0% found this document useful (0 votes)
5 views20 pages

Introduction To Topological Quantum Computation With Anyons: Abstract

This document introduces topological quantum computing (TQC), a theoretical model where information is encoded in the braiding of particles, providing protection against local errors and decoherence. It discusses the basic principles of quantum computation, the role of anyons, and outlines the mathematical framework and examples of TQC models. The paper aims to provide a simplified overview of TQC for readers with minimal prior knowledge of physics or quantum computing.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
5 views20 pages

Introduction To Topological Quantum Computation With Anyons: Abstract

This document introduces topological quantum computing (TQC), a theoretical model where information is encoded in the braiding of particles, providing protection against local errors and decoherence. It discusses the basic principles of quantum computation, the role of anyons, and outlines the mathematical framework and examples of TQC models. The paper aims to provide a simplified overview of TQC for readers with minimal prior knowledge of physics or quantum computing.
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 20

INTRODUCTION TO TOPOLOGICAL QUANTUM

COMPUTATION WITH ANYONS

CHIA-HSUN (PAUL) LEE

Abstract. Topological quantum computing (TQC) is a theoretical model for


quantum computation. In this model, information is encoded in the braiding
of particles, therefore protecting it from local errors. TQC is a promising
solution to the problem of decoherence in quantum computing. In this paper,
we introduce the abstract model of topological quantum computation and look
at two basic examples of TQC models.

Contents
1. Introduction 1
2. Minimal Introduction to Quantum Computation 3
2.1. The Qubit 4
2.2. Quantum Operations 4
2.3. Quantum Measurement 5
2.4. Summary 6
3. Ribbon Unitary Fusion Category 6
3.1. Quantum Computation with the Ribbon Unitary Fusion Category 12
4. Braided 6j Fusion Systems 13
4.1. Ising Fusion System 15
4.2. Fibonacci Fusion System 15
4.3. Simulating a Traditional Quantum Computer 16
5. Analog Computation of the Jones Polynomial 16
6. Conclusion 17
Acknowledgements 17
References 17

1. Introduction
One of the main challenges of quantum computation is mitigating errors. A lot
of effort has been put in to developing robust quantum error correcting codes with
minimal overhead. One class of error correcting codes, called topological codes,
utilizes topological properties to encode information in order to prevent local errors
from changing the logical state of a quantum system. Such codes have proven to
be quite resilient against errors but often require large amounts of qubits to encode
one single logical qubit. Topological quantum computing (TQC) is the application
of the same principle to a physical system.

Date: August 25, 2018.


1
2 CHIA-HSUN (PAUL) LEE

Figure 1. The path of a particle travelling around another in 3


or more dimensions is nullhomotopic.

Information in TQC is encoded in the braiding of the worldlines of particles


under exchange. The state of particles is encoded by a wave function, and the
act of exchanging particles acts on this wave function. The effect on the wave
function under exchange is called exchange statistics, or just statistics. The topo-
logical nature of a particle system is given by the Aharonov-Bohm effect: a particle
travelling along a path around a zero magnetic field gains a phase factor given
by the flux of the enclosed magnetic field. In a vacuum with a few particles, the
physical significance of the path a particle takes is only defined up to homotopy.
Thus local disturbances in a quantum system will not affect the overall quantum
state resulting in a topologically protected system. In three or more dimensions,
braids in Bn that get sent to the same permutation in Sn , by sending a braid to
the permutation of its end points, turn out to be homotopic. This means, that are
no non-trivial particle exchanges, shown in Figure 1. Thus, in such a situation, it
is the group Sn that is acting on the wave function, resulting in only two particle
types, fermions and bosons. In two dimensions, the full braid group Bn acts on
the wave function. This allows there to be infinitely many different particles that
have arbitrary exchange statistics, called anyons. Topological quantum computing
is based on this property that there are infinitely many different particle types.
While there is strong experimental evidence that anyons do exist, we have yet to
produce any. A slightly more detailed discussion of the mathematical justification
for the existence of anyons can be found in my previous REU paper1.

In principle, particles with different exchange statistics are distinguishable and


particles can be split or fused together to form other particles. Further, a pair of
particles may have multiple possible fusion outcomes. When necessary, we will call
particles that exhibit different exchange statistics as having distinct “topological
charges.” As the name suggests, the topological charge of a system is a conserved
quantity. That is, if we bring all the particles together in a system, no matter what
we do, we will always end up with a particle of the same type. As stated in the
Aharonov-Bohm effect, in two dimensions, particle exchanges act non-trivially on
the wave function of a system. This turns out to influence the probabilities of the
different possible outcomes of fusion. We utilize this to do quantum computing: we
begin by preparing a set of particles in a known state; we act on it by exchanging
1https://math.uchicago.edu/ may/REU2016/REUPapers/Lee.pdf
~
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 3

particles around; in the end we observe the outcomes of fusing particles together.
Thus, the information of a system is encoded precisely by the world lines of the
particles under exchange. So long as we keep the particles sufficently far apart at
all times, the topological nature of the system means that small local errors will
not change the information encoded in the system. For the encoded information to
change, a large enough error must occur such that the the topological property of
a braid is altered.
In this paper, we will be looking at a simplified overview of the model of topo-
logical quantum computing. Enough to understand the abstract model and gain
working knowledge of how quantum computation is carried out in TQC. The field
is at the intersection of many fields in computer science, mathematics and physics
and this exposition by no means does it justice. Many details have been left out to
keep this at a reasonable length.
In principle, no prior knowledge of physics or quantum computing is required.
However, some background may be useful in orienting the reader in the subject.
We will begin with a very brief introduction to the basics of quantum computing.

