot K(σ2 , σ2−1 )


Chapter 5. Kanenobu Knots 59

5.7 Observations
Definition 5.7. If a link is equivalent to a 3-braid, closed as in figure 5.18, it
is called a 2-bridge link.

Figure 5.18: The 2-bridge link obtained from β ∈ B 3 .

For a generalised Kanenobu knot K(T, U ), the knot K(0, 0) is always of the
form L#L? where L is a 2-bridge link.
In the case where K(0, 0) ∼ K#K ? is a connected sum of 2-bridge knots with
more that 3 crossings, such a K is generated by taking the 2-bridge closure of
an element α ∈ B3 . Such a braid generates the Kanenobu braid i 0 (α)i2 (sα),
and taking the closure
|i0 (α)i2 (sα)| = K(T, U )
with U = T ? gives rise to the evaluation matrix
 
hK#K ? i δ
K=
δ δ2

since K#K ? = K(0, 0). Now K ∈ Fix(σ2 ), and with the specification that
U = T ? , the familly of knots
 −1

K T σ2 , U σ2

share the common Jones polynomial

V n −n = hK#K ? i.
K(T σ2 ,U σ2 )

By recycling the argument of theorem 5.3, we can reduce the comparison of


the knots  
−1
K(T, U ) and K T σ2 , U σ2
to the comparison of the HOMFLY polynomials

PK(0,0) and PK(σ2 ,σ−1 ) .


2
Chapter 5. Kanenobu Knots 60

As shown by the previous examples, this generates further pairs of distinct


knots that are not Conway mutants despite sharing the same Jones polynomial.
The notable exception is the square knot, obtained from the connected sum of
trefoil knots 31 # 3?1 . This is the connected sum of 2-bridge knots. It can be
seen as K(0, 0) in the closure

K(T, U ) = (σ1 σ −1 σ1 )(σ −1 σ4 σ −1 )
2 5 5

but another view is given in figure 5.19.

Figure 5.19: The square knot 31 # 3?1 .

From the diagram in figure 5.20 it can be seen that the action of σ 2 can cancel
along a band connecting the tangles.


Figure 5.20: The knot (σ1 σ2−1 σ1 )(σ5−1 σ4 σ5−1 ) .

This cancelation is of the form




∼ 

so in the case that T is a rational tangle, their knot type is unaltered, while
a more general tangle results in a Conway mutant of the original diagram. In
particular, there is no change to the Jones polynomial.
In general however, the set of tangles S 2 together with the set of 2-bridge
knots (generated by B3 ) provide a range of knots (and even links) having
evaluation matricies contained in Fix(σ 2 ). In the cases discussed and the
examples produced, we have seen that the HOMFLY polynomial may be used
to distingush these knots. Thus, we conclude that this method of producing
Chapter 5. Kanenobu Knots 61

famillies knots sharing a common Jones polynomial is distinct from Conway


mutation.
In the next chapter, the results of Eliahou, Kauffman and Thistlethwaite [9]
will be restated using the braid actions introduced in this paper.
Chapter 6
Thistlethwaite Links

6.1 Construction
The group action of braids on tangles presented in this work was originally
discussed by Eliahou, Kauffman and Thistlethwaite [9] in the course of study
of the recently discovered links due to Thistlethwaite [33]. While it is still
unknown whether there is a non trivial knot having Jones polynomial V = 1,
Thistlethwaite’s examples allow us to answer the question for links having
more than 1 component.
Theorem (Thislethwaite). For n > 1 there are non trivial n-component
links having trivial Jones polynomial V = δ n−1 .

In the exploration of these links [9], it is shown that this is in fact a corollary
of a much stronger statement.
Theorem (Eliahou, Kauffman, Thislethwaite). For every n-component
link L there is an infinite family of (n + 1)-component links L 0 such that VL0 =
δVL .

While these assertions are discussed at length in [9], the goal of this chapter is
to present some of the examples in light of the group actions discussed in this
work.
Definition 6.1. A Thislethwaite link H(T, U ) is an external wiring of tangles
T, U ∈ S2 modeled on the Hopf link.

Our first task is to compute the evaluation matrix


 
hH(0, 0)i hH(0, ∞)i
H= .
hH(∞, 0)i hH(∞, ∞)i

62
Chapter 6. Thistlethwaite Links 63

Figure 6.1: The Thislethwaite link H(T, U )

It is easy to see, by applications of the Reidemeister move R 2 , that

hH(0, ∞)i = hH(∞, 0)i = δ 2

and
hH(∞, ∞)i = δ.

For the non trivial link H(0, 0), the computation of hH(0, 0)i requires a little
more work.

Figure 6.2: The link H(0, 0)

For this, the following switching formula (stated in [20]) will be useful.
Switching Formula. The equality
         
− = a4 − a−4 −
         
2 −2
+ a −a + − −

holds for the Kauffman bracket, giving rise to equivalence in TL 4 .


