+ All Categories
Home > Documents > Topological Phases, Entanglement and Boson Condensation -...

Topological Phases, Entanglement and Boson Condensation -...

Date post: 24-Mar-2020
Category:
Upload: others
View: 15 times
Download: 0 times
Share this document with a friend
240
Topological Phases, Entanglement and Boson Condensation Huan He A Dissertation Presented to the Faculty of Princeton University in Candidacy for the Degree of Doctor of Philosophy Recommended for Acceptance by the Department of Physics Adviser: Bogdan Andrei Bernevig June 2019
Transcript

Topological Phases, Entanglement and Boson

Condensation

Huan He

A Dissertation

Presented to the Faculty

of Princeton University

in Candidacy for the Degree

of Doctor of Philosophy

Recommended for Acceptance

by the Department of

Physics

Adviser: Bogdan Andrei Bernevig

June 2019

c© Copyright by Huan He, 2019.

All rights reserved.

Abstract

This dissertation investigates the boson condensation of topological phases and the

entanglement entropies of exactly solvable models.

First, the bosons in a “parent” (2+1)D topological phase can be condensed to

obtain a “child” topological phase. We prove that the boson condensation formalism

necessarily has a pair of modular matrix conditions: the modular matrices of the

parent and the child topological phases are connected by an integer matrix. These

two modular matrix conditions serve as a numerical tool to search for all possible

boson condensation transitions from the parent topological phase, and predict the

child topological phases. As applications of the modular matrix conditions, (1) we

recover the Kitaev’s 16-fold way, which classifies 16 different chiral superconductors

in (2+1)D; (2) we prove that in any layers of topological theories SO(3)k with odd

k, there do not exist condensable bosons.

Second, an Abelian boson is always condensable. The condensation formalism in

this scenario can be easily implemented by introducing higher form gauge symmetry.

As an application in (2+1)D, the higher form gauge symmetry formalism recovers

the same results of previous studies: bosons and only bosons can be condensed in

an Abelian topological phase, and the deconfined particles braid trivially with the

condensed bosons while the confined ones braid nontrivially. We emphasize again

that the there exist non-Abelian bosons that cannot be condensed.

Third, the ground states of stabilizer codes can be written as tensor network

states. The entanglement entropy of such tensor network states can be calculated

exactly. The 3D fracton models, as exotic stabilizer codes, are known to have several

features which exceed the 3D topological phases: (i) the ground state degeneracy

generally increase with the system size; (ii) the gapped excitations are immobile or

only mobile in certain sub-manifolds. In our work, we calculate, for the first time,

the entanglement entropy for the fracton models, and show that the entanglement

iii

entropy has a topological term linear to the subregions’ sizes, whereas the topological

phases only have constant topological entanglement entropies.

iv

Acknowledgements

First and foremost, I would like to thank my advisor Prof. B. Andrei Bernevig for

guidance and support. I have benefited enormously from his curiosity, dedication to

physics and perseverance. He provides me with a complete freedom to work on the

problems and topics that I am interested in, and more importantly, he never allows

me to give up easily until we all find a satisfying answer. It is fair to say that rigorous

and solid research is the trademark of our group, which I now completely agree is

truly valuable. I will bear this spirit in mind for the rest of my career.

During the journey, I am indebted to the help from many professors and friends

who make me a better physicist. Among all of them, I owe the most to Prof Nicolas

Regnault, Prof German Sierra and Yunqin Zheng. Many papers were not possible

without their help and efforts. In fact, Nicolas and German are the two aspects of

modern physicists. Nicolas’s numerical technologies are unbelievable. I still remember

how I was amazed when he finished a new program and computation before I finished

speaking my speculations. It turned out that I was wrong and he was correct. I am

lucky to meet and learn from German. Every time I discussed with German, I would

enjoy the aesthetic of his physics and formulas, which I have almost forgotten over

the years of research. His formulas reminded me of my initial motivation of becoming

a physicist: the simple and beautiful laws of nature. Yunqin has been my major

collaborator since the end of my first year in Princeton. In fact, we have known each

other since undergraduate (10 years ago from now). I would never forget all the time

we spent together with chalks on blackboards, pens on papers, and Mathematica on

screens.

The PCTS program and our department certainly bring in brilliant and optimistic

young physicists globally. I am grateful for the collaborations with the postdoctoral

researchers in Princeton: Prof Curt von Keyserlingk, Prof Titus Neupert, Prof Barry

Bradlyn, Prof Jennifer Cano, Prof Yi Li, Dr Abhinav Prem, Dr Lian Biao, Dr Yizhi

v

You and Dr Kiryl Pakrouski. I hope that they also enjoy discussing and working with

me.

It was joyful and also surprisingly efficient to work with Prof Jia-Wei Mei and Dr.

Ji-Yao Chen. A powerful tensor network MATLAB program was delivered in only

a few weeks over the summer. I need to thank Xue-Da Wen, Apoorv Tiwari, Peng

Ye for analytically calculating entanglement entropies, and being reliable and helpful

collaborators for this project. Particularly, Xue-Da is the person who I would like to

discuss all the time because of his enthusiastic passion.

It would be my fault if I did not mention my group members, who I have in-

teractions with, for many enlightening and refreshing conversations and discussions:

Yang-Le Wu, Aris Alexandradinata, Sanjay Moudgalya, Zhi-Jun Wang, Jian Li, Zhi-

Da Song, Fang Xie. I also would like to thank the physicists that I met during confer-

ences, visit, emails and Skype: Rui-Xing Zhang, Chen Fang, Hong Yao, Yuan-Ming

Lu, Meng Cheng, Yang Qi, Yuan Wan, Xie Chen, Giuseppe Carleo, Alan Morningstar,

Yi-Chen Huang, Juven Wang, Yidun Wan, Heidar Moradi, Wen Wei Ho, Tian Lan

and many others.

My life would be miserable if without my friends in Princeton: Jing-Yu Luo, Jie

Wang, Zheng Ma, Jing-Jing Lin, Yu Shen, Wu-Di Wang, Zhen-Bin Yang, Jun Xiong,

Liang-Sheng Zhang, Bin Xu, Zhao-Qi Leng, Tong Gao, Suerfu, Yin-Yu Liu, Jia-Qi

Jiang, Si-Hang Liang, Xin-Ran Li, Xiao-Wen Chen, Bo Zhao, Jun-Yi Zhang, Xue

Song, Xin-Wei Yu, Hao Zheng, Zi-Ming Ji, Ho Tat Lam, Wen-Li Zhao, Po-Shen Hsin,

Xiao-Jie Shi, Chaney Lin and many many others.

I appreciate Prof Duncan Haldane, Prof Waseem Bakr and Prof Silviu Pufu for be-

ing my committee members, and Prof Robert Austin for leading me the experimental

project.

vi

Last but not the least, I appreciate the support from my parents during all these

years. And most importantly, it is beyond my language to describe my gratitude for

many years’ accompany with my love, Yina. This dissertation is dedicated to you.

vii

To my love Yina

viii

Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiv

1 Preliminaries 1

1.1 Stabilizer Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Entanglement Entropy and Tensor Network States . . . . . . . . . . . 4

1.3 Abelian Topological Field Theory . . . . . . . . . . . . . . . . . . . . 7

1.4 Modular Tensor Category . . . . . . . . . . . . . . . . . . . . . . . . 9

2 An Example: Toric Code Model 14

2.1 Stabilizer Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.2 Tensor Network State . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.3 Modular Tensor Category . . . . . . . . . . . . . . . . . . . . . . . . 18

2.4 BF Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3 Boson Condensation 23

3.1 Definitions and Assumptions . . . . . . . . . . . . . . . . . . . . . . . 26

3.2 The Condensation Matrix Mab . . . . . . . . . . . . . . . . . . . . . . 29

3.2.1 Proof that M commutes with T matrix of the A theory . . . . 31

3.2.2 Proof that M commutes with S matrix of the A theory . . . . 31

ix

3.3 The Modular Tensor Category after condensation . . . . . . . . . . . 33

3.4 Simple currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

3.4.1 Introduction to simple currents . . . . . . . . . . . . . . . . . 36

3.4.2 Simple current condensation . . . . . . . . . . . . . . . . . . . 38

3.5 One confined particle theories . . . . . . . . . . . . . . . . . . . . . . 43

3.6 Formalism and implementation . . . . . . . . . . . . . . . . . . . . . 44

3.6.1 Solutions for M . . . . . . . . . . . . . . . . . . . . . . . . . . 45

3.6.2 Solutions for n . . . . . . . . . . . . . . . . . . . . . . . . . . 48

3.6.3 The modular matrices of the new theory . . . . . . . . . . . . 51

3.7 Layer constructions and uncondensable bosons . . . . . . . . . . . . . 54

3.7.1 Theories with Zm-graded condensations . . . . . . . . . . . . . 55

3.7.2 Theories with Z-fold way: Fibonacci TQFT . . . . . . . . . . 60

3.8 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4 No-Go Theorem for Boson Condensation in Topologically Ordered

Quantum Liquids 65

4.1 First No-Go Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.2 Example (i): Multiple layers of Fibonacci . . . . . . . . . . . . . . . . 69

4.3 Example (ii): Single layer of SO(3)k . . . . . . . . . . . . . . . . . . . 71

4.4 Example (iii): Multiple layers of SO(3)k . . . . . . . . . . . . . . . . 74

4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

5 Abelian Boson Condensation in Field Theory 76

5.1 Abelian Boson Condensation Formalism . . . . . . . . . . . . . . . . 77

5.2 Condensations in K-Matrix Chern-Simons Theories . . . . . . . . . . 81

5.2.1 Condensable Condition . . . . . . . . . . . . . . . . . . . . . . 82

5.2.2 Confinement/Deconfinement . . . . . . . . . . . . . . . . . . . 84

x

6 Fracton Models, Tensor Network States and Their Entanglement

Entropies 87

6.1 Stabilizer Code Tensor Network States . . . . . . . . . . . . . . . . . 91

6.1.1 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

6.1.2 Stabilizer Code and TNS Construction . . . . . . . . . . . . . 94

6.1.3 TNS Norm . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

6.1.4 Transfer Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 100

6.2 Entanglement properties of the stabilizer code TNS . . . . . . . . . . 102

6.2.1 TNS as an exact SVD . . . . . . . . . . . . . . . . . . . . . . 102

6.2.2 Summary of the results . . . . . . . . . . . . . . . . . . . . . . 107

6.3 3D Toric Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

6.3.1 Hamiltonian of 3D Toric Code Model . . . . . . . . . . . . . . 109

6.3.2 TNS for 3D Toric Code . . . . . . . . . . . . . . . . . . . . . . 111

6.3.3 Concatenation Lemma . . . . . . . . . . . . . . . . . . . . . . 115

6.3.4 Entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

6.3.5 Transfer Matrix as a Projector . . . . . . . . . . . . . . . . . . 121

6.3.6 GSD and Transfer Matrix . . . . . . . . . . . . . . . . . . . . 124

6.4 X-cube Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

6.4.1 Hamiltonian of X-cube Model . . . . . . . . . . . . . . . . . . 129

6.4.2 TNS for X-cube Model . . . . . . . . . . . . . . . . . . . . . . 131

6.4.3 Concatenation Lemma . . . . . . . . . . . . . . . . . . . . . . 134

6.4.4 Entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

6.4.5 Transfer Matrix as a Projector . . . . . . . . . . . . . . . . . . 141

6.4.6 GSD and Transfer Matrix . . . . . . . . . . . . . . . . . . . . 142

6.5 Haah Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

6.5.1 Hamiltonian of Haah code . . . . . . . . . . . . . . . . . . . . 146

6.5.2 TNS for Haah Code . . . . . . . . . . . . . . . . . . . . . . . 148

xi

6.5.3 Entanglement Entropy for SVD Cuts . . . . . . . . . . . . . . 154

6.5.4 Entanglement Entropy for Cubic Cuts . . . . . . . . . . . . . 161

6.6 Conclusion and Discussion . . . . . . . . . . . . . . . . . . . . . . . . 169

A Appendix for Boson Condensation 172

A.1 Quantum dimensions of A and T . . . . . . . . . . . . . . . . . . . . 172

A.1.1 Proof of da =∑

r∈T nradr . . . . . . . . . . . . . . . . . . . . . 172

A.1.2 Proof of dr = 1q

∑a∈A n

rada . . . . . . . . . . . . . . . . . . . . 173

A.2 Chiral algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

A.3 Condensations in SU(2) CFTs . . . . . . . . . . . . . . . . . . . . . . 181

A.3.1 SU(2)16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184

A.3.2 SU(2)28 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

B Appendix for Uncondensable Bosons 189

B.1 No-go theorem with Abelian sector . . . . . . . . . . . . . . . . . . . 189

B.2 Proof for Example (iii), Multiple layers of SO(3)k . . . . . . . . . . . 191

B.3 General constraints on boson condensation . . . . . . . . . . . . . . . 194

C Appendix for Tensor Network States and Fracton Models 197

C.1 Proof for the Concatenation Lemma for the 3D Toric Code Model . . 197

C.2 Proof for the Concatenation Lemma for the X-cube Model . . . . . . 200

C.3 GSD for the X-cube Model . . . . . . . . . . . . . . . . . . . . . . . . 204

Bibliography 208

xii

List of Tables

xiii

List of Figures

1.1 A link in (2+1)D space. . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 The Hamiltonian terms of the 2D toric code model. Panel (a) is Av

which is a product of 4 Z operators around the vertex v, and Panel

(b) is Bp which is a product of 4 X operators around the plaquette p. 15

2.2 The TNS for the 2D toric code model on a square lattice. On each

bond, we associate a projector g tensor, and on each vertex, we as-

sociate a local T tensor. The connected lines are contracted virtual

indices. The lines with arrows are the physical indices. . . . . . . . . 17

2.3 Left panel: The line operator in the first equation of Eq. (2.12). The

blue line is the path for C. Right panel: The line operator in the

second equation of Eq. (2.12). The blue line is the path for C on the

dual lattice. The red dots represent the excitation created. W (C)

creates two excitations at the ends of C which have −1 eigenvalues of

Av. W (C) creates two excitations at the ends of C which have −1

eigenvalues of Bp. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

xiv

4.1 Tunneling processes mediated by an anyon condensate. The gray re-

gion is a phase in which a boson B is condensed. a) Vertex of a boson

B that localizes a zero mode of anyon ai. In the condensed phase, B

can be converted into an identity particle world line (not shown). By

the axioms of anyon condensation, processes a) and b) are equivalent,

i.e., B can be converted into ai by tunneling through the condensate. 67

4.2 Quantum dimensions and bosons (blue columns) for SO(3)k theories

with a) k = 13 and b) k = 103. These are the smallest k, for which

SO(3)k contains two and four bosons, respectively. Indicated are also

the ranges I–III defined in Eq. (4.10). The maximum quantum dimen-

sion coincides with the boundary between range I and II in Eq. (4.10).

For instance, to apply the no-go theorem to the j = 5 boson in a),

choose Ij=5 = {j = 2} and use that d5 ≈ 3.6 is smaller than d2 ≈ 4.2. 72

5.1 An illustration of integer current and Hodge dual in 3 dimensional

space. In Panel (a), jS1 is the integer which only has support on the

loop S1. It is a 1-form represented by the red arrow. Its Hodge dual

?jS1 is represented by the red square penetrated by S1, which is a 2-

form perpendicular to jS1 . In Panel (b), jS is the integer 2-current

which only has support on the sphere S. In 3 dimensions, ?jS is then

a 1-form going out S denoted by the red dot. . . . . . . . . . . . . . 79

6.1 Examples of TNS lattice wave functions in 1D and 2D. Each node is a

tensor whose indices are the lines connecting to it. The physical indices

- of the quantum Hilbert space - are the lines with arrows, while the

lines without any arrows are the virtual indices. Connected lines means

the corresponding indices are contracted. Panel (a) is an MPS for 1D

systems. Panel (b) is a PEPS on a 2D square lattice. . . . . . . . . . 88

xv

6.2 An illustration of the TNS gauge in MPS. (a) A part of an MPS. A1

and A2 are two local tensors contracted together. (b) We insert the

identity operator I = UU−1 at the virtual level - it acts on the virtual

bonds. The tensor contraction of A1 and A2 does not change. (c) We

further multiply U with A1 and U−1 with A2, resulting in A1 and A2

respectively in Panel (d). The tensor contraction of A1 and A2 is the

same as the tensor contraction of A1 and A2. The TNS does not change

as well. Similar TNS gauges also appear in other TNS such as PEPS. 93

6.3 (a) A plane of TNS on a cubic lattice. (b) TNS on a cube. The

lines with arrows are the physical indices. The connected lines are

the contracted virtual indices, while the open lines are not contracted.