2. Minimal Introduction to Quantum Computation


Before we describe the computational power of anyons, it would be useful to
familiarize ourselves with the basic concepts and terminology used in quantum
computing. In this section, we will introduce the bare minimum of quantum com-
puting needed in order to understand quantum computation with anyons. A reader
already familiar with quantum computation may safely skip this section. A more
detailed description of quantum computation may be found in [7], [1], and [5]. Our
formalization of quantum computing will closely follow the circuit model of quan-
tum computation. A Quantum Turing Machine formalization exists, but we will
not be discussing that in this section. We refer the reader to the references above
for details.
We will assume that the reader is familiar with linear algebra and tensor prod-
ucts. References for those topics can be found at [10]. A brief description of Dirac’s
bra-ket notation will follow now. Dirac’s notation is standard among quantum
physics and quantum computing.
Notation 2.1. A “ket” written |ψi, is just a vector in a Hilbert space Cn . A “bra”

is the conjugate transpose of a “ket,” that is, hψ| = |ψi . Lastly, |ai ⊗ |bi is usually
written as |ai |bi or just |abi. To avoid confusion, we will refrain from writing hab|
or the likes. We can apply a “ket” to a linear operator, just as you would with a
regular vector. For example, A |ψi. We can also apply it to a “bra,” resulting in
another “bra” like so: hψ| A = (A |ψi)† . Familiarize yourself with this notation by
considering the following objects:
• A |ψi where A : Cn → Cm is a linear map.
• hψ|ϕi
• hψ|B|ψi where B : Cn → Cn is an endomorphism.
• |ψi hψ|
Let’s recall the circuit model of classical computation. The basic building block
of classical computation is the bit. A bit can exist in two states, on or off. Com-
putation is carried out by preparing a collection of bits initialized to a certain
state in {0, 1}n , manipulated with a sequence of logic gates or boolean functions,
4 CHIA-HSUN (PAUL) LEE

such as AND, OR, NOT, and measured at the end to read the output of compu-
tation. Quantum computing operates similarly, consisting of initialization of an
initial state, manipulation of that state by a sequence of gates, and a measurement
of the final state. The key difference is, instead of bits, we have qubits.

2.1. The Qubit. The state of a qubit is a unit vector, denoted |ψi, residing in
a Hilbert space C2 . Thus, there are continuum many possible states for a single
qubit, as opposed to just 2 classically. We choose a set of orthonormal basis vectors
|0i , |1i corresponding to the off and on states of a classical bit. We call this the
computational basis, or basic states, and all matrices written below are written in
this basis. In general, the state of a qubit may be written as,
|ψi = α |0i + β |1i
where |α|2 +|β|2 = 1. Adding more qubits corresponds to tensoring the state spaces
of multiple qubits together. This is expected as, in quantum mechanics, the Hilbert
space of the composition of two subsystems is given by the tensor product of the
Hilbert spaces of the two subsystems. For example, the state of a 2-qubit system is
described by a unit vector residing in C2 ⊗ C2 . As we fix a computational basis for
each of the components, the Hilbert space will have the basis |00i , |01i , |10i , |11i.
Hence, a state of a 2-qubit system is written as,
|ψi = α |00i + β |01i + γ |10i + δ |11i
where |α| + |β| + |γ| + |δ|2 = 1. In general, the state space of an n-qubit system
2 2 2

is (C2 )⊗n = C2 ⊗ · · · ⊗ C2 with basis |xi where x ∈ {0, 1}n and a state is written
as, X
|ψi = αx |xi
x∈{0,1}n
where, X
|αx |2 = 1
x

2.2. Quantum Operations. Now we’ve described the state of a collection of


qubits, we need a way to manipulate it. Since any operation must take a state
of qubits to another state of qubits and be norm preserving, the operation must be
a unitary map from (C2 )⊗n to itself. As is with the case of classical computation,
where gates are boolean functions, we would like to be able to build complex gates
from a small set of simple gates, in order to minimize the cost of building a quantum
computer. In the classical setting, one single gate is enough, the NAND gate, to
build all possible boolean functions from n-bits to m-bits. In the case of quantum
computation, it can be shown that a finite set of unitary maps is not enough to
generate all unitary maps on n-qubits through a simple cardinality argument. The
set of all unitary maps generated from a finite set of maps is countable, while the
set of all unitary maps on n-qubits is uncountable. Hence the result. However, two
important theorems say that we can do so if we relax certain conditions:
Theorem 2.2 ([1]). Any unitary map can be decomposed into single qubit gates
and the CNOT gate given by,
CNOT : |x1 i |x2 i 7→ |x1 i |x1 ⊕ x2 i
where ⊕ is addition modulo 2 and x1 , x2 ∈ {0, 1}.
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 5