Chapter 6. Thistlethwaite Links 64

proof of the switching formula. First note that the double crossing
may be viewed as the braid σ2 σ1 σ3 σ2 ∈ B4 . Thus, as

σi 7−→ a + a−1 ei

gives a representation of the braid group in TLn , we can represent this element
of B4 as

(a2 + e1 + e2 + a−2 e2 e1 )(a2 + e2 + e3 + a−2 e3 e2 )


=a4 + a2 e2 + a2 e3 + e3 e2
+ a2 e1 + e1 e2 + e1 e3 + a−2 e1 e3 e2
+ a2 e2 + e22 + e2 e3 + a−2 e2 e3 e2
+ e2 e1 + a−2 e2 e1 e2 + a−2 e2 e1 e3 + a−4 e2 e1 e3 e2
=a4 + a2 e1 + (2a−2 + δ + 2a2 )e2 + a2 e3
+ a−2 e1 e3 e2 + a−2 e2 e1 e3 + a−4 e2 e1 e3 e2
+ e 1 e2 + e 1 e3 + e 2 e1 + e 2 e3 + e 3 e2 .

Similarly, the double crossing may be viewed as (σ2 σ1 σ3 σ2 )−1 ∈ B4 so


that
σi−1 7−→ a−1 + aei
gives the representation

(a−2 + e1 + e2 + a2 e2 e1 )(a−2 + e2 + e3 + a2 e3 e2 )
=a−4 + a−2 e1 + (2a−2 + δ + 2a2 )e2 + a−2 e3
+ a 2 e1 e3 e2 + a 2 e2 e1 e3 + a 4 e2 e1 e3 e2
+ e 1 e2 + e 1 e3 + e 2 e1 + e 2 e3 + e 3 e2 .
Chapter 6. Thistlethwaite Links 65

Therefore
   

=(a2 + e1 + e2 + a−2 e2 e1 )(a2 + e2 + e3 + a−2 e3 e2 )


− (a−2 + e1 + e2 + a2 e2 e1 )(a−2 + e2 + e3 + a2 e3 e2 )
=a4 − a−4 + (a2 − a−2 )e1 + (a2 − a−2 )e3
+ (a−2 − a2 )e1 e3 e2 + (a−2 − a2 )e2 e1 e3
+ (a−4 − a4 )e2 e1 e3 e2
=(a4 − a−4 )(1 − e2 e1 e3 e2 )
+ (a2 − a−2 )(e1 + e3 − e1 e3 e2 − e2 e1 e3 )
     
4 −4
= a −a −
         
2 −2
+ a −a + − −

as required.

Applying the switching formula twice,

* +
     
3 4 −4
=δ + a −a −

= δ 3 + (a4 − a−4 )
 
−(a4 − a−4 )(δ − δ 3 ) − (a2 − a−2 )(2 − 2δ 2 )
 
= δ 3 + (a4 − a−4 )(δ 2 − 1) δ(a4 − a−4 ) + 2(a2 − a−2 )
 
= δ 3 + (a4 − a−4 )(δ 2 − 1) a−6 − a−2 + a2 − a6

= −a−14 − a−6 − 2a−2 − 2a2 − a6 − a14


so the evaluation matrix is
 −14 
−a − a−6 − 2a−2 − 2a2 − a6 − a14 δ 2
H= .
δ2 δ

Now consider the braid ω = σ22 σ1−1 σ22


Chapter 6. Thistlethwaite Links 66

Figure 6.3: The braid ω ∈ B3

giving rise to the matrices


 −1 
−a + a3 − a7 −a−11 + 2a−7 − 2a−3 + 2a − a5
Φ(ω) =
a−3 a−13 − a−9 + a−5
 −7 
−a + a−3 − a −a−5 + 2a−1 − 2a3 + 2a7 − a11
Φ(ω −1 ) = .
a3 a5 − a9 + a13
It can be checked (using MAPLE, for example) that
Φ(ω) H Φ(ω −1 )> = H
so that H ∈ Fix(ω), giving rise to a comutative diagram
S2 × S2P
PPP
PPP
P((
ω Z[a, a−1 ]
n66
nnnnn
 nn
S2 × S 2
and
−1
hH(T, U )i = hH(T ω , U ω )i.
As a first example, consider the tangles

T = and U =

forming an unlink H(T, U ). Under the action of ω, we have

ω //

ω −1 //
Chapter 6. Thistlethwaite Links 67

−1
Therefore, the link H(T ω , U ω ) is two linked trefoils and, depending on ori-
−1
entation, w(H(T ω , U ω )) = ±8

−1
Figure 6.4: The link H(T ω , U ω )

so that
−1
hH(T ω , U ω )i = δ
and
VH(T ω ,U ω−1 ) = −a±24 δ.