On each vertex, there lives a T tensor, and on each bond, we have a

projector g tensor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

6.4 Transfer matrix (red dashed square) of a 1D MPS. The connected lines

are the contracted virtual indices. The connected arrow lines are the

contracted physical indices. The MPS norm (or any other quantities)

can be built using the transfer matrix. Higher dimensional transfer

matrices are similarly defined for TNS on a cylinder or a torus, by

contracting in all directions except one. This leads to a 1D MPS with

a bond dimension exponentially larger than the TNS one. . . . . . . . 100

6.5 The Hamiltonian terms of the 3D toric code model. Panel (a) is Av

which is a product of 6 Z operators, and Panel (b) is Bp which is a

product of 4 X operators. The circled X and Z represent the Pauli

matrices acting on the spin-1/2’s. The toric code Hamiltonian includes

Av terms on all vertices v and Bp terms on all plaquettes p. . . . . . . 110

6.6 Contraction of two local T tensors in the z-direction. We emphasize

that there is no projector g tensor in this figure. . . . . . . . . . . . . 115

xvi

6.7 The splitting of tensors near the entanglement cut. . . . . . . . . . . 118

6.8 The Hamiltonian terms of the X-cube model: (a) Av,yz, (b) Av,xy, (c)

Av,xz and (d) Bc. The circled X and Z represent the Pauli matrices

acting on the physical spin-1/2’s. . . . . . . . . . . . . . . . . . . . . 128

6.9 Figures for several regions A for which we calculate the entanglement

entropies. (1) Region A is a cube of size lx × ly × lz. (2) Region A is a

cube of lx× ly × lz with a hole of size l′x× l′y × l′z in it. (3) Region A is

a cube of size lx× ly × lz and a small cube of height l′z on top of it. (4)

Region A is a cube of size lx × ly × lz carved on top of it a small cube

of height l′z. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

6.10 Derivations for the first equation in Eq. (6.107). The rest two equations

can be proved similarly. In the first step, the physical X operators can

be transferred to the virtual level by using Eq. (6.36), and in the third

step, all the virtual X operators are exactly canceled in pairs (dashed

red rectangles in the third figure) due to Eq. (6.87). . . . . . . . . . 141

6.11 Examples for the X-cube TNS in a xy-plane, obtained by acting

WZ [Cz,x] and WZ [Cz,y] on the constructed TNS. The intersection of

one WZ [Cz,x] operator and one WZ [Cz,y] operator with the xy-plane is

only one Pauli Z operator, i.e., the circled blue Z in this figure. . . . 144

6.12 We act WZ [Cx,y] and WZ [Cy,x] operators on Panel (a) in Fig. 6.11.

Hence, we have four TNSs in this xy-plane that can be related to each

other. See the text for the explanations. . . . . . . . . . . . . . . . . 144

6.13 Tensor contraction for the Haah code TNS. (a) The lattice size is 2×

3× 3. (b) The lattice size is 3× 3× 3 . . . . . . . . . . . . . . . . . 148

6.14 Region A contains all the spins connecting with l− 1 T tensors which

are contracted along z direction. The figure shows an example with

l = 4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

xvii

6.15 Region A contains all the spins connecting with T tensors which are

contracted in a “tripod-like” shape, where three legs extend along

x, y, z directions. There are three legs extending along x, y, z direc-

tions respectively. In general, three legs can have different length, each

with lx−1, ly−1, lz−1 cubes along three directions. This figure shows

an example where lx = ly = lz = 3. . . . . . . . . . . . . . . . . . . . 155

6.16 Transferring the Pauli X operators of the Bc operator from the region

A (a) to the region A (b). . . . . . . . . . . . . . . . . . . . . . . . . 162

xviii

Chapter 1

Preliminaries

The topological phase has been one of the central topics in condensed matter physics

for the past decades since the discovery of integer and fractional quantum Hall

effects[1, 2]. The definition of topological phases has been evolving over the years

of research[3, 4, 5, 6, 7, 8, 9]. It generally refers to the gapped phases of matter

which ground states are different due to the boundary conditions. For instance,

the ground state degeneracy can vary when the system is on torus or sphere[4],

or the boundary can support protected gapless modes when the system is on a

disk[10, 11, 12, 13, 14, 15, 16]. Another significant signature is that the bulk of

the topological phases supports excitations which have nontrivial braiding statistics

and topological spins[17, 18, 19, 20] than those of bosons and fermions. They are

dubbed as “anyons”. In this dissertation, we mostly focus on the topological phases

with anyons.

This dissertation studies two aspects of topological phases in 2 dimensions and

3 dimensions: boson condensations[21] and entanglement entropy[22, 23]. It will be

based on several different languages for topological phases, namely, stabilizer codes,

topological field theories and modular tensor categories. The stabilizer codes and

topological field theories provide the exactly solvable lattice Hamiltonian and field

1

theory models respectively. The modular tensor category, on the other hand, “for-

gets” the specific models and details such as the lattice or the couplings, but only

“remembers” the universal data such as the fusion rules and the braiding statistics

of the gapped excitations. In this chapter, each of the languages will be briefly re-

viewed to provide the foundations for the rest of the dissertation. In Chapter 2, the

2 dimensional toric code model will be described in each of the languages.

1.1 Stabilizer Codes

The stabilizer code refers to a class of exactly solvable Hamiltonian models specified

as follows. The key ingredient is the notion of qubit, which is nothing but a 2-level

system. For instance, a qubit can be a spin-12. On the qubit, the operators can be

generated by the Pauli X or Pauli Z operators satisfying anti-commutation relations:

XZ = −ZX, X =

0 1

1 0

, Z =

1 0

0 −1

. (1.1)

Conventionally the Pauli Z basis is preferred:

|0〉 =

1

0

, |1〉 =

0

1

. (1.2)

Hence,

Z|0〉 = |0〉, Z|1〉 = −|1〉, X|0〉 = |1〉, X|1〉 = |0〉. (1.3)

The stabilizer code is a many-qubit system, and the total Hilbert space is defined as

a tensor product of the Hilbert spaces of each qubit. Hence, the Pauli operators on

different sites commute:

[Xi, Zj] = 0, [Xi, Xj] = 0, [Zi, Zj] = 0, ∀ i 6= j. (1.4)

2

where the subscripts label different qubits. Each of the Hamiltonian terms is a product

of Pauli X and Pauli Z operators. The Hamiltonian is exactly solvable because all

local Hamiltonian terms are required to be commutative with each other. One simple

example for the stabilizer codes is a 3-qubit system, and its Hamiltonian is:

H = −Z1Z2Z3 −X1X2 −X2X3 (1.5)

Note that all three terms are commutative:

[Z1Z2Z3, X1X2] = 0, [Z1Z2Z3, X2X3] = 0, [X1X2, X2X3] = 0. (1.6)

These commutation relations imply that:

[H,Z1Z2Z3] = 0, [H,X1X2] = 0, [H,X2X3] = 0. (1.7)

The eigenstates of the Hamiltonian are thus also the eigenstates of the local Hamil-

tonian terms. Particularly, the ground state satisfies:

Z1Z2Z3|GS〉 = |GS〉, X1X2|GS〉 = |GS〉, X2X3|GS〉 = |GS〉. (1.8)

The ground state wave function, in terms of the Pauli Z basis, contains all the con-

figurations satisfying the first equality of Eq. (1.8):

|GS〉 =1

2(|000〉+ |110〉+ |101〉+ |011〉) (1.9)

The ground state wave function is much simplified if we use Pauli X basis instead:

|GS〉 =1√2

(|000〉x + |111〉x) , (1.10)

3

i.e. a GHZ state.[24]

Another example for stabilizer codes, extremely useful in the topological phases,

is the toric code model on a 2 dimensional lattice[19]. We defer the introduction of

this model to Chapter 2. There we will thoroughly describe this model in each of the

languages.

1.2 Entanglement Entropy and Tensor Network

States

The entanglement (von Neumann) entropy for a wave function |ψ〉 is defined as:

S(A) = TrA (ρA ln (ρA)) (1.11)

where ρA is the reduced density matrix for the subregion A by tracing out A’s com-

plement A:

ρA = TrA|ψ〉〈ψ|. (1.12)

The eigenvalues of the ρA are commonly referred as the entanglement spectrum. An

equivalent method to calculate the entanglement entropy is to perform the singular

value decomposition to the wave function coefficients M :

|ψ〉 =∑ij

Mij|i〉A ⊗ |j〉A,

M = UΛV †.

(1.13)

where |i〉A and |j〉A form the orthonormal basis for the subregion A and A respectively.

U , V are unitary matrices and Λ is a diagonal matrix containing the singular values.

It is easy to show that the eigenvalues of the reduced density matrix are the square

of the singular values.

4

A breakthrough was the conjecture of the area law[25]: the ground states of a

local and gapped Hamiltonian has entanglement entropy linear to the subregion’s

perimeter length/area:

S(A) ∼ Area (∂A) (1.14)

where S(A) is the entanglement for the subregion A and Area (∂A) refers to the A’s

perimeter length/area. For instance, in 1 dimension, the entanglement entropy of the

gapped local Hamiltonian is proved to be[26]:

S(A) ∼ Const. (1.15)

since in 1 dimension, the subregion A is a line segment, and ∂A only contains two

end points.

Further, the entanglement entropy and spectrum, measured using the ground state

wave functions, serve as the smoking gun to numerically and theoretically distinguish

2 dimensional topological phases. Two results need to be highlighted:

1. For a topological phase, the entanglement entropy scales as:

S(A) = α Area (∂A)− γ (1.16)

where the constant γ is a universal constant independent of the subregion’s

size[23, 22]. “Universal” means that γ is the same at any point in the gapped

phase. By computing γ from a microscopic model, most of the topological

phases are eliminated, and only few of them can be the candidates which require

further investigation.

2. In fractional quantum Hall systems, the entanglement spectrum degeneracy

provides the “fingerprint” to identify the different quantum Hall states[27]. The

5

entanglement spectrum degeneracy is argued to be the same as the conformal

field theories on the fractional quantum Hall edge[28].

Inspired by the area law, an efficient ansatz for the ground state wave function was

proposed and widely applied both numerically and analytically. This set of ansatz

is known as tensor network states. To name a few, they are matrix product state

(MPS)[29], projected entangled pair states (PEPS)[30], multi-entanglement renor-

malization ansatz[31, 32]. In this section, we briefly introduce MPS in 1 dimension.

PEPS is the immediate generalization of MPS in higher dimensions and will be heav-

ily used in Chapter 6. Suppose we have an MPS with periodic boundary condition

whose local tensor is A:

∑s1,s2,...,sN

∑i1,i2,...,iN

As1i1,i2As2i2,i3

. . . AsNiN ,i1 |s1, s2, . . . , sN〉 (1.17)

where the tensor A has one physical index sn of dimension d and two virtual indices

in of dimension D. The digram representation for the tensor A at the n-th site is:

(1.18)

where each line represents an index of the tensor. The physical index sn is dis-

tinguished from the virtual indices by the arrows. Diagrammatically, the MPS is

represented as:

(1.19)

6

The connected lines imply the contraction over the two indices. Note that the entan-

glement of such an MPS is upper bounded by the virtual bond dimension D:

S ≤ 2 ln(D), (1.20)

because the rank of the reduced density matrix or the rank of the singular value

matrix is upper bounded by D2 where the number 2 comes from the two cuts.

1.3 Abelian Topological Field Theory

In this section, we review a class of topological field theories, K-matrix Chern-Simons

theories[33], which can describe any Abelian topological phases. The Lagrangian of

a K-matrix Chern-Simons theory is:

L =KIJ

4πεµνλaIµ∂νaJ,ρ (1.21)

where gauge fields aI are U(1) gauge fields, and the matrix KIJ is symmetric and all

elements are integers. In a differential geometry language, it can also be written as

differential forms and exterior derivatives.

L =KIJ

4πaI ∧ daJ (1.22)

When the diagonal elements of the matrix K are all even integers, this type of K-

matrix Chern-Simons theories is called “bosonic”. For a K-matrix whose diagonal

elements have odd integers, it is refereed as “fermionic”. In this dissertation, we

mostly focus on the bosonic K-matrix Chern-Simons theories.

7

Figure 1.1: A link in (2+1)D space.

K-matrix Chern Simons theories have gauge invariant loop operators:

U{n}(S1) = exp i

˛S1

nIaI

, nI ∈ Z (1.23)

where S1 is a closed loop in (2+1)D spacetime and is also the worldline of the corre-

sponding pair of excitations. The linking correlation function of two loop operators

can be calculated by the Green function method[34]:

〈U{n}(S11)U{m}(S

12)〉 = exp i

(2πm ·K−1 · n

)(1.24)

where loops S11 and S1

2 form a link as in Fig. 1.1, and “·” is the matrix multiplication.

The physical interpretation of the linking correlation functions is that they are the

braiding statistics of the two excitations represented by U{n}(S11) and U{m}(S

12). In

the modular tensor category, this linking correlation is encoded in the modular S

matrix. The topological spin for each loop operator is:

Θ(U{n}

)= exp i

(πn ·K−1 · n

)(1.25)

In the modular tensor category, the topological spin is encoded in the modular T

matrix.

One observation is that U{K·l} is a topologically trivial operator when l is an

integer vector: it has trivial linking correlation with all other loop operators, and it

8

has trivial topological spin:

〈U{K·l}(S11)U{m}(S

12)〉 = 1

Θ(U{K·l}

)= 1

(1.26)

Moreover, if the charge vectors of two operators differ byK·l, i.e., considering U{n} and

U{n+K·l}, then they are topologically identical, since their linking correlation functions

with all other loop operators are the same and they have the same topological spin:

〈U{n}(S11)U{m}(S

12)〉 =〈U{n+K·l}(S

11)U{m}(S

12)〉

Θ(U{n}

)=Θ

(U{n+K·l}

) (1.27)

Hence, we only need to consider the loop operators U{n}(S1) up to the shift K · l

where l is an arbitrary integer vector.

1.4 Modular Tensor Category

In this section, we present a short review of the modular tensor category description

of a (2+1)-dimensional TQFT. This approach only describes the low energy exci-

tations of the TQFT, i.e., the anyons. The anyons are usually labeled by objects

a, b, c, . . . and are supplemented by other data, such as the fusion coefficients N cab.

For a comprehensive overview of the category theory approach of TQFT, we refer the

reader to Refs. [20, 35, 36]. Here, we only present a brief and simple review of the

important properties that we frequently use in this dissertation.

Fusion rules and quantum dimension The anyons of a TQFT can fuse. When

two anyons come close to each other spatially, they can fuse into other anyons. An

analogy can be drawn to the algebra of spins: if we take two spin 12

particles, they

can fuse into either spin 0 and spin 1 particle. For this, we would write, in group

9

representation theory, the fusion rule 12× 1

2= 0 + 1. In general, the fusion of anyons

in a TQFT is represented via

a× b =∑c

N cabc, (1.28)

where a, b, c are labels for the anyons, and the fusion coefficients N cab are non-negative

integers. The fusion can be represented by a state |a, b; c, µ〉 in the fusion vector space

V abc . Here µ = 1, · · · , N c

ab labels the vectors that form a basis of the N cab-dimensional

fusion vector space V abc .

Just like the fusion of spins, we require that the fusion rules are symmetric or

commutative, that is, a× b is equivalent to b× a. This translates to

N cab = N c

ba. (1.29)

Moreover, fusion rules are also associative. Suppose we take three anyons a, b, c

and try to fuse them. Then we have two ways to do so: (a× b)× c and a× (b× c).

We require the fusion rule to be associative by requiring that the two fusions yield

the same result. In terms of the fusion coefficients, this translates to

∑d,e

NdabN

edc =

∑d,e

N eadN

dbc. (1.30)

Other important data associated with anyons are their so-called quantum dimen-

sions da, db, · · · . This concept appears because anyons are associated with nontrivial

internal Hilbert spaces. Again, we can take the example of spins to illustrate this. In

the case of spin 12, where 1

2× 1

2= 0 + 1, spin 1

2is associated with a two-dimensional

Hilbert space, and meanwhile spin 0 is associated with a one-dimensional Hilbert

space, spin 1 a three-dimensional Hilbert space. As we can see, the total dimension

of Hilbert space does not change after fusion. The product 12× 1

2has a (2× 2 = 4)-

10

dimensional Hilbert space while 0 + 1 has a (1 + 3 = 4)-dimensional Hilbert space.

Similarly, in a TQFT, we also have

dadb =∑c

N cabdc. (1.31)

The above equation can be viewed as an eigenvalue equation of a matrix Na whose

entries are (Na)bc = N cab. The eigenvector is (db), the eigenvalue is da. Equation (1.30)

says that all matrices Na, Nb, . . . commute and thus they have common eigenvectors,

one of which is the vector of all quantum dimensions. The total quantum dimension

of a TQFT D is defined as the norm of the quantum dimension vector, D =√∑

a d2a.

If all anyons of a TQFT have quantum dimension 1, we call such a theory Abelian.

If there exist anyons with quantum dimension larger than 1, we call such a theory non-

Abelian. This is intimately related to the Perron-Frobenius theorem, where da is a

Frobenius eigenvalue, and hence has to satisfy minb∑

c(Na)bc ≤ da ≤ maxb∑

c(Na)bc.

Hence da > 1 implies that there exists a b such that∑

c(Na)bc > 1, so a× b contains

more than one particle.