Theorem 2.3 ([7]). We say S is a set of universal gates if any unitary map can be
approximated up to  error in the operator norm, for any  > 0, in polylogarithmic
in 1/ many gates from S. The following forms a set of universal gates,
 
1 1
(1) The Hadamard gate, H = √12
  1 −1
1 0
(2) K =
0 i
(3) K −1
(4) CNOT
(5) The Toffoli gate, also known as the controlled CNOT, given by,
TOFFOLI : |xyi |zi 7→ |xyi |(x × y) ⊕ zi
where x, y, z ∈ {0, 1} and × the usual multiplication. That is, flip the bit z
if x and y are both 1, otherwise do nothing.
Now that we understand how quantum gates look, we need to address an impor-
tant issue. All quantum gates are invertible, while classical gates are not necessarily
invertible. It is useful, and necessary, to be able to implement classical gates in the
quantum computation model. Fortunately, the following theorem allows us to do
so,
Theorem 2.4 (Garbage Removal Lemma, [7]). For any classical gate f : 2n → 2m
computable by a circuit of size L, there is a permutation f⊕ on L + m + n bits com-
putable with a circuit of size 2L+m such that, f⊕ (x, 0, . . . , 0) = f⊕ (x, f (x), 0, . . . , 0)
where x ∈ 2n .
The proof is simple and can be found in [7]. Usually, a classical boolean function
f will be realized by the operator,
Uf : |xi |yi |0L i 7→ |xi |f (x) ⊕ yi |0L i
by lifting the classical circuit implementation of f⊕ directly to quantum gates.

2.3. Quantum Measurement. At this point you should be able to see that a
single qubit holds vast amounts of information. You could even say that it holds
infinite information, as the state of a single qubit generally requires infinitely many
classical bits to describe. However, there is a catch, every qubit we prepare only
reveals a single bit of information everytime we try to “observe” it. The limitation
is given in the following postulate of quantum mechanics:
Postulate 2.5. Any physical observable is associated with a self-adjoint operator
A, and the possible outcome of the measurement of an observable A is one of its
eigenvalues. Further, if we write ai as the eigenvalue for the eigenvector |ii, then
in this basis, a state |ψi is represented by,
X X
|ψi = αi |ii , where |αi |2 = 1
i i

Then, the probability that a measurement yields the outcome ai is given by |αi |2 .
Additionally, if the measurement yields the eigenvalue ai , the state is projected on
to the eigenvector |ii. That is, subsequent measurements with the same operator
will yield the same results.
6 CHIA-HSUN (PAUL) LEE

In the case of a qubit, we would like to observe it in its computational basis.


This corresponds to the following operator,
 
1 0
σz =
0 −1
with eigenvalues 1 corresponding to |0i and −1 corresponding to |1i and the ex-
pected value of the measurement of a state |ψi is given by hψ|σz |ψi. Hence, with
every single qubit you prepare, you can only ever get one single bit of information
out of it. To make matters worse, the no-cloning theorem [1] says that the map
∆ : H ⊗ H → H ⊗ H that behaves by |ψi |ϕi 7→ |ψi |ψi on any states |ψi , |ψi is not
realizable by a unitary map. Thus, in order to prepare multiple identical qubits,
one must start from the beginning from a known state that we can reproduce.
For simplicity sake, if we know that a vector |ϕi is observable, we will just write
hϕ|ψi to denote the probability of observing the state |ψi in the state |ϕi without
going through the quantum measurement machinery.
Measurement of multiple qubits at the same time is possible. Detailed discussion
of the measurement of multiple qubits is given in [4]. There are subtle differences
between measuring a single qubit individually or a bunch at once. Details may be
found in the references for quantum computing listed before.
Additionally, we should mention that a global phase factor on a quantum state
has no physical significance, i.e., the two states |ψi , eiϕ |ψi are physically identical.
Hence, we usually factor out global phase by rotating the system so that the |0i
vector is on the positive real line if possible. That is, the state of a single qubit up
to global phase factor, is written as,
|ψi = cos(θ/2) |0i + eiϕ sin(θ/2) |1i
where θ ∈ [0, π] and ϕ ∈ [0, 2π). This representation is not too important, we
usually opt for the more simple representation given earlier, as no measument is
able to distinguish a global phase anyway. However, this provides a method of
visualizing a single qubit in three dimensions, called the Bloch sphere, given by
considering θ, ϕ as the polar coordinates of the state on the unit sphere. We will
not be going through it in this paper but I encourage the reader to look it up [1] if
they have trouble thinking about qubits.
2.4. Summary. In summary, quantum computation consists of the following steps,
(1) The preparation of n-qubits in a certain state |ψi ∈ (C2 )⊗n .
(2) Application of a sequence of unitary operations, |ψi 7→ |ϕi = Un · · · U0 |ψi.
(3) Measurement of qubits of the final state ϕ.