However, the action of ω 2 leaves the writhe unchanged. This gives rise to a
family of 2-component Thisltlethwaite links, all having Jones polynomial δ.
Taking tangles T, U as in the previous example, the links
2n −2n
H(T ω , U ω )

have Jones polynomial δ for all n ∈ Z. Moreover, for n 6= 0 the links obtained
are non-trivial, since each component is the numerator closure of a tangle,
giving rise to a pair of 2-bridge links that are geometrically essential to a pair
of linked solid tori [32].
The first two links in this sequence (for n = 1, 2) are shown in figures 6.5 and
6.6.
Chapter 6. Thistlethwaite Links 68

Figure 6.5: The result of ω 2 acting on H(T, U )

Figure 6.6: The result of ω 4 acting on H(T, U )


Chapter 6. Thistlethwaite Links 69

6.2 Some 2-component examples


Thistlethwaite’s original discovery [33] consisted of links that had fewer cross-
ings than those of the infinite sequence constructed above. Starting with the
pair of tangles  

(T, U ) =  , 

we obtain a trivial link H(T, U ) such that w(H(T, U )) = −3. Applying the
action of ω to this link gives rise to a non-trivial link

Figure 6.7: A non-trivial, 2-component link

such that
−1
hH(T, U )i = hH(T ω , U ω )i
and  
−1
w H(T ω , U ω ) = −3.
The result is a non-trivial link with trivial Jones polynomial δ.
Similarly, starting with the pair of tangles
 

(T, U ) =  , 

gives rise to another trivial link H(T, U ), in this case having w(H(T, U )) = −1.
Applying the action of ω to this link gives rise to a non-trivial link
such that
−1
hH(T, U )i = hH(T ω , U ω )i
and  
−1
w H(T ω , U ω ) = −1.
Chapter 6. Thistlethwaite Links 70

Figure 6.8: A non-trivial, 2-component link

Again, the result is a non-trivial link with trivial Jones polynomial δ.


It has been shown that these examples are also members of an infinite family
of distinct 2-component links having trivial Jones polynomials [9].

6.3 A 3-component example


It is possible to construct a 16-crossings non-trivial link with trivial Jones
polynomial if we consider links of 3 components.
Starting with the pair of tangles
 

(T, U ) =  , ,

gives a 3 component trivial link H(T, U ). In this case, w(H(T, U )) = −2 and


applying the action of ω, the orientation of the resulting link may be chosen
so that  
−1
w H(T ω , U ω ) = −2
also. Thus, with this orientation,
VH(T ω ,U ω−1 ) = δ 2 .
In fact, with orientations chosen appropriately, this choice of tangles produces
another infinite family of links
n −n
H(T ω , U ω )
for n ∈ Z, each having trivial Jones polynomial [9].
The 16-crossing example is interesting, as it is may be constructed by linking
two simple links: the Whitehead link (5 21 ), and the trefoil knot (31 ).
Chapter 6. Thistlethwaite Links 71

Figure 6.9: A non-trivial, 3-component link

6.4 Closing Remarks


While the search for an answer to question 3.2 continues, the method of muta-
tion developed in this work provides a new tool in the pursuit of an example of
a non-trivial knot having trivial Jones polynomial. Not only has this type of
mutation produced Thistlethwaite’s examples, it is also able to produce pairs
of distinct knots sharing a common Jones polynomial that are not related by
Conway mutation (theorem 5.3). In light of the fact that Conway mutation
cannot alter an unknot so that it is knotted, it is desirable to have more general
forms of mutation such as this braid action at our disposal.
We have produced pairs of knots sharing a common Jones polynomial. As these
examples can be distinguished by their HOMFLY polynomials, they cannot be
Conway mutants. In our development, it is shown that further such examples
may be obtained either by altering the choice of tangles made, or by forming
a special closure |β| of a Kanenobu braid β ∈ B 6 . In addition, it is shown that
such a β may be produced from any given 3-braid.
It is hoped that further study of this new form of mutation will lead to a better
understanding of the phenomenon of distinct knots sharing a common Jones
polynomial. As well, it is possible that a better geometric understanding of this
braid action could give rise to a better understanding of the Jones polynomial
itself.
Bibliography
[1] J. W. Alexander. A lemma on systems of knotted curves. Proc. Nat.
Acad. Sci. USA, 9:93–95, 1923.

[2] J. W. Alexander. Topological invariants of knots and links. Trans. AMS,


30(2):275–306, 1928.

[3] Emil Artin. Theorie der Zöpfe. Abh. Math. Sem. Univ. Hamburg, 4:47–72,
1925.

[4] Emil Artin. Theory of braids. Ann. of Math., 48:101–126, 1947.