Braiding, topological spin and modular matrices Another physically impor-

tant concept in a TQFT is braiding. This allows us to determine how a state trans-

forms when its anyons are adiabatically moved around each other. In Abelian theories,

when we adiabatically move an anyon a fully around another anyon b, the state trans-

forms through multiplication by a universal monodromy phase. For example, if we

take a fermion around a π flux, the wave function obtains a topological minus sign

−1. Another special case is when we exchange two identical abelian anyons a. This

process defines the topological spin θa for the particle a.

In non-Abelian theories, the braiding operation Rab between two anyons a and b is

an operator that acts on the Hilbert space V abc which describes states of a and b that

fuse into a fixed anyon c. If we denote a basis of V abc by |a, b; c, µ〉 with µ = 1, · · · , N c

ab,

11

then Rab has the representation

Rab|a, b; c, µ〉 =∑ν

[Rabc ]µν |b, a; c, ν〉. (1.32)

In this notation, the topological spin for an anyon a is defined as

θa =1

da

∑c

dcTrc[Raac ], (1.33)

where Trc[· · · ] is the trace taken in the fusion vector space V aac .

Given the braiding Rab, we can construct the modular matrices S and T which

are the same modular matrices encoding the global data of S and T in a CFT. They

are given by

Sab =∑c

N cabTr[Rab

c Rbac ]dc, (1.34a)

Tab = θaδab. (1.34b)

By definition, S is a symmetric matrix. Moreover, in a modular tensor categories, S

and T are unitary matrices satisfying S†S = SS† = 1, T †T = TT † = 1.

In Refs. [37, 38, 39, 40], the S matrix is used as an order parameter to detect

topological phase transitions and anyon condensations. The implicit assumption in

doing so is that the S matrix represents physical, measurable properties of the state,

unlike, say, the gauge-dependent F -symbol.

F -symbol F -symbol is in fact the generalized version of Wigner 6j-symbol, and

guarantees that all the fusion processes in the MTC are all consistent:

|a, b; e, α〉|e, c; d, β〉 =∑f,µ,ν

[F abcd ](e,α,β)(f,µ,ν)|b, c; f, µ〉|a, f ; d, ν〉 (1.35)

12

The F -symbol also needs to satisfy pentagon and hexagon equations, the details of

which are omitted in this dissertation.

13

Chapter 2

An Example: Toric Code Model

The toric code model and its generalizations have been extensively studied in the

literature. We will present the toric code model in different languages including the

stabilizer code, the tensor network state, the topological field theory and the modular

tensor category.

2.1 Stabilizer Code

The 2D toric code model, realizing a discrete Z2 gauge symmetry on lattice[41], is a

stabilizer code defined on any 2D random lattice. It exhibits the topological order

and supports gapped anyonic excitations. For simplicity, the 2D toric code introduced

in this section is defined on a square lattice with physical spins on all bonds of the

lattice, and the Hamiltonian consists of vertex terms and plaquette terms:

H = −∑v

Av −∑p

Bp. (2.1)

14

Z Z

x

x

xx

x

y

Z

Z

(a) (b)

Figure 2.1: The Hamiltonian terms of the 2D toric code model. Panel (a) is Av whichis a product of 4 Z operators around the vertex v, and Panel (b) is Bp which is aproduct of 4 X operators around the plaquette p.

Here, Av is the product of four Pauli Z matrices around a vertex v and Bp is the

product of four Pauli X matrices around a plaquette p.

Av =∏i∈v

Zi, Bp =∏i∈p

Xi. (2.2)

These two terms are illustrated in Fig. 2.1. Note that all these operators will commute

with each other:

[Av, Av′ ] = 0, ∀ v, v′,

[Bp, Bp′ ] = 0, ∀ p, p′,

[Av, Bp] = 0, ∀ v, p.

(2.3)

Hence, the ground states of the 2D toric code model need to satisfy the constraints:

Av|GS〉 = |GS〉, Bp|GS〉 = |GS〉, (2.4)

for all vertices v and plaquettes p. Each of the eigenvalue equations is a constraint

for the ground state wave function. Note that the local Hamiltonian terms satisfy the

15

following redundancy on a closed manifold:

∏v

Av = 1,∏p

Bp = 1. (2.5)

The ground state degeneracy (GSD) satisfying Eq. (2.4) is counted as follows:

GSD =2# of qubits

2# of indep. GSD constraints=

2# of bonds

2# of vertices + # of plaquettes - 2= 22−χ (2.6)

where

χ = # of vertices + # of plaquettes - # of bonds (2.7)

is the Euler characteristic. For a torus, χ = 0. Thus the GSD on a torus is 4. The

number 2 in the exponent comes from the redundancy in Eq. (2.5). The calculation

can certainly be generalized to other manifolds, and the GSD only depends on the

Euler characteristic and hence topological.

2.2 Tensor Network State

In order to conveniently construct the TNS for the 2D toric code model, we introduce

projector g tensors on each bond of the square lattice and local T tensors on each

vertex of the square lattice. The physical or virtual index is of dimension 2 and

labeled as 0, 1. The g tensor has 1 physical index and 2 virtual indices, and the T

tensor has 4 virtual indices. The TNS is depicted in Fig. 2.2. The same construction

will be heavily used in Chapter 6.

The projector g tensor identifies the physical index with the virtual indices

gpij =

1 if p = i = j ∈ {0, 1}

0 otherwise

(2.8)

16

where p is the physical index, and i and j are thr virtual indices. The rest of the

calculations is to construct a T tensor such that the TNS satisfies Eq. (2.4) on the

TNS. We choose to directly act Av and Bp operators on the TNS. Note that g tensor

identifies the physical index and the virtual indices. The actions of X and Z Pauli

matrices on the physical indices are transferred to the virtual indices:

TZ

Z

Z

Z

Z

Z

Z

Z

g

g

g

g

g

=

Tg

gg g

x

x

x x

x xx

xx x

x

x=

. (2.9)

Tg

x

y

Figure 2.2: The TNS for the 2D toric code model on a square lattice. On each bond,we associate a projector g tensor, and on each vertex, we associate a local T tensor.The connected lines are contracted virtual indices. The lines with arrows are thephysical indices.

17

In order to implement Eq. (2.4), we require the four Pauli Z matrices (in the first

line) and two Pauli X matrices (in the second line) acting on the virtual indices in the

dashed red squares to be identity operations. Then we have two (strong) conditions

respectively:

Txx,yy = (−1)x+x+y+yTxx,yy

Txx,yy = T(1−x)x,(1−y)y = T(1−x)x,y(1−y)

= Tx(1−x),(1−y)y = Tx(1−x),y(1−y).

(2.10)

where x, x are the two indices of the T tensor in the x-direction, and the y, y are the

two indices of the T tensor in the y-direction. As a result, the local T tensors are

fixed up to an overall factor:

Txx,yy =

0 if x+ x+ y + y = 1 mod 2

1 if x+ x+ y + y = 0 mod 2.

(2.11)

Hence, we have constructed the TNS for the ground state. It is a natural question

that how to construct all degenerate ground state on a torus, since this method only

gives rise to one tensor. The solution to this puzzle is that the boundary conditions

of the TNS can be twisted in both directions without breaking the conditions in

Eq. (2.4)[30].

2.3 Modular Tensor Category

The excited states are also labeled by the eigenvalues of Av and Bp, due to the

fact that all local Hamiltonian terms commute with each other. The excited state

has eigenvalues −1 for some local Hamiltonian terms, while the ground state has all

eigenvalues 1 for all local terms. To create such excitations from the ground state, an

18

Figure 2.3: Left panel: The line operator in the first equation of Eq. (2.12). Theblue line is the path for C. Right panel: The line operator in the second equation ofEq. (2.12). The blue line is the path for C on the dual lattice. The red dots representthe excitation created. W (C) creates two excitations at the ends of C which have−1 eigenvalues of Av. W (C) creates two excitations at the ends of C which have −1eigenvalues of Bp.

observation is that line operators can be constructed as follows:

W (C) =∏i∈C

Xi, W (C) =∏i∈C

Zi, (2.12)

where C is a path on the lattice while C is a path on the dual lattice. See Fig. 2.3 for

an illustration. When the line operators act on the ground states, the Hamiltonian

terms near the two ends of the line have eigenvalues −1. If C or C is closed, then the

energy remains the same as the ground state energy. Thus the excitations are always

created by pairs. Moreover, if C or C winds around the cycles of the lattice, for

instance the two cycles of a torus, then W (C) or W (C) commutes with all the local

terms in the Hamiltonian, and cannot be generated by the local Hamiltonian terms.

Hence, these winding operators serve as good quantum numbers to distinguish the 4

ground states on torus.

The excitation, measured by an Av operator and created by W (C), is commonly

dubbed as e (charge), while the excitation, measured by an Bp operator and created by

N(NC), is dubbed as m (flux). Their composite particle is dubbed as f . Combining

the trivial particle denoted as 1, there are four excitations: 1, e, m and f . Their

19

fusion rules can be derived from their operators:

e× e = 1, m×m = 1, f × f = 1, e×m = f. (2.13)

The braiding statistics of these excitations can be derived from their operators’ com-

mutation relation. Their braiding statistics is encoded in the modular S matrix in

the basis of (1, e,m, f):

S =1

2

1 1 1 1

1 1 −1 −1

1 −1 1 1

1 −1 −1 1

(2.14)

The matrix element Sab is the braiding statistics of the anyon a and the anyon b.

Their topological spins are encoded in the modular T matrix:

T =

1 0 0 0

0 1 0 0

0 0 1 0

0 0 0 −1

(2.15)

The diagonal matrix element Taa is the topological spin for the anyon a.

The F -symbols is omitted in this dissertation since it is trivial for the toric code

model.

2.4 BF Field Theory

In this section, we show that a Z2 BF theory will reproduce the same physics as that

in the toric code Hamiltonian model. The Lagrangian for a ZN BF theory is:

L =N

2πεµνρbµ∂νaρ. (2.16)

20

For the reader who are familiar with differential forms, it is equivalent to write:

L =N

2πb ∧ da. (2.17)

Note that this Lagrangian can be casted in the form of K-matrix Chern-Simons

theory. Restricting ZN to Z2, the Lagrangian reduces to:

L =2

2πεµνρbµ∂νaρ. (2.18)

This Lagrangian exhibits the gauge symmetry as follows:

aµ 7→ aµ + ∂µα, bµ 7→ bµ + ∂µβ. (2.19)

where α and β can be any functions.

Applying canonical quantization for this gauge theory will reproduce the same

results for the toric code model. We start with gauge fixing. Tuning α and β can fix

the temporal gauge:

at = 0, bt = 0. (2.20)

The Lagrangian becomes:

L =1

π(bt (∂xay − ∂yax) + at (∂xby − ∂ybx)− bx∂tay + by∂tax)

=1

π(−bx∂tay + by∂tax) .

(2.21)

Temporal gauge fixing enforces the Gauss constraints:

∂xay − ∂yax = 0, ∂xby − ∂ybx = 0. (2.22)

21

The rest two terms in the Lagrangian imply the commutation relations:

[bx(x, y), ay(x′, y′)] = −iπδ (x− x′) δ (y − y′) ,

[by(x, y), ax(x′, y′)] = iπδ (x− x′) δ (y − y′) .

(2.23)

Therefore, the gauge fields bµ serve as the electric fields for the gauge fields aµ.

The lattice Hamiltonian language is in fact exactly the same as the field theory

language. The observation is that when the gauge fields aµ is identified with the

Pauli X operators on the lattice, then using Eq. (2.23) the gauge fields bµ will be

identified with Pauli Z operators on the dual lattice. The two Gauss constraints

in Eq. (2.22) are in fact the two constraints for the ground states of the toric code

model in Eq. (2.4): the first one of Gauss constraints being the Bp while the second

one being Av.

The line operators of Eq. (2.12) in the lattice Hamiltonian language are the Wilson

operators in the field theory language:

W (C) = exp

(i

ˆC

aµdxµ

),

W (C) = exp

(i

ˆC

bµdxµ

).

(2.24)

Note that these two operators are not gauge invariant when C and C are open, which

indicates that there are additional degrees of freedom attaching to the ends of C or C.

This corresponds to the lattice Hamiltonian language that the open line operators will

cost energy. When C and C are closed, then the two operators are gauge invariant in

the field theory language. This corresponds to the lattice Hamiltonian language that

the closed line operators will not cost energy.

22

Chapter 3

Boson Condensation

Prior to the discovery of topological order, it was well known that bosons can macro-

scopically occupy a single quantum state, a fact which allows for the possibility of a

Bose-Einstein condensation phase transition. In a topologically ordered phase, bosons

are more complicated particles: They can have nontrivial braiding behavior with other

anyons[20, 35], and even more exotically they can carry nonlocal internal degrees of

freedom[20], in which case they are called non-Abelian bosons. Notwithstanding,

such bosons can sometimes condense[42, 21, 37, 43, 44, 45, 46, 47, 48, 49, 50, 51]. It

is then natural to ask how this condensation affects the topological order, namely,

what is the fate of the other anyons in the phase. The answer is that anyon con-

densation induces transitions between different topologically ordered phases in such

a way that universal properties of the anyons of the condensed phase can be in-

ferred from those of the initial phase, together with a list of condensed bosons. This

framework of anyon condensation transitions found many applications in the study

of topological order, [52, 53, 54, 55, 56, 57] in particular in solving the question of

bulk-boundary correspondence [58, 59, 59, 60, 61, 62], or in deducing the univer-

sal properties of domain walls [63, 64, 65, 66, 67, 68, 69, 70, 71] and other external

defects [72, 73, 74, 75, 76, 77, 78, 79].

23

The universal aspects of topologically ordered phases are captured by topological

quantum field theories[80]. Among these, the axiomatic approach of category the-

ory [81, 82, 83, 84, 85], more concretely the formulation of modular tensor categories

(MTCs), is particularly powerful and, to our knowledge, provides a complete char-

acterization of topological order in two-dimensional space. [20, 35] At a basic level,

MTC’s are characterized by the types of anyons that appear in the phase as well as

their interrelations in the form of fusion and braiding information, the so-called “F

moves” and “R moves”.

In correspondence with the different descriptions of topological order itself, sev-

eral formulations of anyon condensation were developed. In the context of MTCs,

the phase after condensation is found by studying commutative separable Frobenius

algebras [86, 87, 88, 89, 90] of the initial theory. [91, 62] Bais and Slingerland trans-

lated this procedure into the language of anyon models [42], but their formulation

did not give a systematic method for determining properties of the phase after con-

densation. This was later achieved by Eliens et al. in Ref. [21] via a diagrammatic

formulation of the condensation problem that makes use of the so-called vertex lifting

coefficients. These allowed them to embed the fusion and braiding processes of the

condensed phase in the initial anyon model. However, all these approaches fall short

of providing an algorithmic formulation of boson condensation in a way that could,

for example, be implemented in a computer algebra program allowing for systematic

studies of possible condensations.

In this chapter, we reformulate the problem of boson condensation in anyon mod-

els axiomatically and purely algebraically. The resulting formalism is based on a small

number of natural assumptions such as the commutativity of fusion and condensation

as well as an assumption about the topological spins of the anyons after condensation.

Our approach puts the modular matrices S and T of the initial anyon model center

stage, instead of focusing on the F and R moves, which are the key objects of interest

24

for the diagrammatic approach. [21] The F and R moves are in general notoriously

hard to compute even for relatively simple theories. Our goal is to find the modular

matrices S and T of the final theory after condensation. Using our algebraic formula-

tion, we propose an algorithm that determines all possible condensation instabilities

of an anyon model and can be efficiently implemented on a computer. We solve for

the condensation via a series of linear algebra problems, involving the factorization

of nonnegative integer matrices.

Besides its utility for computer-aided calculations, our algebraic formulation of

condensation also facilitates analytical derivations. As an example, we discuss layer

constructions of topologically ordered states and easily reproduce the known result

that 5 and 10 layers of the Fibonacci anyon model cannot undergo a condensation

transition.

This chapter is structured as follows. In Sec. 3.1 we formulate the condensation

problem along with the axioms relating to fusion rules. In the following Sec. 3.2, we

present the assumptions that allow us to deduce the braiding properties of the theory

after the condensation transition, and several implications are derived. In Sec. 3.3,

we derive central equations which constrain S and T . Subsequently, we show in

Secs. 3.4 and 3.5 that a weaker set of axioms suffices, if the condensate consists of so-

called simple currents and if only one particle is confined through the condensation

transition, respectively. We formulate an algorithm for solving the condensation

problem in Sec. 3.6. The final Sec. 3.7 gives examples of condensation transitions

in multi-layered anyon models and discusses obstructions to boson condensation in 5

and 10 layers with Fibonacci anyons.

25

3.1 Definitions and Assumptions

In this section we present the formalism underpinning anyon condensation, following

Refs. [42] and [21] closely. Our discussion is self-contained with respect to the previous

literature on anyon condensation, but assumes that the reader is familiar with the

basic concepts of MTCs [19] (see Sec. 1.4 for a short review).