3. Ribbon Unitary Fusion Category


Quantum computation is more powerful than classical computation. It can
speedup many problems that are hard classically. Typical examples of quantum
speedup are the unstructured database search problem and the integer factoriza-
tion problem. However, quantum computation as it stands has its limitations.
Keeping a quantum system intact remains a huge technological hurdle. Current
quantum error correcting methods add a huge amount of overhead to algorithms
that make implementing them not feasible or barely advantageous. Topological
quantum computing promises to fix this by encoding information topologically,
therefore protecting information from being destroyed by local errors.
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 7

Recall that topological quantum computing with anyons is facilitated by braiding


particles. Particles can be fused together to form another particle and braiding
changes the probabilities of the outcomes of fusion. The act of fusing the particles
together and observing the outcomes is the measurement of quantum state. The
abstract model of topological quantum computation is captured in the unitary
modular tensor category. The structure of it is depicted roughly in Figure 2. For
the purposes of a simple introduction, we will relax some constraints and look just
at ribbon unitary fusion categories. It turns out that this category already holds a
huge amount of data.

Fusion Category

+ pivotal structure

Pivotal Fusion Category

+ compatible braiding

Ribbon Fusion Category + unitary structure

Ribbon Unitary Fusion Category


+ non-degeneracy

Modular Tensor Category

+ unitary structure

Unitary Modular Tensor Category


Figure 2. The structure of a unitary modular tensor category (UMTC)

We begin by looking at fusion categories. We start with a few basic definitions.

Definition 3.1 (Tensor Category). A category C is a tensor category if,

(1) there is a bifunctor ⊗ : C × C → C,


(2) there is a unit object 1,
(3) for every object x, y, z there is a natural isomorphism αx,y,z : (x ⊗ y) ⊗ z ∼
=
x ⊗ (y ⊗ z), that satisfies the pentagon equations,
8 CHIA-HSUN (PAUL) LEE

(a ⊗ b) ⊗ (c ⊗ d)
αa,
b,c⊗
d

a ⊗ (b ⊗ (c ⊗ d))

,d
b,c

((a ⊗ b) ⊗ c) ⊗ dαa ida ⊗αb,c,d
αa
,b,
c

a ⊗ ((b ⊗ c) ⊗ d)
id d

c,d
α a,b⊗
(a ⊗ (b ⊗ c)) ⊗ d

(4) for every object x there are natural isomorphisms ρx : 1 ⊗ x ∼


= x, λx :
x⊗1∼ = x such that the triangle diagram commutes,

x⊗y λx ⊗idy
(x ⊗ 1) ⊗ y y⊗x idy ⊗ρx
y ⊗ (1 ⊗ x)

idx ⊗ρy αx,1,y λy ⊗idx αy,1,x

x ⊗ (1 ⊗ y) (y ⊗ 1) ⊗ x

For convenience, we will require all the above isomorphism to be identities, so


that we can write a ⊗ b ⊗ c without ambiguity. In principle, it is not required for
topological quantum computing. This gives us the following definition,

Definition 3.2 (Strict Tensor Category). A tensor category C is a strict tensor


category if, the natural isomorphisms αx,y,z , λx , ρx are all identities.

Let us now add structure that captures the notion of fusion.

Definition 3.3 (Strict Fusion Category). A category C is a strict fusion category


(over C) if,

(1) it is C-linear, i.e. every Hom-set is a C-vector space,


(2) it is a strict tensor category such that ⊗ is bilinear on morphisms,
(3) the unit object 1 is simple, i.e. Hom(1, 1) ∼ = C,
(4) there are finitely many isomorphism classes of simple objects,
(5) it is semisimple, i.e. every object is a finite direct sum of simple objects,
(6) every object has left and right duals, i.e. for every object x, there are objects
x∗ , ∗ x along with morphisms, ηx : 1 → x⊗x∗ , x : x∗ ⊗x → 1 (right rigidity)
and ηx0 : 1 → ∗ x ⊗ x, 0x : x ⊗ ∗ x → 1 (left rigidity) such that the following
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 9