[5] Dror Bar-Natan. The knot atlas. www.math.toronto.edu/ ∼ drorbn.

[6] Joan S. Birman. Braids, Links and Mapping Class Groups. Number 82
in Annals of Mathematics Studies. Princeton University Press, Princeton,
1974.

[7] W. Burau. Über Zöpfgruppen und gleichsinnig verdrillte Verkettunger.


Abh. Math. Sem. Hanischen Univ., 11:171–178, 1936.

[8] J. H. Conway. An enumeration of knots and links. In J. Leech, editor,


Computational problems in abstract algebra, pages 329–358. Pergamon
Press, 1970.

[9] Shalom Eliahou, Louis Kauffman, and Morwen Thistlethwaite. Infinite


families of links with trivial Jones polynomial. preprint.

[10] Roger Fenn. An elementary introduction to the theory of braids. Notes


by Bernd Gemein.

[11] Roger Fenn. Hecke algebras, 2004. Notes from ‘Knots in Vancouver’.

[12] P. Freyd, D. Yetter, J. Hoste, W. Lickorish, K. Millet, and A. Ocneanu.


A new polynomial invariant of knots and links. Bull. AMS, 12:183–312,
1985.

[13] B. Hartley and T. O. Hawkes. Rings, Modules and Linear Algebra. Chap-
man and Hall Mathematics Series. Chapman and Hall, 1971.

72
Bibliography 73

[14] Jim Hoste and Jósef H. Przytycki. Tangle surgeries which preserve Jones-
type polynomials. Int. J. of Math., 8(8):1015–1027, 1997.

[15] Vaughan F. R. Jones. A polynomial invariant of knots via von Neumann


algebras. Bul. Amer. Math. Soc., 12:103–111, 1985.

[16] Vaughan F. R. Jones. Hecke algebra representations of braid groups and


link polynomials. Ann. of Math., 126:335–388, 1987.

[17] Taizo Kanenobu. Module d’Alexandre des noeuds fibré et polynôme de


Hosokawa des lacements fibré. Math. Sem. Notes Kobe Univ., 9:75–84,
1981.

[18] Taizo Kanenobu. Infinitely many knots with the same polynomial invari-
ant. Proc. AMS, 97(1):158–162, 1986.

[19] Louis Kauffman. State models and the Jones polynomial. Topology,
26:395–407, 1987.

[20] Louis Kauffman and Sóstenes Lins. Temperley-Lieb recoupling theory and
invariants of 3-manifolds. Number 134 in Annals of Mathematics Studies.
Princeton University Press, Princeton, 1994.

[21] W. B. R. Lickorish. Three-manifolds and the Temperly-Lieb algebra.


Math. Ann., 290:657–670, 1991.

[22] W. B. R. Lickorish. Calculations with the Temperly-Lieb algebra. Com-


ment. Math. Helvetici., 67:571–591, 1992.

[23] W. B. Raymond Lickorish. An Intoduction to Knot Theory. Number 175


in Graduate Texts in Mathematics. Springer, 1991.

[24] A. A. Markov. Über die freie Äquivalenz geschlossener Zöpfe. Rec. Math.
Moscou, 1:73–778, 1935.

[25] John W. Milnor. Infinite cyclic coverings. In John G. Hocking, editor,


Conference on the Topology of Manifolds, volume 13 of Complementary
series in mathematics, pages 115–133. Prindle, Weber and Schmidt, 1968.

[26] H. R. Morton. Treading knot diagrams. Mth. Proc. Camb. Phil. Soc.,
99:247–260, 1986.
Bibliography 74

[27] Kunio Murasugi and Bohdan I. Kurpita. A Study of Braids. Mathematics


and Its Applications. Kluwer Academic Publishers, 1999.

[28] K. Reidemeister. Elementare Begründung der Knotentheorie. Abh. Math.


Sem. Univ. Hamburg, 5:24–32, 1926.

[29] K. Reidemeister. Knotentheorie. Springer-Verlag, 1948.

[30] Dale Rolfsen. The quest for a knot with trivial Jones polynomial: Dia-
gram surgery and the Temperley-Lieb algebra. In M.E. Bozhüyük, editor,
Topics in Knot Theory, volume 399 of Nato ASI Series C, pages 195–210.
Kluwer Academic Publishers, 1993.

[31] Dale Rolfsen. Gobal mutation of knots. J. of Knot Theory, 3(3):407–417,


1994.

[32] Dale Rolfsen. Knots and Links. AMS Chelsea, Providence, third edition,
2003.

[33] Morwen Thistlethwaite. Links with trivial Jones polynomial. preprint.

[34] Pierre Vogel. Representations of links by braids: a new algorithm. Com-


ment. Math. Helvetici, 65:104–113, 1990.

You might also like