The input for our approach to anyon condensation is a MTC A (the uncondensed

theory), and a set of restriction and lifting coefficients, which relate the particle

excitations in A to those in T (the condensed theory). In general, T is only a

fusion category, because it may contain some excitations which are confined by the

surrounding condensate. Projecting out these confined excitations, we are left with a

deconfined condensed MTC that we denote as U . Our goal is to find possible MTCs

U given A and some basic information about the condensate, such as which bosons

condense.

In what follows, we will consider the Bose condensation of a collection of bosons in

the original theory A. This collection of anyons is called the condensate and becomes

part of the vacuum in the new intermediate fusion category T . In condensing these

bosons, a generic anyon a ∈ A will become (or “restrict to”) a superposition of

particles t ∈ T

ar7−→∑t∈T

ntat, ∀a ∈ A (3.1)

with the coefficients nta ∈ Z≥0, where we assume that nta = nta and bars denote

antiparticles (see Sec. 1.4). Equation (3.1) defines the “restriction map”. We will

also use the phrase “a restricts to∑

t ntat” to describe Eq. (3.1). It is possible that

more than one particle t appears on the righthand side of Eq. (3.1), in which case

we say that “a splits into∑

t ntat”. Condensed particles (bosons b in the condensate)

have the additional special property that nϕb 6= 0, where ϕ is the vacuum particle in

26

T , that is, their restriction contains the identity of the new T theory. If nϕb 6= 0, then

nϕb = nϕb, i.e., both the boson and its antiparticle must condense at the same time.

The reverse (or, more precisely, adjoint) operation to restriction is called “lifting”.

For a particle t ∈ T , all the particles in A which restrict to t are defined to be the lifts

of t. The lifting coefficients are the same nta that we used in defining the restriction.

Formally, lifting is defined by

tl7−→∑a∈A

ntaa, ∀t ∈ A. (3.2)

Finally, we define particles in T whose lifts do not share a common topological spin

θa as confined, that is

t : confined ⇔ ∃ a, b such that ntantb 6= 0 with θa 6= θb. (3.3)

Conversely, the deconfined particles in T are the particles whose liftings do share a

common topological spin, which becomes identified with the spin of the deconfined

particle, that is

t : deconfined ⇔ ∀ a, b such that ntantb 6= 0 then θa = θb. (3.4)

Obviously, any particle t ∈ T is either deconfined (t ∈ U) or confined (t ∈ T /U).

With these definitions in place, we now make a fundamental assumption from which

we will derive the structure of the theory after condensation. We assume that the

restriction A → T commutes with fusion. This is represented by the diagram

A⊗A f //

r⊗r��

Ar��

T ⊗ Tf

// T

27

in which f represents fusion and r represents restriction.

More explicitly, the commuting diagram can be written in terms of anyon basis

∑r,s∈T

nransbN

trs =

∑c∈A

N cabn

tc, (3.5)

where N cab and N t

rs are the fusion coefficients in A and T , respectively. This elemen-

tary constraint is surprisingly restrictive. For instance, it immediately leads us to

two constraints on the quantum dimensions of particles in the A and T theories (see

Appendix A.1)

da =∑r∈T

nradr, ∀ a ∈ A, (3.6a)

dt =1

q

∑a∈A

ntada, ∀ t ∈ T , (3.6b)

where q :=∑

a nϕada. Diagrammatically, Eq. (3.6b) is

(3.7)

It will also be useful to define the quantity

βt :=∑a∈A

θadanta, (3.8)

where θa is the topological spin of a ∈ A. Given a particle t ∈ U , it follows from the

aforementioned definition of a deconfined particle Eq.(3.4)

βt = qdtθt, ∀t ∈ U , (3.9)

as a useful corollary to Eq. (3.6b).

28

3.2 The Condensation Matrix Mab

So far, our formalism does not differ appreciably from that of Refs. [42] and [21]. How-

ever, in what follows we opt to not introduce the so-called “vertex lifting coefficients”

on which the approach of [21] is based. Instead, we find that we can extract a sur-

prising amount of information from supplementing the algebraic relations in Sec. 3.1

with two additional assumptions. First, by assumption, we are only interested in

cases where U is a TQFT, so that its anyons form a braided fusion category. Second,

we assume that

βt = 0, ∀t ∈ T /U , (3.10)

where t ∈ T /U runs over all confined anyons. To motivate this equation let us

pictorially represent the lefthand side of Eq. (3.10) as

(3.11)

where a particle t is braided around itself. This process is equivalent to braiding

the lifts a of the particle t (namely nta 6= 0). Each of these braidings is given by

the phase θa, while the loop with particle a is equal to the quantum dimension da.

The result we obtain is the quantity βt which we assume vanishes when t ∈ T /U as

confined particles cannot form a braided category. This is in contrast with Kirillov-

Ostrik [91], Kong [62] and Eliens et al. [21]. In these works, the authors present

the process of boson condensation as the identification of a commutative separable

subalgebra ϕ of A. The condensed theory T is identified as a module over φ living

in A. Using this formalism, which allows one to relate braiding processes in the

T theory to braiding processes in the original theory A, it is possible to show that

βt vanishes when t ∈ T /U . Using our stripped down algebraic formalism, we are

29

currently unable to interpret braiding processes in T in terms of those in A, and so

we are unable to mimic the procedures in the chapter above. As a result, we are

inclined to simply assume βt vanishes when t ∈ T /U . In certain special cases we

can show that Eq. (3.10) follows from the assumptions: in Sec. 3.1, e.g., we do so for

the so-called simple current condensates (see Sec. 3.4). It is so far unclear how to

generally prove Eq. (3.10) in our framework.

With these additional assumptions in place, we define some useful quantities. The

vacuum component t = ϕ of Eq. (3.5) will be a central object in our analysis, the left

hand side of which reads:

M ′ac :=

∑t∈T

ntantc

(=∑b∈A

N cabn

ϕb

), (3.12a)

as will be

Mac :=∑t∈U

ntantc . (3.12b)

Notice how the two above definitions of the matrices M ′ and M with nonnegative

integer entries differ subtly but crucially: The expression for M ′ involves a summation

over T while that for M involves a summation over the deconfined condensed theory

U .

The matrices, which can be factorized as in Eq. (3.12a) and (3.12b), are called

completely positive matrices over the ring of positive integers. We will discuss com-

pletely positive matrix factorization later in the chapter. In the following sections,

we will demonstrate two important properties of M , namely [M,S] = [M,T ] = 0,

where S and T are modular matrices of the A theory. For a discussion of the role of

the matrix M in CFTs, we refer the reader to Appendix A.2.

30

3.2.1 Proof that M commutes with T matrix of the A theory

In the following, we will prove that the M matrix we defined in Eq. (3.12b) commutes

with the modular T matrix of the A theory. The T matrix of the A theory is

Tab = θaδab. Note that

[M,T ]ac =∑b∈A

(MabTbc − TabMbc) = Macθc −Macθa =∑t∈U

ntantc(θc − θa). (3.13)

Since t ∈ U , the spins of all the lifts of t in A are the same, hence θa = θc and each

term in the final line vanishes identically. It follows that

[M,T ] = 0 . (3.14)

Note that this is not valid if the sum in the last line of Eq. (3.13) was not restricted

to the U theory, i.e., [M ′, T ] 6= 0, with M ′ defined in Eq. (3.12a).

3.2.2 Proof that M commutes with S matrix of the A theory

In this section, we will prove that the M matrix commutes with the modular S matrix

of A theory

[M,S] = 0 . (3.15)

We start from the expression of the S matrix for a braided fusion category A (e.g.,

see Ref. [36])

Scb =1

DA

∑x∈A

Nxcb

θxθcθb

dx, (3.16)

31

where DA is the total quantum dimension of the A theory and b denotes the antipar-

ticle of b. From the definition of M , we express the commutator [M,S] as

[M,S]ab =1

DAθaθb

∑c,x∈A

∑t∈U

θxdx(ntan

tcN

xcb − n

tbn

tcN

xac

)=

1

DAθaθb

∑x∈A

∑t∈U

θxdx

[nta

(∑c∈A

ntcNxcb

)− ntb

(∑c∈A

ntcNxac

)]

=1

DAθaθb

∑x∈A

∑t∈U

θxdx

[nta

(∑c∈A

ntcNcbx

)− ntb

(∑c∈A

ntcNcax

)].

(3.17)

In the first line we have used that if ntantc 6= 0 with t ∈ U , then θc = θa and if ntcn

tb 6= 0

with t ∈ U , then θc = θb which yields the term θaθb in the denominator. To obtain

the last line, we have used the equalities N cab = N b

ac = N cab

and N cab = N c

ba. We now

use Eq. (3.5) to replace the terms in the round brackets and find

[M,S]ab =1

DAθaθb

∑s∈T

∑x∈A

θxdxnsx

∑t∈U ;r∈T

(ntan

rbN

trs − ntbnraN t

rs

), (3.18)

where we have used the equality nsx = nsx (the assumption that the restriction of x’s

antiparticle x is the antiparticle of the restriction of x) to transfer the antiparticle on

the N tr,s. We can now split up the r sum in Eq. (3.18) into a sum over U and a sum

over T /U . For the first contribution, we have

∑r,t∈U

(ntan

rbN

trs − ntbnraN t

rs

)=∑r,t∈U

ntanrb

(N trs − N r

ts

)(3.19)

by exchanging the labels r and t in the second term. Since N rts = N t

rs = N trs,

Eq. (3.19) vanishes identically. Thus, r in Eq. (3.18) can only take values in T /U .

By assumption, U is a closed fusion category. This implies that no trivalent vertex

with a single leg in T /U exists in T . As a result, the s-sum in Eq. (3.18) may only

run over T /U . However, for s ∈ T /U we can use the assumption Eq. (3.10) to find

32

∑x∈A θxdxn

sx = 0 for the remaining terms in Eq. (3.18). We conclude that Eq. (3.18)

vanishes identically and thus [S,M ] = 0.

These equations are essential to the theory of condensation, as they establish

that the condensation matrix Mab is a particular symmetry of the S and T modular

matrices. While there exist other such symmetries, for example automorphisms that

are represented by permutation matrices, these matrices are not ‘completely positive’

integer matrices i.e., they cannot be factorized as nnT in terms of a nonnegative

integer matrix n.

3.3 The Modular Tensor Category after condensa-

tion

In the previous section, we identified a matrix M which commutes with the modular

matrices S and T of the A theory. In this section we prove a stronger pair of results,

namely that

nS = Sn, (3.20a)

nT = Tn, (3.20b)

where S and T are the modular matrices of the U theory and n is the matrix of coef-

ficients that enter the restriction and lifting maps, (n)at = nta, ∀ a ∈ A, t ∈ U . Our

assumption Eq. (3.10) will be crucial for these proofs. The second equality Eq. (3.20b)

is the statement that, component by component, whenever nta 6= 0, θt = θa; this is true

by recalling our definition of deconfined particles of the U (⊂ T ) theory, Eq. (3.4).

33

Before starting the proof of the first equality Eq. (3.20a), we note the following

equalities derived in the appendix of Ref. [92]

∑t∈U

βtβ∗t = q2D2

U (3.21a)

and ∑t∈T

βtβ∗t = D2

A. (3.21b)

It then follows from assumption Eq. (3.10), that∑

t∈U βtβ∗t =

∑t∈T βtβ

∗t and thus

q2D2U = D2

A or [91]

q = DA/DU , (3.22)

To prove Eq. (3.20a), we multiply Eq. (3.5) by θada and sum both sides over a ∈ A

to obtain

∑r,s∈T

N trsn

sbβr =

∑a,c∈A

N cabdaθan

tc = θbDA

∑c∈A

Sbcθcntc, (3.23)

where we have used the definition of S from Eq. (3.16) and that θa = θa. For particles

t ∈ U , we furthermore have

∑r,s∈T

N trsn

sbβr =

∑r,s∈U

N trsn

sbqθrdr, ∀t ∈ U , (3.24)

because (i) only r ∈ U contributes to the sum (as βr = 0 if r ∈ T /U) and (ii) only

s ∈ U contributes since U is closed under fusion by assumption. We furthermore

used Eq. (3.9) to rewrite the righthand side of Eq. (3.24). Since we assumed that U

forms a braided fusion category with its S matrix, we use the usual definition of the

S matrix to write ∑r,s∈U

(N trsθrdr

)nsb =

∑s∈U

SstDUθtθsnsb. (3.25)

34

Since s, t ∈ U , we furthermore have θs = θb if nsb 6= 0 which allows us to combine

Eq. (3.23) and Eq. (3.25) into

1

q

DADU

∑c∈A

Sbcθcntc =

∑s∈U

Sstnsbθt. (3.26)

Since for all ntc 6= 0, we have θc = θt (t ∈ U) and using Eq. (3.22), this expression

reduces to

∑c∈A

Sbcntc =

∑s∈U

Sstnsb. (3.27)

We have thus proven Eq. (3.20a) and Eq. (3.20b) within our algebraic formulation

of the condensation transition. These two equations have a well known parallel in

the study of chiral algebra extensions, which we detail in Appendix A.2. As a side

remark, let us derive a consequence of Eq. (3.22), namely

DA > DU , (3.28)

which follows from the fact that the embedding dimension q =∑

a∈A danϕa > 1 and

q = 1 if no condensation is happening. This is always true even if the assumption

Eq. (3.10) is not used. If Eq. (3.10) is used, then Eqs. (3.21a) and (3.21b) imply

DA = qDU . By analogy with the Zamolodchikov c-theorem of Ref. [93] one can call

this result the D-theorem which can be interpreted as the disappearance of some

anyons upon condensation. There is a stronger connection between this result and

the g-theorem, according to which the Affleck-Ludwig boundary entropy of an open

conformal system decreases under the renormalization group flow of the boundary

as long as the bulk theory remains critical throughout the flow.[94, 95, 96] (This

situation is, however, distinct from the case of a condensation transition in which the

bulk is only critical at the transition.) The boundary entropy is in turn related to the

35

quantum dimension of the primary field that characterizes the boundary condition,

which suggests a relation between the g-theorem and Eq. (3.28).

3.4 Simple currents

Simple currents are abelian anyons that, when raised to a certain power by fusion,

equal the identity, see Refs. [97, 98, 99, 100, 101, 102, 103]. The precise definition

follows below. In this section, we consider a condensate that is composed of simple

currents only. In this situation, we can prove that Eq. (3.10), i.e., βt = 0, ∀t ∈ T /U ,

follows from the assumptions in Sec. 3.1.

3.4.1 Introduction to simple currents

There are several equivalent definitions of simple currents in the context of rational

conformal field theory (RCFT). First, a simple current is a primary field J that has

a unique fusion channel with any other primary field of the RCFT

J × φ = φ′, ∀φ. (3.29a)

A second definition is that a simple current is a primary field J that when fused with

its antiparticle or conjugate field J only fuses to the identity (see Ref. [99])

J × J = 1. (3.29b)

A third definition is that the quantum dimension of J is 1,

dJ = 1. (3.29c)

One can show that all these definitions are equivalent. [98]

36

Given two simple currents J1 and J2, their fusion product J1J2 is also a simple

current. The number of primary fields of a RCFT is finite, therefore each simple

current J has an associated integer N such that JN = 1 by using Eq. (3.29a). The

smallest integer N > 0 with this property is called the order of J . A simple current

J generates a set of simple currents {Jm|m = 0, 1, . . . , N − 1}, which is isomorphic

to the abelian group ZN . A RCFT may contain simple currents generated by more

that one primary field. The collection of all of them form an Abelian group which

is isomorphic to the product ZN1 × · · · × ZNr . One can choose a basis of simple

currents such that Ni are of the form pnii , ni ∈ Z, with pi a prime number. This is

the fundamental theorem of finite abelian groups.

As an example, consider the RCFT constructed from the Kac-Moody algebra

SU(2)k. The primary fields are denoted by φ` where ` = 0, 1, . . . , k is twice the

topological spin. The field φk is a simple current because its fusion rule is φk × φ` =

φk−`. Indeed φk is the only non-trivial simple current, and satisfies φk×φk = φ0 = 1.

The simple currents form a Z2 = {φ0, φk} subcategory.

When acting on a primary field φ, J generates an orbit formed by the fields Jnφ

[φ] = {φ, Jφ, J2φ, . . . , Jd−1φ}, Jdφ = φ. (3.30)

Here, d is the smallest positive integer such that Jdφ = φ. The orbit Eq. (3.30) is

denoted by a representative field φ but one can choose another field belonging to

the orbit. In general d need not equal N , the order of the current J , but d must

divide N . In the example of SU(2)k, if k is odd, all the orbits have two elements,

(d = N = 2), while for k even, φk×φ k2

= φ k2

for the action of φk and so the orbit has

only one element, φ k2. Generally, we will simply call the anyon a “fixed point” when

it is invariant under fusion with a simple current, or equivalently, if its orbit contains

only the anyon itself. In the SU(2)k (k even) example, φ k2

is a fixed point under the

37

fusion with φk. As we shall see below, the existence of fixed points is crucial for the

construction of the condensed theory.