are identities,
ηx ⊗idx id ⊗
x x ⊗ x∗ ⊗ x x

idx∗ ⊗ηx x ⊗idx∗


x∗ x∗ ⊗ x ⊗ x∗ x∗

0
idx ⊗ηx 0x ⊗id
x x ⊗ ∗x ⊗ x x

0
∗ ηx ⊗id∗ x ∗ id∗ x ⊗0x
x x ⊗ x ⊗ ∗x ∗
x
Let us examine the structure of this category and how it relates to topological
quantum computing. In the category, a particle is represented by a simple object
and a type of particle is represented by an isomorphism class of simple objects.
Suppose we have two simple objects, or particles, a and b. If there is a morphism
a ⊗ b → c, where c is another simple object, we say a, b may fuse to c. If there is
a morphism c → a ⊗ b, we say c may split to a and b. Suppose a ⊗ b = c ⊕ d ⊕
e, where c, d, e are simple objects due to semisimplicity, then it means there are
morphisms a ⊗ b to c, d, e respectively. Alternatively, we could say c, d, e are the
possible outcomes of fusing particles a and b. As one would expect of fusion, the
outcome of fusion should be independent of the order we fuse the particles, hence,
⊗ is associative. Semisimplicity also says that the possible outcomes of fusion is
finite. Further, rigidity says that there exist particle/antiparticle pairs. The term
particle/antiparticle pair implies some sort of isomorphism between a particle and
its double dual. Indeed, they are automatically isomorphic [3]. However, this
isomorphism is not necessarily natural. In order to get the full power of quantum
computing, we require that there be a nice isomorphism between objects and their
double duals. This is called a pivotal structure and it is essential for more advanced
operations (quantum trace). We will define it below after we look at the properties
of the fusion category. The following is an important consequence of semisimplicity.
Proposition 3.4. Every Hom-set in a fusion category is finite dimensional.
Proof. This is a direct consequence of (1), (3), and (4) in the definition of a fusion
category. We have Hom(x, x) ∼ = C if, and only if, x is simple. If y is not simple, it is
a direct sum of finitely many simple objects, y = a1 ⊕ · · · ⊕ an . So, Hom(x, y) ∼
= Ck
is finite dimensional where k is exactly the number of i such that x ∼ = ai . 
This result justifies the later use of intermediate fusion outcomes to label basis
vectors for Hom(x, y). For instance, suppose that y ⊗ z can fuse to i, j and both
can fuse with x to 1. Then x ⊗ (y ⊗ z) = x ⊗ (i ⊕ j ⊕ · · · ) = 1i ⊕ 1j ⊕ · · · where
1i = 1j = 1 and the labels are to identify its source. In particular,
Hom(1, x ⊗ (y ⊗ z)) ∼ = Hom(1, 1i ) ⊕ Hom(1, 1j ) ⊕ · · ·
So we have at least two basis vectors of Hom(1, x ⊗ (y ⊗ z)) distinguished by the
outcomes of the fusion of y ⊗ z. In particular, when we add a unitary structure,
these fusion trees are going to be labeling an orthonormal basis.
To ease talking about fusion categories, we employ diagrams. In our diagrams, we
following the convention in physics that time flows upwards. We denote x1 ⊗· · ·⊗xn
as points on a line. For instance, the following is the diagram for x1 ⊗ · · · ⊗ x5 :
10 CHIA-HSUN (PAUL) LEE

x1 x2 x3 x4 x5

As mentioned above, we identify vectors in a basis given by some order of fusion


with the intermediate fusion outcomes. Hence, we use fusion trees to denote basis
vectors. The following denotes a basis vector in Hom(1, (a ⊗ b) ⊗ c):

x1 x2 x3

We denote morphisms with just boxes. For example, the diagram for f : x → y is
given by,
y

A plain line denotes the identity morphism. We have diagrams for right birth and
death,
x x∗
x∗ x
To denote composition, we stack diagrams together from bottom to top consistent
with the convention that time flows upwards. In particular, we have intepretations
for the following, corresponding to one of the rigidity constraints.
x x

x x

Note, at the moment, there is no interpretation for the following diagram. Since
we have no diagram for left birth and death.

In order to define the above, we need a pivotal structure.

Definition 3.5 (Strict Spherical Fusion Category). A strict fusion category C is


a strict spherical fusion category if it has a nice pivotal structure. Make ∗ into a
contravariant functor by defining f ∗ : y ∗ → x∗ by the diagram,
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 11

x∗

f
y

A pivotal structure is a collection of isomorphisms, ϕx : x → x∗∗ such that,


(1) for all objects x, y, ϕx⊗y = ϕx ⊗ ϕy
(2) for all morphisms f , f ∗∗ = f
Additionally, for all morphisms f : x → x the following two diagrams, called the
trace, must agree.

ϕx f

f ϕ−1
x

For convenience, we will denote the above by just drawing the box for f with the
understanding that the ϕx is inserted appropriately.

With this we can interpret the circle as the trace of the identity. Let us now add
structure for braiding.

Definition 3.6 (Strict Braided Fusion Category). A strict fusion category C is


braided if for every pair of objects x, y, there is a natural isomorphism cx,y : x⊗y ∼
=
y ⊗ x such that the hexagon diagram below commutes,
αb,a,c
(b ⊗ a) ⊗ c b ⊗ (a ⊗ c)
c a,c
,b
ca

(a ⊗ b) ⊗ c b ⊗ (c ⊗ a)
α a,

a
,c,
b,c

αb

ca,b⊗c
a ⊗ (b ⊗ c) (b ⊗ c) ⊗ a

We denote braiding x ⊗ y with cx,y by the following diagram,


12 CHIA-HSUN (PAUL) LEE

Note, the braid pictured is in fact a right handed braid, where the strand on the
right crosses over the strand on the left. Since time is flowing upwards this may
be contrary to other diagrams where braids grow downwards. Arrows have been
added to this particular diagram as a reminder of that fact.
Note, the definition differs from that of a symmetric tensor category since cy,x cx,y
is not necessarily the identity. Note that braiding gives another isomorphism ψx :
x∗∗ → x between an object and its double dual, given by the diagram,
x

x∗∗
Again this isomorphism need not be natural nor does it need to be compatible with
the pivotal structure, if it exists, in general. If a spherical fusion category has a
compatible braiding then it is called a ribbon fusion category.

Definition 3.7 (Strict Ribbon Fusion Category). A braided spherical fusion cat-
egory is a ribbon fusion category if for all objects x, θx = ψx ϕx : x → x satisfies
θx∗ = θx∗ .