3.4.2 Simple current condensation

We consider a condensation transition, where the condensate consists only of the

set of bosonic simple currents, generated by n simple currents J1, . . . , Jn with orders

N1, . . . , Nn. Any anyon in the condensate can thus be represented as J i11 . . . J inn , where

il = 0, . . . , Nl − 1 and l = 1, . . . , n (note that the fusion product of simple currents is

unique.) We use the shorthand notation i = (i1, . . . , in) and

Ji := J i11 · · · J inn . (3.31)

The initial theory might have simple currents which are not bosons. We do not

consider these, as they cannot condense. We consider the group generated by the

powers of all the bosonic simple currents, which is sometimes called the bosonic cen-

ter C of the RCFT. Powers of a condensed bosonic simple current, or the products of

different condensed bosonic simple currents are also bosonic and condensed. To see

this, examine the Ji, Jj component of Eq.(3.5), where Ji, Jj are assumed to be con-

densed. Recalling that Ji+j := Ji × Jj is automatically a simple current (see above),

and making use of the fact that condensed simple currents like Ji have quantum

dimension 1 so that ntJi = δtϕ, we find

1 = nϕJinϕJj

= NJi+j

Ji,JjnϕJi+j

= nϕJi+j. (3.32)

As a result nϕJi+j= 1, indicating that Ji+j restricts solely to the vacuum, so it must be

a boson. Therefore the product of any two condensed simple current is a condensed

(hence bosonic) simple current.

38

As an aside, we note that for general bosonic currents which are not necessarily

condensed: (i) as before, any power of such a simple current is a bosonic simple

current; (ii) however, the product of two such bosonic simple currents does not have

to be bosonic. For example, in the toric code that we will discuss in detail in Sec. 3.6,

e and m are bosonic simple currents while their product f is actually a fermion (which

cannot condense). To prove (i), one can use the symmetry Sab = S∗ab

of the S matrix

and the fact that for any anyon θa = θa. Choosing a = b = J with J a simple current

gives

θJ2

θJθJ=

θ1

θ∗Jθ∗J

. (3.33)

If J is bosonic, then so is J and the above implies J2 is also a bosonic simple current,

i.e., θJ2 = 1. This argument can be iterated by assuming that up to some n0 all Jn,

n = 1, . . . , n0, are bosonic (then so are all JN−n, n = 1, . . . , n0, with N the order of

J). Solving the equality SJ,Jn0 = S∗J,JN−n0

for θJn0+1 yields that Jn0+1 is also bosonic.

Vafa’s theorem

We first aim to find information about the topological spins of some of the particles

in the theory by analyzing the implications of Vafa’s theorem [104]

∏p

(θpθxθy

)NpxyN

upz∏

q

(θqθxθz

)NuyqN

qxz

=∏r

(θuθxθr

)NryzN

uxr

(3.34)

∀x, y, z, u. For the case of the simple current condensate, we pick a particle x = a, a

particle y = Ji and a particle z = Jj . Note that a can be any particle in the A theory,

not necessarily a simple current. This choice of the anyons uniquely fixes all other

anyons in the equation (p = a × Ji, u = a × Ji+j , q = a × Jj , r = Ji+j). Using the

39

fact that the simple currents and their powers are all bosons, Vafa’s theorem gives

θa×Jiθa

θa×Jjθa

=θa×Ji+j

θa. (3.35)

This equation implies that the fractions θa×Ji/θa are irreducible characters of the

group ZN1 ⊗ ZN2 ⊗ . . . ⊗ ZNn , the bosonic center of RCFT which condenses. The

one-dimensional characters of this group can be written as

θa×Jiθa

= ωi11 ωi22 ...ω

inn , (3.36)

where ωi’s satisfy ωNii = 1. Also note that the ωi’s secretly depend on the subindex

a. There are two cases:

Case 1 : θa×Ji/θa = 1, (3.37a)

Case 2 : θa×Ji/θa 6= 1. (3.37b)

In the latter case, if particles a and a× Ji restrict to the same particle t ∈ T

then this particle is confined (as θa×Ji 6= θa). Moreover, from the orthogonality of

characters Eq. (3.36) we know that in this case

N1,N2,...,Nn∑i1,i2,...,in

θa×Jiθa

=

N1−1∑i1=0

ωi11

N2−1∑i2=0

ωi22 ...

Nn−1∑in=0

ωinn

= 0.

(3.38)

This happens when at least one ωil is not equal to 1.

40

Condensation

Without loss of generality, we will assume J1, J2, . . . , Jn condense. If only a subset of

the simple currents condense then the same analysis applies to just the bosons that

condense (the others factor out). Since dJ1 = . . . = dJn = 1, the bosons restrict only

to the new vacuum ϕ with coefficients unity

nϕJ1 = . . . = nϕJn = 1 (3.39)

and do not split. Using the reasoning in Sec. 3.4.2 it follows that all products of these

simple currents also condense – indeed, all bosonic simple currents Ji condense.

We will now proceed to prove a few crucial lemmas for any a, b ∈ A: (i) nta =

nta×Ji , ∀i, for all t ∈ T and (ii)∑

t ntan

tb 6= 0 if and only if b = a× Jj for some j.

(i) is easily proved by examining the b = Ji component of Eq. (3.5). To show

(ii) we examine the t = ϕ component of Eq. (3.5) and note that all bosonic simple

currents condense giving

∑t∈T

ntantb =

N1,N2,...,Nn∑i1,i2,...in

N ba,Ji

nϕJi

=

N1,N2,...,Nn∑i1,i2,...,in

δb,a×Ji .

(3.40)

for any a, b ∈ A. To prove (ii), note that:

• If b 6= a × Jj for all j, then∑

t∈T ntan

tb = 0 and particles a, b do not have any

common restrictions. Let us write this result as

If b /∈ [a] =⇒ ntantb = 0, ∀t, (3.41)

where [a] = {Jja, Jj ∈ C} is the orbit obtained acting on a with all the bosonic

simple currents.

41

• If b = a× Jj for some j then

∑t∈T

ntanta×Jj =

N1,N2,...,Nn∑i1,i2,...in

δa×Jj ,a×Ji

=Ra ∈ Z+.

(3.42)

But from (i), nta×Jj = nta, so the LHS of this equation is positive. Hence Ra > 0.

For example, for n = 1, Ra = N1/d, with d defined in Eq. (3.30).

Hence we have proved (ii). From (ii) we know that if a and b are in the lift of t then

b = a× Jj for some j. On the other hand from (i), if a is in the lift of t, so is a× Jj .

Hence t is deconfined iff θa = θa×Jj for all j, where a is any particle in the lift of t.

In other words, given an a ∈ A, the character θa×Jj/θa 6= 1 for some j iff a restricts

only to confined particles.

Let us now prove the assumption (3.10). We first multiply Eq. (3.40) by dbθb and

sum over all particles b in the A theory to obtain

∑t∈T

βtnta =

N1,N2,...,Nn∑i1,i2,...,in

da×Jiθa×Ji

= daθa

N1,N2,...,Nn∑i1,i2,...,in

θa×Jiθa

,

(3.43)

where we have used the fact that the quantum dimension of any product of a particle

with simple currents remains the same. Now if θa×Ji/θa is not the identity character

of the trivial representation, then the particle a restricts only to confined particles

and we have from Eq. (3.38)

∑t∈T

βtnta =

∑t∈T /U

βtnta = 0. (3.44)

42

In fact, the second equality holds even if θa×Ji/θa = 1 for all j, because in that case

nta = 0∀t ∈ T /U – as a result, the second equality holds for all a. Multiplying the

second equality by θ∗ada and summing over all particles a we obtain

0 =∑t∈T /U

βt∑a∈A

θ∗adanta =

∑t∈T /U

βtβ∗t . (3.45)

The unique solution is βt = 0 for t confined, coinciding with our assumption (3.10).

3.5 One confined particle theories

In this section we study a simple boson condensation with just one confined particle

t0 in the T theory. Furthermore, assume that the confined particle t0 has only two

lifts a1 and a2 with lifting coefficients both 1, i.e.,

nt0a1 = nt0a2 = 1, (3.46)

otherwise nt0a = 0,∀a 6= a1, a2. (3.47)

With these assumptions, we can prove that the condensate has only one boson

besides vacuum, and this condensed boson has quantum dimension 1. This implies

that the boson is a simple current, so the results of the previous section imply βt0 = 0.

However, we choose to prove this equation through another method which gives more

information about bosonic condensation theories with only one confined particle.

Further we find that dt0 = da1 = da2 , which means that a1 and a2 only restrict to one

particle t0 in the T theory, with no other particles in T . Finally, in this special one-

confined particle case, we prove that βt0 :=∑

a∈A nt0a daθa = 0 which clearly support

the assumption we used in previous sections. The detailed proof can be found in the

appendix of Ref. [92].

43

3.6 Formalism and implementation

We now present an algorithmic prescription, which can be implemented on a com-

puter, and which strongly contrains the possible condensation transitions starting

from a TQFT with given modular matrices S and T . We then apply this procedure

to several example TQFTs. The algorithm is performed in 3 steps:

1. Search for the symmetric matrices M with nonnegative integer entries and

M1,1 = 1 satisfying

[M,S] = [M,T ] = 0. (3.48)

2. For each M , find all nonnegative integer rectangular matrices n such that M =

nnT.

3. For each M and n, find the putative modular matrices S and T of the TQFT

after condensation by solving

Sn = nS and Tn = nT . (3.49)

One subtlety is that we need to make sure that the S, T matrices we obtain

are valid. In this chapter, we use the necessary conditions for a valid S matrix:

it should be symmetric, unitary, and it should generate non-negative fusion

coefficients by Verlinde formula. These are always satisfied if U is a MTC.

This algorithm sidesteps the discussion of the theory T that contains confined anyons

and directly yields the resulting MTC U formed by the remaining deconfined anyons.

The algorithm provides all condensation solutions of theory A. Another algorithm

which does not sidestep T is to (1) build the matrixM ′ as in the bracket of Eq. (3.12a);

(2) factorize it in nta; (3) keep only the deconfined particle t’s; and then apply step (3)

and Eq. (3.49). Whether the two theories are identical hinges on Eq. (3.10), which

we assume to be true. We now address the above steps one by one.

44

3.6.1 Solutions for M

Since T is diagonal, the equation [M,T ] = 0 is satisfied if and only if M is a block

diagonal matrix with nonzero off-diagonal entries only between particles with the same

topological spin. Imposing this block structure, we can solve [M,S] = 0, imposing

that

1. Mab =Mba ≥ 0, M ∈ Z,

2. M11 =1.

(3.50)

The second condition ensures that the A vacuum restricts to the vacuum ϕ of U . In

this case, the first row (or column) of M is equal to the first column of n and describes

the particles that condense into the vacuum nϕa . (From this, it is also clear that only

solutions with M1a ≤ da can lead to a valid theory.)

With conditions 1 and 2 in Eq. (3.50) in place, we obtain two types of solutions for

M , which we call automorphisms and condensations, aside from the trivial solution

M = 1.

Automorphisms are defined by a fully-ranked matrix M satisfying

∑a

Mab = 1 ∀b. (3.51)

They satisfy M2 = 1 because of the following reasons: Since∑

aMab = 1 and all

entries of M can only be nonnegative integers, for any b ∈ A, there is only one

corresponding particle b′, such as Mb′b = 1. Further, Mab = 0, ∀a 6= b′ and M is

fully-ranked. As a result, if a 6= b then a′ 6= b′. Hence (M2)ab =∑

cMacMcb =∑cMacMbc = δab. An automorphism M is thus a permutation matrix of order two –

it is a symmetry of the S, T data under relabeling of particles. All automorphisms of

A form a group under matrix multiplication, which is used to construct “topological

symmetry group” in the presence of a global symmetry. [105] Automorphism, however,

45

still exists even when any other symmetries (e.g. U(1) charge conservation), are

broken.

On the other hand, solutions M that correspond to a condensation have M1a 6= δa,1

for some a, implying that at least one other boson besides the vacuum restricts

to the new vacuum. All the condensations can be superimposed with any of the

automorphisms, yielding a potentially different condensation. In other words, two

condensations can be related via a permutation of A by multiplying the M matrix of

one condensation from both sides with the M matrix of the automorphism – we will

see an example of this below for the toric code TQFT.

We can prove that any M that satisfies Eq. (3.48) and conditions 1 and 2 in

Eq. (3.50) is either an automorphism or a condensation as follows. We first assume

that M is not a condensation solution, that is, the first row and column of M are

all zeros (M1a = Ma1 = 0, ∀a 6= 1) except M11 = 1 . We show that M must be an

automorphism in this case. From

∑b

MabSbc =∑b

SabMbc (3.52)

we have for c = 1 ∑b

MabSb1 = Sa1 ⇒∑b

Mabdb = da. (3.53)

Thus, da is a strictly positive eigenvector of M with eigenvalue 1. Since every Mab is

integer and larger or equal to zero, Eq. (3.53) can only hold if

fa ≡∑b

Mab ≥ 1. (3.54)

On the other hand, summing Eq. (3.53) over a, and using M = MT, gives

∑b

fbdb =∑a

da. (3.55)

46

Again, since fa ≥ 1 and the da are strictly positive this equation can only be satisfied

if

fa ≡∑b

Mab = 1, (3.56)

which, together with the fact that M is symmetric, implies that M has to be an

automorphism (a permutation matrix).

Let us illustrate how automorphism and condensation solutions for M arise from

condition (3.48) for the example of the toric code (TC) TQFT. It contains the anyons

1, e, m, f and has the modular matrices

STC =1

2

1 1 1 1

1 1 −1 −1

1 −1 1 −1

1 −1 −1 1

(3.57a)

and

TTC = diag(1, 1, 1,−1). (3.57b)

It admits three nontrivial solutions to Eq. (3.48), one automorphism

M (1) =

1 0 0 0

0 0 1 0

0 1 0 0

0 0 0 1

(3.58)

47

that exchanges the e and the m particles and two condensations

M (2) =

1 1 0 0

1 1 0 0

0 0 0 0

0 0 0 0

, M (3) =

1 0 1 0

0 0 0 0

1 0 1 0

0 0 0 0

, (3.59)

of either the e or the m boson. They are related by the automorphism M (2) =

M (1)M (3)M (1) [note that (M (1))−1 = M (1)].

3.6.2 Solutions for n

Next we solve for the integer matrix nta ≥ 0, where t labels the deconfined particles in

the MTC U . It is possible that multiple solutions n exist for a given M . However, for

some solutions, it still might not be possible to find a valid condensed MTC: please

refer to our Appendix A.3.2 for an example of 4-layer Ising model condensation. In

that example, we obtain unitary S and T matrices, but they do not correspond, via

Verlinde’s formula, to integer fusion coefficients.

An efficient first step in solving for n is to realize that any column of M that only

contains zeros and ones is equal to a column in n. While the matrix M may contain

several columns with only zeros and ones that are equal, they all correspond to only

a single column in n (there are no duplicate columns in n). After removing from

M all rows and columns that contain only zeros and ones, an actual factorization

routine can be used on the remaining sub-block of the M matrix. (As we will discuss

for an example below, the factorization does not always yield a unique solution for

n in this case.) In the situations we have encountered, this part of the algorithm is

not limited by computational power. In the particularly simple toric code example,

48

deleting duplicate columns directly yields the solution

M (2) = nnT, nT = (1, 1, 0, 0). (3.60)

There is only one particle in the new theory, the vacuum. Thus, condensation of

either the e or the m particle in the toric code yields the trivial TQFT.

As a less trivial example, consider a bilayer of Ising TQFTs. Each layer contains

the anyon types 1, σ, ψ with modular matrices

SI =1

2

1

√2 1

√2 0 −

√2

1 −√

2 1

, TI = diag(1, eiπ/8,−1). (3.61)

The bilayer S and T matrices are direct products SI(2) = SI⊗SI, TI(2) = TI⊗TI, and the

theory supports 9 particle types which we denote 11, 1σ, 1ψ, σ1, σσ, σψ, ψ1, ψσ, ψψ,

where 11 is the vacuum. There is only one nontrivial solution for M , which reads in

this basis

M =

1 0 0 0 0 0 0 0 1

0 0 0 0 0 0 0 0 0

0 0 1 0 0 0 1 0 0

0 0 0 0 0 0 0 0 0

0 0 0 0 2 0 0 0 0

0 0 0 0 0 0 0 0 0

0 0 1 0 0 0 1 0 0

0 0 0 0 0 0 0 0 0

1 0 0 0 0 0 0 0 1

. (3.62)

49

It is straightforward to obtain the unique solution n that yields M = nnT

nT =

1 0 0 0 0 0 0 0 1

0 0 0 0 1 0 0 0 0

0 0 0 0 1 0 0 0 0

0 0 1 0 0 0 1 0 0

, (3.63)

which shows that this describes the condensation of the ψψ particle. In this process,

the σσ particle (which has quantum dimension 2) splits into two particles of quantum

dimension 1 and both 1ψ, ψ1 restrict to the same particle. All other particles, except

for the vacuum, become confined.

There exists M that solve Eq. (3.48), but cannot be decomposed as M = nnT

with a nonnegative integer matrix n. Some of them still admit an interpretation in

terms of a condensation in the following sense. If the MTC U that is obtained from a

condensation with matrix M = nnT has an automorphism symmetry P , that is equal

to its transpose P = PT, then M = nPnT is also a symmetric matrix that solves

Eq. (3.48). For instance, one necessary condition for a decomposition M = nnT to

be possible is that Maa +Mbb ≥ 2Mab. If the matrix elements of M do not satisfy the

triangle equation Maa +Mbb ≥ 2Mab, then M = nPnT might be possible instead.

If the TQFT corresponds to a CFT, the possible forms of matrices M that solve

Eq. (3.48) are understood with the help of the “naturality theorem” by Moore and

Seiberg [106, 107]. This theorem implies that all M that solve Eq. (3.48) in a CFT

are either automorphisms of A, condensations of the form M = nnT, or of the form

M = nPnT, with P an automorphism of U . As a corollary, we then conclude that

for any solution to Eq. (3.48) of the from M (2) = nPnT, there is another solution

M (1) = nnT, with the same n, since the identity mass matrix of U always exists. For

the purpose of studying condensations, we thus focused on matrices M that admit

the decomposition M = nnT throughout our analysis. If we relaxed this constraint to

50

also include M = nPnT, assumptions such as nta = nta would not be justified anymore.

We discuss the interpretation of condensation transitions for CFTs in Appendix A.2

and relate it to the “naturality theorem”. Subsequently, in Appendix A.3 we give an

example of condensation transitions in SU(2)16, for which two solutions M = nPnT

and M = nnT to Eq. (3.48) exist.

The decomposition M = nnT is generally not unique. For example, if Maa = 4 for

some particle with quantum dimension 4 or larger, it can either split in 4 particles

with nta = 1 for each or restrict to one particle with nta = 2 (this issue was discussed

previously in Sec. 3.4). However, in all examples we studied, at most one of all

possible decompositions of M lead to a consistent TQFT with valid solutions for S

and T . Thus, the uniqueness of this step in the condensation is an open question.

We mentioned that factorizing M = nnT is a well-known problem in the field of

completely positive matrices. In our cases, the factorization happens over the ring of

positive integers. This problem is known to be NP-hard. With the exception of small

dimension matrices, it has not yet been solved. Some outstanding questions are the

characterization of when a matrix M is completely positive (sufficient and necessary

condition), as well as what is the minimal number of rows in n (called CP rank),

which is translated in our case to the minimal number of particles in U that can be

obtained.

3.6.3 The modular matrices of the new theory

Having obtained the matrix n, we now solve the equations

Sn = nS, Tn = nT (3.64)

for S and T . These equations can have spurious solutions unless we impose a list of

additional constraints. For modular theories, these constraints are

51

• S† = S−1,

• S2 = Θ(ST )3 = C, where C is a permutation matrix that squares to the

identity and Θ = e−iπc/4 with c the chiral central charge of A, which we can

prove remains unchanged (mod 8) during condensation.

• T is a diagonal matrix with complex phases on the diagonal,

• the fusion coefficients obtained from the Verlinde formula

Nt = SDtS−1, (3.65)

with (Dt)rs = δr,sStr/S1r have to be nonnegative integers.

We do not prove that any solution that obeys the above list of conditions is indeed

a valid MTC U . However, any allowed condensation will be a solution to these

conditions. Therefore, if we do not find a solution for a given MTC A, we can

conclude that no condensation transition to a modular U theory out of A exists (we

will discuss a nontrivial example for this situation in Sec. 3.7.2).

For the example of the double layer Ising theory, we have for SI(2)n = nS (skipping

columns of zeros, which correspond to the confined particles)

12

12

1 12

12

12−1

20 −1

212

12−1

20 −1

212

12

12−1 1

212

=

S11 S14 S12 + S13 S14 S11

S21 S24 S22 + S23 S24 S21

S31 S34 S32 + S33 S34 S31

S41 S44 S42 + S43 S44 S41

. (3.66)

Note that all |Sab| = 1/2 since the theory contains only Abelian anyons. Thus,

Eq. (3.66) determines all matrix elements of S, except for S22 = −S23 = −S32 =

S33. (The equality −S23 = −S32 follows from the fact that a modular S matrix is

symmetric.) At the same time, we have from Tn = nT that θ1 = 1, θ2 = θ3 = eiπ/4,

52

θ4 = −1. Furthermore, the (2,2) component of the equation S2 = Θ(ST )3 reads

1

2

(1 + 4S2

22

)=

1

2

(1 + 8iS3

22

)(3.67)

yielding the unique solution S22 = −i/2, that also satisfies |S22| = 1/2. We can use

the thus obtained S matrix to compute the fusion coefficients from Eq. (3.65), and

we find that they are all non-negative integers. The new fusion rules are

2× 2 = 3× 3 = 4, 2× 3 = 1, (3.68)

which are distinct from the toric code fusion rules. The resulting TQFT coincides

with the gauged Chern number 2 superconductor from Kitaev’s 16-fold way [20]. We

have thus shown that this TQFT is obtained in a unique way through condensation

in a double layer of Ising theories (two gauged Chern number 1 superconductors).

In fact, one can iterate this procedure to obtain all TQFTs appearing in Kitaev’s

16-fold way. A natural open question is to find out which TQFTs exhibit such a

closed structure with unique condensations. Using the formalism developed above,

we will show below that another simple non-Abelian TQFT, the Fibonacci category,

does not admit a similar structure, since it does not allow for any condensation.

One may wonder whether Eq. (3.65) needs to be imposed as a separate condition

on the possible solutions for S, or whether it follows from the other conditions in

the above list. To show that Eq. (3.65) is required, we discuss the example of four

layers of Ising TQFTs in Appendix A.3.2, for which there exist a unitary and sym-

metric S matrix, except that the fusion coefficients generated from S by Verlinde’s

formula in Eq. (3.65) are not integer. Therefore, it does not correspond to an allowed

condensation transition and the list of conditions is not complete without Eq. (3.65).

53

3.7 Layer constructions and uncondensable bosons

In this section, we apply the condensation formalism to TQFTs A(N) that are tensor

products of N identical layers of a TQFT A. There are several motivations to study

such a construction:

(1) Some TQFTs are characterized by a Zm grading under layering: N = m

layers can be physically equivalent to the trivial TQFT in the bulk. For a theory to

be condensable to nothing, m is constrained by the fact that the chiral central charge,

which is conserved under condensation, must vanish (mod 8). Condensation provides

a way to determine the grading m as well as all the TQFTs for N = 1, . . . ,m layers.

See the following Sec. 3.7.1 for discussions and details of examples.

(2) The grading of TQFTs has an immediate physical implication: Kitaev’s 16-

fold way, which we discuss below, characterizes 16 different chiral superconductors in

(2+1) dimensions.

(3) Abelian bosons are always condensable in any A. However, there exists a

non-Abelian boson b in an A which cannot be condensed. Furthermore, there exists a

non-Abelian boson b in an A, such that any tensor product of the non-Abelian boson

b(n), ∀ n = 1, 2 . . . , N cannot be condensed in A(N), ∀ N . For instance, the Fibonacci

category studied in Sec. 3.7.2. This phenomenon will be studied more systematically

in Chapter 4.

(4) Layer constructions have been proposed to gain insight into (3+1)-

dimensional phases with topological order, for which there is currently no systematic

understanding[108]. The idea is to couple N layers of a TQFT A by a condensation

transition in such a way that the number of anyons after condensation does not scale

with N . Some of the anyons that restrict to deconfined particles have a nontrivial

particle in every layer. Their restriction is then interpreted as a string excitation of

the (3+1)-dimensional theory. We discuss an example in the following Sec. 3.7.1.

54

Before condensation, the general structure of A(N) is

S ′A(N) = SA ⊗ · · · ⊗ SA︸ ︷︷ ︸N times

, T ′A(N) = TA ⊗ · · · ⊗ TA︸ ︷︷ ︸N times

, (3.69)

for the modular matrices and

N ′ca,b =N∏i=1

N ciai,bi

, d′a =N∏i=1

dai , θ′a =N∏i=1

θai , (3.70)

for the fusion matrices, quantum dimensions, and topological spins. Here, N ca,b, da

and θa, are the fusion coefficients, quantum dimensions, and topological spins of A

and the respective primed quantities belong to A(N). We have labeled the anyons

in A(N) by a vector a = (a1, · · · , aN)T of anyons in each layer, 1, . . . , N , where each

entry ai can be any of the anyons in A.

3.7.1 Theories with Zm-graded condensations

SU(3)1: 4-fold way

As a simple example, let us consider the SU(3)1 TQFT. It has three Abelian anyons

1, 3, 3 with fusion rules

3× 3 = 3, 3× 3 = 3, 3× 3 = 1 (3.71)

and topological spins θ3 = θ3 = ei2π/3. Now, we consider multiple layers of SU(3)1.

Notice that since each layer has a automorphism symmetry 3 ↔ 3, all statements

below should be understood modulo this automorphism symmetry applied to every

layer.

Clearly, the m = 2 layer theory SU(3)1 × SU(3)1 has no bosons and therefore no

condensation transition is possible.

55

The m = 3 layer theory SU(3)1 × SU(3)1 × SU(3)1 has 8 bosons. However, up to

the automorphism, there is a unique condensation corresponding to bosons (1, 1, 1),

(3, 3, 3) and (3, 3, 3) restricting to the vacuum 1′ and all other bosons confined. Besides

the vacuum, two more particles are deconfined: 3′ with lifts (3, 3, 1), (1, 3, 3), (3, 1, 3),

and 3′ with lifts (3, 3, 1), (1, 3, 3), (3, 1, 3). Together, 1′, 3′, and 3′ furnish SU(3)1,

which differs from SU(3)1 by complex conjugation of the topological spins. It might

seem unusual that the condensation of multiple layers of chiral theories results in an

anti-chiral theory, but we remind the reader that the chiral central charge is only

conserved modulo 8 under condensation transitions and hence −2− 2− 2 = 2 mod 8

is allowed.

Then, the m = 4 layer theory is SU(3)1 × SU(3)1 which can be condensed to

the trivial TQFT by condensing simultaneously (3′, 3) and (3′, 3), which confines all

other particles. We have thus shown that condensation induces in a unique way a Z4

grading in the layered SU(3)1 TQFTs.

Ising: Kitaev’s 16-fold way

We want to couple N layers of the Ising TQFT, which is defined in Eq. (3.61). For

condensation, the simplest boson that we can build consists of the ψ-particles in two

consecutive layers n+ 1 and n+ 2,

Bn := (1n, ψ, ψ, 1N−n−2), (3.72)

where 1n stands for the vacuum particle in n consecutive layers. All bosons Bn,

n = 0, ..., N − 2, are condensed. We will identify all bosons of this form with the

vacuum, building a simple current condensate. From

(1n, ψ, ψ, 1N−n−2)× (· · · , ψ, 1, · · · ) = (· · · , 1, ψ, · · · ), (3.73)

56

we see that consistency requires that any pair of anyons (· · · , ψ, 1, · · · ) and

(· · · , 1, ψ, · · · ) restrict to the same anyon after condensation. Here, · · · stands

for any sequence (that agrees between the two particles). Furthermore, by fusion

with the condensate we have

(1n, ψ, ψ, 1N−n−2)× (· · · , σ, 1, · · · ) = (· · · , σ, ψ, · · · ). (3.74)

However, θ(···σ1··· ) = −θ(···σψ··· ), implying that the restrictions of (· · ·σ, 1 · · · ) are con-

fined, because they have another lift (· · ·σ, ψ · · · ) with different topological spin. By

that argument we have shown that the set Q consisting of particles with least one

and at most N − 1 σ’s restricts only to confined particles. On the other hand, we

know that particles containing no σ’s (i.e., only 1’s or ψ’s) restrict to single deconfined

particles:

• By closure of the condensate, any particle with even number of ψ and otherwise

1 restricts to the new vacuum 1′.

• Any particle with odd number of ψ and otherwise 1 restricts to the deconfined

particle ψ′. Their fusion rule is

ψ′ × ψ′ = 1′. (3.75)

The only particle left to consider is σ(N) ≡ (σ, . . . , σ). It is easy to show that

σ(N) × Q ⊆ Q. It then follows from the a = σ(N), b ∈ Q, t = ϕ component of

Eq. (3.5) that σ(N) and particles in Q restrict to disjoint sets of particles, because

the righthand side of Eq. (3.5) is zero in this case, as none of the particles in Q

restrict to the vacuum. But then the restriction of σ(N) cannot possibly contain

confined particles as those confined particles would have just a single lift σ(N), which

57

is impossible from the definition of confined particle. Hence σ(N) restricts only to

deconfined particles, and we can identify Q as the set of lifts of all confined particles.

We can say more about the restriction of σ(N). Note DU = DA/q = 2N/2N−1 =

2, because q is equal to the number of condensed bosons i.e., q = 2N−1. As we

already know 1′, ψ′ are deconfined, DU =√

1 + 1 + . . . = 4, where . . . are additional

contributions from the restriction of σ(N). When N is not a multiple of 8, there are

just two options. Either case (1) (σ, · · · , σ) splits into just two Abelian particles

distinct from 1′, ψ′, or case (2) (σ, · · · , σ) has a single restriction with quantum

dimension√

2. (When N is a multiple of 8 the σ-string is itself a fermion or boson

and could restrict to the ψ′ and the vacuum, respectively. However, by DU = 2 it is

not possible that ψ′ and the σ-string have a common restriction in the case where N

is an odd-integer multiple of 8 (since DU =√

3 in that case). The case where N is a

multiple of 16 will be discussed separately below.) Consider now from Eq. (3.64) the

matrix element that corresponds to any particle t in the restriction of (σ, · · · , σ) and

the identity in A,

nt(σ,··· ,σ) =

√2N

2dt, (3.76)

since we know from the discussion following Eq. (3.74) that t has only one lift,

(σ, · · · , σ).

From the condition that nt(σ,··· ,σ) is integer, we conclude that case (1) applies to

even N and case (2) to odd N . We now analyze the two cases separately.

Case: N odd According to Eq. (3.76), we have (σ, · · · , σ)→ 2(N−1)/2σ′. It follows

from the fusion rules of the original theory, i.e., from Eq. (3.5) by choosing a = b =

(σ, · · · , σ), that

σ′ × σ′ = 1′ + ψ′. (3.77)

58

Thus, 1′, σ′, ψ′ furnish the same (Ising) fusion algebra as 1, σ, ψ do in every layer. The

spin factors of the deconfined restrictions are given by

θ1′ = 1, θσ′ = e2πiν/16, θψ′ = −1, (3.78)

where ν = N mod 16 is an odd integer, for N is odd. We have thus obtained all

TQFTs with Ising fusion rules that appear in Kitaev’s 16-fold way.

Case: N even If N is even, Eq. (3.76) yields the restriction (σ, · · · , σ)→ 2N/2−1a′+

2N/2−1b′ with equal coefficients. To find the fusion rules for a′ and b′, we solve

Eq. (3.64). This leaves two possibilities

a′ × a′ = b′ × b′ = 1′, a′ × b′ = ψ′, (3.79)

a′ × a′ = b′ × b′ = ψ′, a′ × b′ = 1′. (3.80)

Here, Eq. (3.79) are the toric code fusion rules. Which of the two cases applies can

be determined from the equation S2 = Θ(ST )3 = C, by using the topological spins

θa′ = θb′ = e2πiN/16. (3.81)

For N = 2 mod 4 one finds the solution Eq. (3.80) and for N = 4 mod 4 one finds

the solution Eq. (3.79).

The case where N is a multiple of 16 has to be considered separately. The con-

densation described here leads to the toric code TQFT in which a′ and b′ are bosons.

We have shown above that the toric code can be condensed to the trivial TQFT by

condensing either a′ or b′ (which were called e and m before). Thus, in the case where

the σ string is a boson, two condensations are possible: one leads to the toric code

and in the other one, in which the σ string restricts in part to the vacuum, leads to

59

the trivial TQFT. The toric code is also the TQFT that was proposed to describe a

gauged s-wave superconductor without topological edge modes. [109]

Together, this Z16 grading represents Kitaev’s 16-fold way, yielding a (non-

)Abelian fusion category for the vortices of even (odd) layer length. From the point

of view of layer construction[108], we note that ψ′ is a point-like fermionic excitation

in 3D space, while σ′, a′ and b′ are to be interpreted as vortex or line-like excitations

in 3D, because their lift has a nontrivial anyon in each layer.

It is tempting to consider the topological orders that have been proposed in

Refs. [110, 111] as the possible symmetry-preserving gapped surface terminations

of time-reversal symmetric (3+1)-dimensional superconductors as another example

of a theory with Z16 grading under condensation. The topological index ν of the

bulk superconductor has been shown to be only meaningful mod 16 in the presence

of interactions. The ν = 1 surface topological order was proposed to be nonmodular

category SO(3)6, while that for ν = 2 is the so-called T-Pfaffian state. We do not

further elaborate on possible condensations in this theory here, as the focus of the

present work is on condensation in modular categories. However, if we were to apply

the formalism of Eq. (3.64) to this problem, none of the possible condensation tran-

sitions in a double layer SO(3)6×SO(3)6 would lead to the T-Pfaffian. Rather, one

can condense all bosons in SO(3)6×SO(3)6 to obtain the trivial nonmodular TQFT

{1, f} with only one Abelian fermion f .