Lastly, we add a unitary structure in order to perform some basic quantum


computation.

Definition 3.8 (Strict Ribbon Unitary Fusion Category, [9]). A strict braided
fusion category C is a strict braided unitary fusion category if it has the additional
structure,
(1) Hom-sets are Hilbert spaces,
(2) there is conjugation, i.e. a contravariant endofunctor − acting as the iden-
tity on objects and such that f = f , f ⊗ g = f ⊗ g, f g = gf for all
morphisms f, g,
(3) for all f , f f = 0 implies f = 0
(4) ηx = 0x and x = ηx0
(5) cx,y = c−1
x,y
(6) θx = θx−1
(7) for all f , the trace of f f is non-negative.

3.1. Quantum Computation with the Ribbon Unitary Fusion Category.


Now we briefly describe how to perform quantum computation in this category.
Recall that there are three steps to quantum computation. Initial state preparation,
unitary evolution of the state, final measurement. These actions correspond to the
following,
(1) To prepare the initial state, we pick a state vector |ψi in some Hom(b, a1 ⊗
· · · ⊗ an ).
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 13

(2) We operate on this initial state by picking a sequence of unitary morphisms


Hom(a1 ⊗ · · · ⊗ an , aσ(1) ⊗ · · · ⊗ aσ(n) ) and composing it with |ψi to get a
new state |ϕi ∈ Hom(b, aσ(1) ⊗ · · · ⊗ aσ(n) ).
(3) Measure |ϕi in Hom(b, aσ(1) ⊗ · · · ⊗ aσ(n) ).
In particular, the morphism in (2) is usually a sequence of braiding operators.
Hence, a quantum algorithm in this model is given by a braid and a fusion tree
(the choice of basis).
As you can see, all the components of quantum computation are there. However,
it’s not very helpful if you actually want to do quantum computation as there is a
lot of structure to specify. It turns out that the structure of a ribbon unitary fusion
category is determined (almost completely) by how different types of particles fuse
together. This allows us to specify a model of topological quantum computing with
far less data. We now turn our attention to that formalization.

4. Braided 6j Fusion Systems


In a fusion category, there are finitely many isomorphism classes, or particle
types. The structure of a braided fusion category turns out to be determined by
how these particle types fuse pairwise. To describe how particles of different types
fuse together, we write them down as fusion rules. The following formalization is
given in [9].
Definition 4.1 (Braided 6j Fusion System). A 6j Fusion System consists of the
following,
(1) A set of symbols L, whose elemnts we will call particle types, with a dis-
tinguished element 1 and an involution −∗ : L → L. We will call the
distinguished element the vacuum.
(2) A binary operation ⊗ : L×L → NL , called the fusion rule. For convenience,
c a b
we denote (a ⊗ b)(c) by Nab and write formally a ⊗ b = Nab a ⊕ Nab b ⊕ ···
do denote the function given by a ⊗ b.
(3) A total fusion channel t ∈ L.
satisfying the following,
(1) 1∗ = 1.
x d
P
For all a, b, c ∈ L, (a ⊗ b) ⊗ c = a ⊗ (b ⊗ c). In other words, x Nab
(2) P Nxc =
x d
N N
x bc ax for all d ∈ L.
(3) For every particle type, there is exactly one other particle type that corre-
sponds to the antiparticle. That is, for all a, b ∈ L,
(
1 a∗ = b
Nab = Nba =
0 otherwise
(4) The vacuum fuses trivially with all particle types. That is, for all a, c ∈ L,
(
c c 1 a=c
Na1 = N1a =
0 otherwise
Additionally, the structure needs to be compatible with that of a unitary braided
fusion category. Associated with every fusion, a1 ⊗ · · · ⊗ an → b is a Hilbert
space Cn where n is the number of “fusion paths” to b, i.e. the number of ways
to fuse to b. Given a fixed order of fusion the Hilbert space has an orthonormal
basis parameterized by the intermediate outcomes of fusion [9]. Further, there
14 CHIA-HSUN (PAUL) LEE

d
are unitary matrices Fabc , Rab , such that analogues of the triangle diagrams, the
pentagon diagrams, and the hexagon diagrams given in the section above commute.
These equations are given explicitly below.
We use fusion trees to denote vectors of a particular basis of our Hilbert space.
The F -matrix relates the two different choices of basis for a ⊗ b ⊗ c, it acts on the
basis vectors as follows:
a b c a b c
k

Fabc i,j
i j

k k

The R-matrix is the braiding matrix. It is always diagonal and acts on the basis
vectors as follows:
a b a b

(Rab )k,k

k k
d
In particular, the triangle equation is simple: Fa,b,c = 1 if any one of a, b, c is 1.
This is what we expect: suppose b = 1, then the intermediate outcome of fusing
a ⊗ 1 must be a and the intermediate outcome of fusing 1 ⊗ c must be c. Now
given the above diagrammatic representation, we are able to give the pentagon and
hexagon equations diagramatically, given in Figures 3 and 4.