3.7.2 Theories with Z-fold way: Fibonacci TQFT

Not every TQFT has a Zm-graded structure under condensation. The simplest

counter-example is the Fibonacci TQFT with the single nontrivial anyon τ and the

fusion rule

τ × τ = 1 + τ. (3.82)

60

It has topological spin θτ = ei4π/5 and quantum dimension dτ = φ, where φ =

(1 +√

5)/2 is the golden ratio.

First, we want to show that no condensation is possible in 5 layers of Fibonacci,

despite the presence of the boson (τττττ). We will show that there is no matrix M

that describes a condensation and satisfies Eq. (3.48). To see this, consider the (1,b)

component of the equation MSFib(5) = SFib(5)M ,

∑a

nϕa (SFib(5))a,b =1

(2 + φ)5/2

∑a

daMa,b. (3.83)

Observe that the righthand side is nonnegative for any b. Specializing to b =

(τ, 1, 1, 1, 1), we find the lefthand side

(2 + φ)−5/2(φ− nϕ(τττττ)φ

4), (3.84)

which is negative for any nϕ(τττττ) ≥ 1, i.e., for any condensation. Therefore, no

condensation transition is possible in 5 layers of Fibonacci (see Ref. [21] for an

alternative proof).

Second, let us show further that no condensation is possible in 10 layers of Fi-

bonacci. Besides the vacuum, there is a boson with a τ anyon in every layer, which

we denote by (10τ), and 252 =(

105

)bosons with τ anyons in exactly 5 layers.

Again, we will show that there is no matrix M that describes a condensation and

satisfies Eq. (3.48). To see this, we consider the (1,b) component of the equation

MSFib(10) = SFib(10)M , but this time for the choice b = (10τ). Up to an overall factor

of the total quantum dimension, the equation reads

nϕ1φ10 +

∑a∈5τ bosons

(−1)5φ5nϕa + (−1)10nϕ(10τ)

= nϕ(10τ) +∑

a∈5τ bosons

φ5Ma,(10τ) + φ10M(10τ),(10τ).

(3.85)

61

Using nϕ1 = 1, it simplifies to

0 = φ5(M(10τ),(10τ) − 1

)+

∑a∈5τ bosons

(nϕa +Ma,(10τ)

). (3.86)

We can see that Eq. (3.86) has no nontrivial solution: Since φ5 is irrational, the first

term needs to be zero on its own, which requires M(10τ),(10τ) = 1. This implies that

(10τ) does not condense, as it has noninteger quantum dimension and would therefore

have to split in order to condense. However, the second term in Eq. (3.86) is a sum of

nonnegative numbers that can only vanish if nϕa = 0, ∀a. Hence, none of the bosons

condenses.

In fact, one can show that no condensation is possible for any number of layers N

of the Fibonacci TQFT [112]. We will reformulate this proof much more easily using

the formalism developed in this chapter elsewhere in a way that also generalizes to

other TQFTs.

Obstructions against the condensation of bosons within our formalism can only

ever occur in theories that contain non-Abelian anyons. In Abelian theories, any

potentially condensing boson J is a simple current (of order d), and one can explicitly

construct a theory in wich J is condensed as follows: Form all the orbits [a] with

respect to J , as defined in Eq. (3.30). The orbit of the identity is the condensate. If

all anyons in an orbit [a] have the same topological spin, the orbit labels a particle t[a]

in the theory U , otherwise all particles in the orbit are confined. If t[a] is unconfined,

choose nt[a]b = 1 if b ∈ [a] and n

t[a]b = 0 otherwise. Further, choose St[a],t[b] = dSa,b and

Nt[c]t[a],t[b]

=d∑

n=0

N c×Jna,b , (3.87)

for t[a], t[b], and t[c] unconfined. In can be readily shown that this choice is a consis-

tent solution to Eqs. (3.64) and (3.65) and therefore a valid condensation within our

formalism.

62

3.8 Conclusions

In summary, we derived a framework for the condensation of anyons that is applicable

to modular tensor category models of topological order. Our derivation is based

on a small number of physical assumptions and focuses on the computation of the

modular matrices S and T of the theory after condensation. Based on this, we propose

an algorithm to carry out this computation. This algorithm first seeks symmetric

nonnegative integer matrices M that commute with the modular matrices S and

T of the original theory. It then proceeds by factorizing M = nnT in a product

of a nonnegative integer matrix n with itself. Finally, the equations Sn = nS and

Tn = nT are solved. Our algorithm has proven to be practically useful in all examples

that we studied. We finally demonstrated that the equations that are central to our

derivation are powerful constraints on condensation transitions in general.

This leads us to several open problems that are not answered by the present

work. One concerns the assumption that βt = 0 for all confined particles t. We have

shown in Secs. 3.4 and 3.5 that this relation follows from weaker assumptions for

certain theories. But a general proof of this statement is lacking, so that it remains

an assumption for us. Other questions concern the uniqueness of solutions and the

transitivity of condensation transitions. For example, given an M , is there a unique

n that solves M = nnT and leads to a valid condensed theory? And given such a

solution n, is there a unique consistent solution S and T? In a similar vein, is the

condensed theory completely characterized by the coefficients nϕa?1 At present, we do

not have counterexamples against affirmative answers to these questions.

Another future direction could be the condensations in the presence of global

symmetries[105]. When we have global symmetries on top of a topologically or-

dered system, the anyons may transform in a projective representation. A direct

1Indeed, we cannot exclude the possibility that additional information, like certain vertex liftingcoefficients, are needed to fully determine the topological order of the condensed phase.

63

consequence is that certain condensations may not be able to happen if all global

symmetries are respected.

64

Chapter 4

No-Go Theorem for Boson

Condensation in Topologically

Ordered Quantum Liquids

One motivation to study condensation transitions is to classify topological order. An

important example are the 16 types of gauged chiral superconductors introduced by

Kitaev [20]. Kiteav showed that while two-dimensional superconductors are classified

by an integer Z, only 16 bulk phases are topologically distinct. This construction

can be understood by considering ` layers of initially disconnected chiral p-wave su-

perconductors, i.e., elementary (Ising) TQFTs. Upon introducing generic couplings

between these layers, one obtains a single layer of a chiral `-wave superconductor,

which corresponds to a specific TQFT in Kitaev’s classification. This physical pro-

cess of coupling the layers (by condensing inter-layer cooper pairs), corresponds to

a condensation transition on the level of the TQFTs. For every ` < 16, there is a

unique condensation possible and one obtains exactly 16 distinct TQFTs including

Ising, the toric code and the double semion model. They determine the nature of the

topologically protected excitations in the vortices of each superconductor, including

65

their braiding statistics. In essence, this Z16 classification can be seen as a property

of the Ising TQFT.

It is imperative to ask whether multi-layer systems of other TQFTs show a similar

collapse of the classification from Z to ZN for some integer N . In this chapter,

we derive a criterion for when this is not the case, i.e., when the Z classification

generated by a given TQFT is stable. This criterion is based on the fact that there

exist bosonic anyons that cannot be condensed. An example are the bosons in multi-

layered Fibonacci topological order [112, 21, 113]. In this chapter, we generalize this

observation by formulating a no-go theorem that constitutes a sufficient obstruction

against the condensation of a boson. Our criterion and its proof are given using the

tensor category formulation of topological order [80, 81, 82, 83, 84, 85, 20, 35, 36],

which we can use to describe the condensation transition axiomatically [42, 21, 113].

We apply our no-go theorem to several examples, including the forementioned multi-

layer Fibonacci TQFTs.

4.1 First No-Go Theorem

The following definition is useful for formulating our no-go theorem: For a given

anyon b, a subset Ib = {a1, . . . , am} of anyons is called a set of zero modes localized

by b [105] if for all i, j = 1, . . . ,m:

1. The fusion products ai × aj do not contain condensable bosons, except the

identity if ai = aj,1

2. all ai are zero modes of b, by which we mean ai × b = b+ . . ., (i.e. N baib> 0)

3. if a particle ai is in Ib then so is its antiparticle.

1In demanding that ai × aj does not contain condensable bosons, as opposed to not containingany bosons at all (except the identity), we are anticipating a inductive application of the no-gotheorem. Once we have shown that a boson B, whose set IB is such that ai × aj , with ai, aj ∈ IB ,does not contain any boson (except the identity), is uncondensable, it is allowed that B appears inthe fusion product ai × aj of the set IB′ of another boson B′.

66

Bai

BBai

a) b)

~phase with condensed B a#

i

Figure 4.1: Tunneling processes mediated by an anyon condensate. The gray regionis a phase in which a boson B is condensed. a) Vertex of a boson B that localizes azero mode of anyon ai. In the condensed phase, B can be converted into an identityparticle world line (not shown). By the axioms of anyon condensation, processesa) and b) are equivalent, i.e., B can be converted into ai by tunneling through thecondensate.

Note that the choice of Ib for a given boson b is not unique and that Ib may or

may not contain the identity. (The above conditions are satisfied in both cases.)

Typically, we will be interest to find a set Ib that is as large as possible. To motivate

the terminology of the set Ib, observe that N bab > 0 implies that a anyons can always

be emitted or absorbed by b. Therefore, b must carry a zero-mode excitation of a.

We can now state our first main result, a general condition under which a boson B

cannot condense. It is an obstruction that is sufficient to show that condensation of

B cannot occur.

No-go theorem — A boson B cannot condense if there exists a set IB, such that the

sum of the quantum dimensions of all anyons in IB exceeds the quantum dimension

of B, i.e., if

dB < da1 + da2 + · · ·+ dam . (4.1)

Proof. We start by showing that all particles in IB do not split, and have distinct

restrictions. This follows from inspection of Eq. (3.5) for t = ϕ, a = ai, b = aj,

∑r∈T

nrainraj

= δi,j +∑c6=1

N caiaj

nϕc , (4.2)

67

where we used nϕc = nϕc and nraj = nraj . By assumption, there are no condensable

bosons in ai × aj, hence N caiaj

and nϕc cannot be both nonzero for any c 6= 1. Thus∑r n

rainrai = 1, implying a single restriction a↓i of ai, with da↓i

= dai using (A.4).

Moreover,∑

r nrainraj = 0 if i 6= j, implying that the restrictions of ai 6= aj are

distinct particles.

With this knowledge about the restrictions of the ai, Eq. (3.5) for t = ϕ, a = ai,

b = B evaluates to

na↓iB

= na↓iB =

∑c

N caiBnϕc ≥ NB

aiBnϕB, (4.3)

where we used N BaiB

= NBaiB

Inserting this inequality in Eq. (A.4) for a = B, and

using dai = da↓i, we have

dB ≥ nϕB

m∑i=1

NBaiBdai . (4.4)

It follows that in a situation where Eq. (4.1) holds, Eq. (4.4) implies nϕB = 0, i.e., B

does not condense. [Note that in the case NBaiB

> 1, a stronger form of Eq. (4.1) with

dai is replaced by NBaiBdai holds.]

To follow up with a pictorial representation of these equations, consider the tun-

neling of anyons across the domain wall as shown in Fig. 4.1, where each particle a

in the uncondensed theory is converted into its restriction a↓ in the gray region. Fig-

ure 4.1 (a) shows a vertex allowed by the fusion rule ai×B → B in the uncondensed

phase. The boson B enters the condensed phase, where it can disappear as it is part

of the condensate (one of its restrictions is the vacuum ϕ, the world lines of which can

be removed at will). By the fundamental assumption that fusion and condensation

commute [which is at the heart of Eq. (3.5)], Fig. 4.1 (a) is equivalent to Fig. 4.1 (b).

The latter represents a coherent tunneling process that is mediated by the conden-

sate and converts B into any of the ai. The existence of this process implies that the

distinct restriction a↓i of any ai must be in the restriction of B. Hence, by Eq. (A.4),

the quantum dimension of B must be large enough to accommodate all the distinct

68

restrictions of the ai, if B condenses. Therefore if we find sufficiently many ai such

that Eq. (4.1) holds, B cannot condense.

Note that the no-go theorem does not a priori require knowing the braiding data

of A – although the modular tensor category structure fixes that data to some ex-

tend. The theorem involves only data obtainable from N cab. We remark that the no-go

theorem can only ever yield an obstruction against the condensation of non-Abelian

bosons. For Abelian bosons, the theory after condensation can be constructed explic-

itly, which is a constructive proof that there is no obstruction. [113]

We now demonstrate that the no-go theorem is practically useful by considering

three examples: (i) multiple layers of the Fibonacci TQFT, (ii) single layers of the

SO(3)k TQFT for k odd, and (iii) multiple layers of the latter. We will show that all

these theories, while containing bosons, do not admit condensation transitions. All

the bosons are noncondensable. Additional general results, concerning for instance

TQFTs with a condensing Abelian sector and with only a single boson, are given in

appendix B.1.

4.2 Example (i): Multiple layers of Fibonacci

The Fibonacci category AFib is a non-Abelian TQFT containing just one nontrivial

particle τ with a fusion rule τ×τ = 1+τ , a topological spin θτ = ei4π/5, and a quantum

dimension dτ = φ given by the golden ratio φ = (1+√

5)/2. As AFib does not contain

any nontrivial boson, it cannot undergo a condensation transition. We are interested

whether the TQFT formed by N identical layers of AFib i.e., the TQFT A⊗NFib , admits

a condensation transition. The TQFT A⊗NFib contains 2N particles corresponding to all

possible distributions of τ -particles over the N layers. For each r = 0, . . . , N there are(Nr

)so-called (rτ) particles with τ ’s in exactly r layers, each with spin θ(rτ) = ei4πr/5

and quantum dimension d(rτ) = φr. The unique r = 0 particle is the identity of A⊗NFib .

69

From the topological spin, the bosons in A⊗NFib are (rτ) particles with r = 5n, n ∈ Z.

Using the no-go theorem, we show that none of these bosons can condense.

Using proof by induction on n ≥ 1, we show that for any (5nτ) boson B, there

exists a set I(5nτ) such that Eq. (4.1) holds. We first consider the case n = 1. Given

a (5τ) boson, we must construct a set I(5τ) for this boson. Consider the set formed

by all (2τ) particles obtained by replacing any 3 τ ’s in the boson with a 1. There are(52

)= 10 such (2τ) particles for a given (5τ) boson. They form a set I(5τ) that obeys

point 1–3 from the definition: point 1 holds as any product of two of these particles

has at most 4 τs and is therefore not a (potentially condensable) boson. Points 2 and

3 can be checked by using the Fibonacci fusion rules in each layer. Finally, Eq. (4.1)

holds because

d(5τ) = φ5 < 10φ2 =∑

ai∈I(5τ)

dai (4.5)

evaluates to about 11.1 < 26.2. We conclude that none of the (5τ) bosons condense

for any number N of layers of Fibonacci TQFT.

For the induction step, we assume that none of the (5nτ) bosons can condense

for n < n0, n0 > 1, and we show that the same holds for the (5n0τ) bosons. Define

r0 := b(5n0 − 1)/2c, where bxc is the largest integer smaller than or equal to x. For

a given (5n0τ) boson, form the set I(5n0τ) out of all (r0τ)-particles that are obtained

by replacing any (5n0 − r0) τ ’s in the boson (5n0τ) with a 1. There are(

5n0

r0

)such

(r0τ) particles. They form a set I(5n0τ) for (5n0τ). In particular their fusion products

can only contain (5nτ)-bosons with n < n0, which cannot condense by assumption.

Equation (4.1) reads for this case

φ5n0 <

(5n0

r0

)φ5n0−r0 . (4.6)

Using that r0 ∼ 5n0/2 and(

5n0

5n0/2

)∼ 45n0/2/

√π5n0/2 for large n0, we obtain that the

right-hand side of Eq. (4.6) grows like 45n0/2φ5n0/2/√n0, asymptotically dominating

70

the left-hand side. An explicit evaluation yields that Eq. (4.6) holds for any n0 ≥ 1

in fact. We have thus shown that none of the (5n0τ) bosons can condense. This

concludes the induction step and the proof that no boson in A⊗NFib can condense.

4.3 Example (ii): Single layer of SO(3)k

Our second example focuses on the (single-layer) TQFTs associated with the Lie

group SO(3) at values of odd level k. They contain bosons for an infinite subset of

k. We show that none of these bosons can condense. The SO(3)k TQFTs with k odd

have (k + 1)/2 anyons j = 0, · · · , (k − 1)/2 with

dj =sin(π 2j+1k+2

)sin [π/(k + 2)]

, θj = e2πij j+1k+2 . (4.7)

We note that for k odd, all particles have distinct quantum dimensions. The fusion

rules are

N j3j1j2

=

1 |j1 − j2| ≤ j3 ≤ min{j1 + j2, k − j1 − j2}

0 else

. (4.8)

The smallest odd k for which SO(3)k contains a boson is k = 13, in which j = 5 is a

boson – an uncondensable one, as we shall see.