It turns out specifying the fusion rule is almost enough to determine the structure
of a braided 6j fusion system.
Theorem 4.2 (Ocneaunu rigidity, [9]). Given a fixed fusion rule, there are only
finitely many braided fusion systems with that fusion rule.
Given a fusion rule, one can find a consistent fusion system by solving the
hexagon and pentagon equations. Generally, solving the pentagon and hexagon
equations given a fusion rule to find a consistent model is a computationally diffi-
cult task [9].
If a braided 6j fusion model gives a ribbon unitary fusion category, we can per-
form quantum computation with it. A fusion model gives a ribbon fusion category
under some conditions (see [9]). For simplicity, assume that the model we are
working with does in fact have the required structure. Quantum computation in a
braided 6j fusion model is carried out in the following steps,
(1) Prepare a set of particles in a known state |ψi. That is, we’ve chosen an
order of fusion, and we know what the state of the particle is in that basis.
INTRODUCTION TO TOPOLOGICAL QUANTUM COMPUTATION WITH ANYONS 15

(2) Operate on this initial state by braiding particles.


(3) Fuse the particles together in our chosen order of fusion and observe the
intermediate outcomes.
Let us consider two simple fusion rules and examine their computation power:
Example 4.3. When we write out fusion rules, we usually omit the trivial ones
and write out the non-trivial ones.
(1) Ising fusion rule: L = {1, σ, ψ} with,
σ⊗σ =1⊕ψ ψ⊗σ =σ⊗ψ =ψ ψ⊗ψ =1
(2) Fibonacci fusion rule: L = {1, τ } with,
τ ⊗τ =1⊕τ
4.1. Ising Fusion System. The Ising fusion rules are simple enough to be able
to compute the F and R-matrices by hand. For detailed computation see [6].
σ
In particular, there is only one non-trivial F -matrix: Fσσσ . This matrix is two
dimensional (σ ⊗ σ ⊗ σ = 2σ) and is,
 
σ 1 1 1
Fσσσ =√
2 1 −1
in the basis given by the fusion trees in the order σ, σ → 1 and σ, σ → ψ dis-
tinguished by the outcome of the first fusion. There is also only one non-trivial
R-matrix: Rσσ . This matrix is again two dimensional (σ ⊗ σ = 1 ⊕ ψ) and is given
by,
 
1 0
Rσσ = eiϕ
0 i
given in the same basis.
In practice, four σ particles are employed to represent a qubit. Fixing an order
of fusion, the state corresponding to the first pair of σ fusing to 1 is the logical |0L i
and the state corresponding to fusing to ψ is |1L i.

4.2. Fibonacci Fusion System. As for the Fibonacci Fusion System, there is
again only one non-trivial F -matrix: Fτττ τ since (τ ⊗ τ ⊗ τ = 1 ⊕ 2τ ). This matrix
is two dimensional and again ordered by intermediate outcomes 1, τ :
 −1 √ 
ϕ ϕ
Fτττ τ = √
ϕ −ϕ−1
There is also only one non-trivial R-matrix. It is two dimensional and given by,
 4πi/5 
e 0
Rτ τ =
0 −e2πi/5
Detailed computation can be found in [8]. Computation of the F and R matri-
ces for this model is simple and the reader can certainly try to derive it. Braids
approximating the Hadamard and CNOT gates can be found in [2].
In practice, three τ particles are grouped together to form a qubit. Fixing an
order of fusion, we use the same convention, the state corresponding to the first
pair fusing to 1 is |0L i and the state corresponding to fusing to τ is |1L i.
16 CHIA-HSUN (PAUL) LEE

4.3. Simulating a Traditional Quantum Computer. The question to ask is


whether or not a topological quantum computer is as powerful as a traditional
quantum computer. In order to simulate a traditional quantum computer, we need
to be able to perform any computation it can do. That is, we need to find a fusion
system with braid group representation ρ : Bn → SU(d) such that for every unitary
U we can find a braid b ∈ Bn such that the following commutes, up to arbitrary
precision,
i
(C2 )⊗n Vn
U ρ(b)

i
(C2 )⊗n Vn
where i is a chosen embedding of the logical n-qubit space in our fusion space Vn .
In particular, if the image of the braid group representation is dense, then the above
is achievable. A conjecture of a sufficient condition for universality can be found
in [9].
In particular, the Ising model does not support universal quantum computation
while the Fibonacci model does. However, one advantage the Ising model has is
that there is a simple braid for implementing the CNOT gate exactly. It turns
out that if we can perform single qubit rotations through external means, we get
quantum computation. Additionally, the Ising model is currently the one closest to
experimental realization.
On the other side of the question whether topological quantum computers are as
strong as traditional quantum computers, is the question whether they are stronger.
As it turns out, they are not, and a quantum computer can efficiently simulate a
topological quantum computer [9].

5. Analog Computation of the Jones Polynomial


The topological nature of TQC provides an analog algorithm for approximating
the Jones polynomial at some root of unity q. Suppose we have a link L given
as the trace closure of some braid B. Recall that the Jones representation ρA is
unitary for A = ±ie±2πi/4r [9] and that the Jones polynomial is given by,
VL (A) = (−A)3 Wr(L) dn−1 Tr(ρA (B))
where Wr is the writhe map and d = −A2 − A−2 . In particular, given an appro-
priate model of TQC with a braid group representation given by ρA (B), the trace
Tr(ρA (B)) is given by the diagram of the trace closure of the braid B. Specif-
ically, let |ψi denote the state corresponding to the pairwise creation of particle
and antiparicle pairs in Hom(1, x1 ⊗ · · · ⊗ x∗1 ). The diagram can be written as
hψ|ρA (B)|ψi ∈ Hom(1, 1). The quantum mechanical interpretation of this value is
the probability that pairwise fusion of the particles after evolution ρA (B) yields the
vacuum. Hence, we can additively approximate this value to compute the Jones
polynomial (see Figure 5 for an example).

The problem of additive approximation of the Jones polynomial of a link L at


e±2πi/r where r > 6 turns out to be BQP-complete [9], where BQP is considered to
be the feasible class of quantum computation. It is the class of languages that can
be decided by a polynomial sized quantum circuit with success probability at least
REFERENCES 17

2/3. It is the quantum analog of BPP. For more information on the complexity of
those classes see [7].

6. Conclusion
TQC is a topologically protected model of quantum computation. Although not
yet experimentally confirmed, the principles have been applied to quantum error
correcting codes. Further, the topological nature facilitates analog computation.
Classically, analog computation has been unsuccessful due to its noisy nature. Small
errors can quickly snowball into large errors. Since TQC is immune to local changes,
there is built in protection from small errors, perfect for analog computation. The
model provides a novel way of thinking about quantum computation that is distinct
from the traditional circuit model.
Further, the topological nature of the model may be advantageous to dealing
with problems that are topological in nature. The approximation of the Jones
polynomial is the prime example for this. Perhaps there are other classically hard
topological problems that may be solved efficiently on a quantum computer.

Acknowledgements
I would like to thank my mentor, Randy Van Why, for providing me guidance
in writing this exposition. I would also like to thank my friends, Josef Klafka and
Anand Abraham, for providing feedback and Claudio Gonzales, for introducing me
to this subject.
All diagrams and figures have been drawn with TikZ, the tikz-cd package and
the braids package. These packages can be found on CTAN.

References
[1] Giuliano Benenti, Giulio Casati, and Giuliano Strini. Principles of quantum
computation and information: Volume II: Basic Tools and Special Topics.
World Scientific, 2007.
[2] N. E. Bonesteel. “Braid Topologies for Quantum Computation”. In: Physical
Review Letters 95.14 (2005). doi: 10.1103/PhysRevLett.95.140503.
[3] César Galindo. “On braided and ribbon unitary fusion categories”. In: arXiv
preprint arXiv:1209.2022 (2012).
[4] Daniel F. V. James et al. “Measurement of qubits”. In: Physical Review A
64.5 (Oct. 2001). issn: 1094-1622. doi: 10.1103/physreva.64.052312. url:
http://dx.doi.org/10.1103/PhysRevA.64.052312.
[5] Alexei Yu Kitaev et al. Classical and quantum computation. 47. American
Mathematical Soc., 2002.
[6] Jiannis K Pachos. Introduction to topological quantum computation. Cam-
bridge University Press, 2012.
[7] Alexander Razborov. Lecture Notes on Quantum Computing. url: http :
//people.cs.uchicago.edu/ ~razborov/teaching/QuantumComputing/
notes.pdf (visited on 08/15/2018).
[8] Simon Trebst et al. “A Short Introduction to Fibonacci Anyon Models”. In:
Progress of Theoretical Physics Supplement 176 (2008), pp. 384–407. issn:
0375-9687. doi: 10.1143/ptps.176.384. url: http://dx.doi.org/10.
1143/PTPS.176.384.
18 REFERENCES

[9] Zhenghan Wang. Topological quantum computation. 112. American Mathe-


matical Soc., 2010.
[10] Takeo Yokonuma. Tensor spaces and exterior algebra. 108. American Mathe-
matical Soc., 1992.
REFERENCES 19

1 2 3 4

a d
F15  1 2 3 4
2d
a,c

5 d

b,d

c

4
5
a3
F
1 2 3 4 5

a
c
(F234 )e,d
b

5 1 2 3 4
F 12
b 3 a

e


1 2 3 4 c
,e


5 b,c
e F 1 e4 5

Figure 3. The pentagon equation for the 6j fusion system. The


diagram must commute. It is the analogue of the pentagon equa-
tion for a fusion category.
20 REFERENCES

2 1 3 k 2 1 3
(F213 )a,b

a b

k k

(R 1
a,a
12 )

3
) b,b
(R

1 2 3 2 3 1

a b

,b
k k

1 )c
(F 12

k
23
k 3) a

(F
2 3 2 3
,c

2 3 1

1 c c 1 =
(R1c )k,k c

k k k

Figure 4. The hexagon equation for the 6j fusion system. The


diagram must commute. It is the analogue of the hexagon equation
for a braided fusion category.

Figure 5. The trace closure of a braid resulting in the trefoil knot.


In an appropriate ribbon unitary fusion category, this diagram cor-
responds exactly to Tr(ρA (B)) of the trefoil knot. Physically, the
value corresponds to the probability of obtaining the vacuum af-
ter pairwise fusion of the particles, created from the vacuum, after
braiding.

You might also like