The topological spins θj yield the condition j(j + 1) = k + 2 for the lowest j

that may correspond to a boson (aside from the vacuum j = 0). (Frequently, this

condition cannot be met with integer j, as in the k = 13 example, and the lowest

boson appears at even higher j.) We conclude that the first boson after j = 0 cannot

occur for j lower than

j0 =⌊√

k + 9/4− 1/2⌋. (4.9)

71

dj dj

j j

I II III I II III

a) b)k = 13 k = 103

Figure 4.2: Quantum dimensions and bosons (blue columns) for SO(3)k theories witha) k = 13 and b) k = 103. These are the smallest k, for which SO(3)k contains twoand four bosons, respectively. Indicated are also the ranges I–III defined in Eq. (4.10).The maximum quantum dimension coincides with the boundary between range I andII in Eq. (4.10). For instance, to apply the no-go theorem to the j = 5 boson in a),choose Ij=5 = {j = 2} and use that d5 ≈ 3.6 is smaller than d2 ≈ 4.2.

We will now discuss separately bosons j in the three ranges (see Fig. 4.2 for two

examples)

I. j0 ≤ j ≤ bk/4c (4.10a)

II. bk/4c < j ≤ k − 1

2−⌊j0 − 1

2

⌋(4.10b)

III.k − 1

2−⌊j0 − 1

2

⌋< j ≤ k − 1

2. (4.10c)

Due to Eq. (4.9), bosons jB in range III have no bosons in their fusion product

jB × jB, other than the identity. Thus, from Eq. (3.5) for t = ϕ, and the fact that B

are their own antiparticles, we conclude that they cannot split. Using Eq. (A.4) and

the fact that they have djB > 1, we conclude that they cannot restrict to the vacuum

i.e., they cannot condense.

We now use our no-go theorem to show that bosons jB in range I are non-

condensable. Specifically, we show that the particles 0 < j < bjB/2c form a set

IjB of jB obeying Eq. (4.1). Before establishing that they satisfy the conditions for a

set IjB , let us show that Eq. (4.1) holds for IjB . For large k, we can rely on the follow-

72

ing asymptotic estimate. Using that the sine function in Eq. (4.7) is monotonously

increasing with negative second derivative for j ≤ bk/4c, the estimate

2jB + 1 <

bjB/2c−1∑j=1

(2j + 1) (4.11)

implies Eq. (4.1) for jB in range I. This inequality holds for all jB ≥ 10. Using

Eq. (4.9) we conclude that it applies to all bosons in range I for k ≥ 109. We

verified explicitly that inequality (4.1) holds (using the exact values of the quantum

dimensions) for all bosons in range I for k < 109. Finally, it is readily verified

using Eq. (4.8) that IjB form a set of zero modes localized by jB provided that all

bosons with j < jB cannot condense. The proof then proceeds straightforwardly by

induction.

We apply our no-go theorem successively to bosons jB in range II in order of

increasing jB. Using the result that all bosons in range I are uncondensable, one

verifies that the particles j with 1 ≤ j ≤ min{k− 2jB, bjB/2c− 1} form a set IjB . As

for range I, we can estimate the quantum dimensions. From the relation sin[π(2jB +

1)/(k+2)] = sin[π(k−2jB+1)/(k+2)] we can estimate the quantum dimension of jB

using sin[π(2jB+1)/(k+2)] < π(k−2jB+1)/(k+2). The quantum dimensions of the

anyons in IjB are estimated as for range I with sin[π(2j+1)/(k+2)] < π(2j+1)/(k+2).

Using these estimates we find that if

k − 2jB + 1 <

min{k−2jB ,bjB/2c−1}∑j=1

(2j + 1) (4.12)

holds, Eq. (4.1) follows. In the case k − 2jB < bjB/2c − 1, Eq. (4.12) reduces to

1 < (k − 2jB)2 + (k − 2jB), which is true for all jB in range II for all k. In the case

k−2jB > bjB/2c−1, Eq. (4.12) simplifies to k+2 < 2jB +(bjB/2c)2, which holds for

all jB in range II if k ≥ 37. We verified explicitly that Eq. (4.1) holds for all bosons

73

in range II if k < 37 (they appear in k = 13, 19, 31). This concludes our proof that

no condensation transition is possible in the SO(3)k TQFT for any odd k.

We note that this result can be readily extended to SU(2)k with k odd, since

SO(3)k is the projection of SU(2)k to anyons with integer j. One simply includes

the half-integer j anyons in the theory (none of which are bosons). The sets Ib as

defined above remain the same and so do all the quantum dimensions. Hence, we

also showed the noncondensability of SU(2)k, with k odd. This is consistent with

the ADE classification of SU(2)k [114]: There are no off-diagonal modular invariant

partition functions for odd k in SU(2)k [115]. Thus, the no-go theorem provides a

proof of this fact that is complementary to the ADE classification.

4.4 Example (iii): Multiple layers of SO(3)k

We can show that any number of layers of SO(3)k, with k odd, does not contain

condensable bosons. Fixing k, the proof proceeds again by induction. As induction

base, we proof that all multi-layer anyons with a nontrivial particle in only a single

layer (and the identity anyon in the other k− 1 layers) cannot condense nor split. To

show that, we can use the single-layer result from Example (ii). For the induction

step, we assume that for a fixed k0 < k all multi-layer anyons with nontrivial particles

in l layers, 1 ≤ l ≤ k0, cannot condense and do not split. We can then show that the

same holds for multilayer anyons with nontrivial particles in k0 +1 layers, completing

the induction. The details of this proof are given in appendix B.2.

4.5 Summary

We have presented a generally applicable no-go theorem against the condensation of a

topological boson and illustrated it with several examples. The proof of our theorem

uses mostly the fusion (as compared to the braiding) information of the TQFT.

74

We showed a connection between our results and the ADE classification of SU(2)k

theories, indicating that the no-go theorem might be useful for the classification of

modular invariant partition functions of conformal field theories more broadly. [113] It

would be interesting to study, whether other obstructions against boson condensation

exist or whether our no-go theorem actually constitutes a necessary condition. In all

examples we know, noncondensability is captured by the no-go theorem.

The no-go theorem can be used to study whether a TQFT is ZN graded under

layering. This provides a way to classify TQFTs depending on whether N is finite or

infinite. As a venue for future work, when restricting the condensations to those that

preserve certain symmetries of the anyon model, one could similarly classify symmetry

enriched topological phases, and with this also symmetry protected topological phases

without intrinsic topological order. The classification of the latter is often related to

the former upon gauging the protecting symmetry. [116, 117]

75

Chapter 5

Abelian Boson Condensation in

Field Theory

In this chapter, we reformulate the boson condensation in a field theory language

when the boson is Abelian. We propose that when an operator is condensed, the

partition function of the condensed phase will be invariant under arbitrarily insertion

of the condensed operators. The Lagrangian can be easily modified to have such an

invariance. The modification of the Lagrangian consists of introducing an integer

gauge field that couples with the condensates.

The organization of this chapter is as follows: In Sec. 5.1, we explain our intuition

and basic formalism of the boson condensation for any TQFT. In Sec. 5.2, we apply

the general formalism for K-matrix Chern-Simons theories. The results obtained

in our formalism are the same as those of the previous studies[60, 61, 11]: bosons

and only bosons can be condensed, and the deconfined operators are those who have

trivial braiding with the condensates.

76

5.1 Abelian Boson Condensation Formalism

In this section, we present the first principle for the boson condensation. For clarity,

we restrict our discussion to the condensation of loop operators. However, it can be

easily generalized to any operators. When an operator U is condensed, we expect

that the correlation functions in the condensed phase stay invariant under arbitrary

insertions of U(S1) for all loops S1 in (2+1)D spacetime:

〈U(S1) . . .〉c = 〈. . .〉c, ∀ S1 (5.1)

where “. . .” represents all other possible operators in TQFT; S1 is an arbitrary loop

in (2+1)D where the operator U(S1) lives; the subindex “c” denotes the expectation

value is taken in the condensed phase with a new yet underived Lagrangian. Our

purpose of the calculations in this section is to derive this Lagrangian. We emphasize

that Eq. (5.1) is required to be true for all possible closed loops.

The intuition for Eq. (5.1) comes from the physical expectation that the operator

U becomes a trivial operator (vacuum) in the condensed phase. Hence it has trivial

correlation functions with all other operators.

In order to realizing Eq. (5.1) at the Lagrangian level, we introduce a dynamical

1-form gauge field “c”, and later couple it with the condensed operator U . More

subtly, we require that gauge field c should be an integer field:

dc = 0 mod 2π,

⇔˛

S1

c = 0 mod 2π, ∀ closed paths S1(5.2)

Otherwise, the gauge field c will introduce more topological operators than we expect

for the condensation, and the central charge of the condensed theory will be generally

changed, both of which are not true in the formalism of the condensation phase

77

transitions. We defer to elaborate on this point when we discuss the deconfined

operators in Eq. (5.25). At the present stage, we take Eq. (5.2) as an assumption.

More explicitly, the condensate U is in the form of exp i(¸

f(a)), where a ab-

stractly represents all fields in TQFT and f is an arbitrary 1-form function of a. The

Lagrangian for the condensed phase is proposed:

Lc = L0 −1

2πf(a) ∧ dc, (5.3)

Before moving on, we point out that it might be confusing to observe that we can

use the gauge symmetry in Eq. (5.5) to gauge fix c = 0. Then the flat gauge field c

no longer appears in the Lagrangian Lc for the condensed phase. It seems that the

flat gauge field c does not play any role and does not change the original theory at

all? We need to explain and emphasize that when we use Eq. (5.5) to gauge away

the gauge field c and canonical quantize the theory in Eq. (5.3), we need to enforce

the corresponding Gauss Law in the Hilbert space:

df(a) = 0 mod 2π (5.4)

Therefore, we can still use the original Lagrangian L0 for the condensed phase, but

only need to enforce such a Gauss Law in the Hilbert space.

We prove the condensation lemma to show that our prescription of Eq. (5.3)

implies the intuition of Eq. (5.1):

Condensation lemma: Eq. (5.1) is true, if Lc has a gauge symmetry:

c 7→ c+ 2πλ, (5.5)

where the gauge parameter λ is an integer 1-form.

78

(a) (b)

Figure 5.1: An illustration of integer current and Hodge dual in 3 dimensional space.In Panel (a), jS1 is the integer which only has support on the loop S1. It is a 1-formrepresented by the red arrow. Its Hodge dual ?jS1 is represented by the red squarepenetrated by S1, which is a 2-form perpendicular to jS1 . In Panel (b), jS is theinteger 2-current which only has support on the sphere S. In 3 dimensions, ?jS isthen a 1-form going out S denoted by the red dot.

Proof for condensation lemma :

The derivations for Eq. (5.1) are as follows:

〈exp i

˛S1

f(a, . . .)

. . .〉c

=

ˆD[c]D[a] exp i

ˆ Lc +

˛

S1

f

. . .

=

ˆD[c]D[a] exp i

(ˆLc + f ∧ ?jS1

). . .

=

ˆD[c]D[a] exp i

(ˆL0 −

1

2πf ∧ d(c+ 2πλ)

). . .

=

ˆD[c]D[a] exp i

(ˆLc

). . .

=〈. . .〉c

(5.6)

The derivations need certain explanations:

79

The first equality to the second in Eq. (5.6): jS1 is the integer current which

only has support on the loop S1, in order to have:

˛

S1

f =

ˆfµj

µS1 =

ˆf ∧ ?jS1 (5.7)

? is 3-dimensional Hodge dual which maps the coordinates of forms to their orthogonal

counterparts. For example ?dt = dx ∧ dy. See Fig. 5.1 for illustrations of the Hodge

dual in 3 dimensions. An example of Eq. (5.7) is when S1 is a loop along t-direction

parametrized in terms of Cartesian coordinates:

S1 : (t, 0, 0) (5.8)

Then, we have:

jS1 = δD(x)δD(y)dt,

?jS1 = δD(x)δD(y)dx ∧ dy,(5.9)

where δD is the Dirac δ function. Hence, in this example, Eq. (5.7) is:

˛t

f =

ˆftδD(x)δD(y)d3r

=

ˆf ∧ ?jS1

(5.10)

The second equality to the third in Eq. (5.6): We use the 1-form gauge trans-

formation with a special gauge parameter satisfying

c 7→ c+ 2πλ, dλ = − ? jS1 (5.11)

where the gauge parameter λ is an integer 1-form. We point out that this equation

of λ and jS1 always has a solution, because it is an equation of the vector potential λ

80

which has a unit of flux in 3 dimensions. Consistently, the solution of λ is an integer

1-form because: ˛

S1

λ =

ˆ

D2

dλ =

ˆ

D2

− ? jS1 = −ˆ

D2∩S1

1 ∈ Z (5.12)

where we have used Stokes’ theorem, D2 is a disk whose boundary is S1, and D2∩S1

is the intersection of D2 and S1 which can only be an integer number of points.

The fourth equality to the fifth in Eq. (5.6): We use the assumption that

Lc respects the symmetry Eq. (5.11). Hence, from Eq. (5.6), (5.7) and (5.11), the

operator exp i(¸

S1 f(a))

satisfies Eq. (5.1) and is thus condensed in the condensed

phase described by Lc.

Therefore we have completed the proof for the condensation lemma. 2

A related observation was coined “generalized global symmetry” in Ref. [118,

119] when the authors discussed higher form global symmetry for a TQFT. As a

result, the gauge fields in TQFT are shifted by constant 1-forms without changing

the actions. We emphasize that in our situation, it is a gauge symmetry instead of

a global symmetry, and the gauge fields in the condensation formalism are shifted

by local 1-forms (not constant 1-forms), in order to implement Eq. (5.1). Hence, our

starting point Eq. (5.1) implies a 1-form gauge symmetry.

5.2 Condensations in K-Matrix Chern-Simons

Theories

In this section, we apply our formalism (the condensation lemma) to the K-matrix

Chern-Simons theories. This section is divided into two parts. (1) We first derive the

condition under which the operators can be condensed. (2) We find the conditions of

whether operators are confined/deconfined after condensation.

81

5.2.1 Condensable Condition

We begin with a general bosonic K-matrix theory and condense the operator U{l}.

We know from the past studies[60, 61, 11] that only bosons can be condensed. This

is the only requirement for the condensed particle in Abelian topological theories. In

other words, the condensable U{l} requires that l ·K−1 · l ∈ 2Z. We derive this boson

condensation condition from our condensation lemma developed in Sec. 5.1.

The Lagrangian for the condensed phase, according to Eq. (5.3), is:

Lc =KIJ

4πaI ∧ daJ −

1

2πlIaI ∧ dc (5.13)

where K is a symmetric matrix. The repeated indices imply summation. We need to

verify that Lc is invariant under gauge transformation Eq. (5.5). We can first slightly

change Lc using integration by part and the fact that K and K−1 are symmetric

matrices:

Lc = − lMK−1MN lN

4πc ∧ dc+

KIJ

(aI − lM(K−1)MIc

)∧ d(aJ − lN(K−1)NJc

)(5.14)

Therefore, in order for Lc to have the higher form gauge symmetry Eq. (5.5), we

need to introduce the transformations for aI fields, when the gauge field c takes the

transformation Eq. (5.5):

c 7→ c+ 2πλ, aI 7→ aI + 2πlMK−1MIλ, ∀ I. (5.15)

The extra terms in Lc after the gauge transformation in Eq. (5.15) only come from

the first term in Eq. (5.14):

δLc = − lMK−1MN lN

(4π2λ ∧ dλ+ 4πλ ∧ dc

)(5.16)

82

where λ∧ dλ is an integer 3-form, and λ∧ dc is a 2π integer 3-form due to Eq. (5.2).

In order to make the variation of action´δLc to be 0 modulo 2π, we conclude that

the coefficient of λ ∧ dλ has to be an integer multiple of 2π. Hence,

l ·K−1 · l ∈ 2Z, (5.17)

The condition for the higher form gauge invariance in Eq. (5.17) is equivalent to the

statement that U{l} is a boson. As we discussed in Eq. (5.6), the higher form gauge

invariance is equivalent to condensation. Hence Eq. (5.17) is indeed the condition for

condensable operator U{l}. To conclude, when Eq. (5.17) holds, U{l} is condensable.

We can also manifest the higher form gauge invariance condition of Eq. (5.17)

without resorting to a particular gauge transformation (Eq. (5.15)). The basic idea is

that it will be easier to find the higher form gauge invariance, if we have an effective

expression for the partition function Zc in terms of only the gauge field c. We integrate

over all aI fields without any source terms in the path integral, and thus obtain an

effective expression for Zc with an effective Lagrangian Leffc (c).

Leffc (c) = − l ·K

−1 · l4π

c ∧ dc (5.18)

Then we only need to examine whether Leffc (c) has the higher form symmetry Eq. (5.5).

Leffc (c) respects it modulo 2π only when Eq. (5.17) holds, which derivations are the

same as in Eq. (5.16). For simplicity, Leffc (c) will be the notation of the effective

Lagrangian in the condensed phase, which implies the calculations of integrating out

all gauge fields except c.

Therefore, we have proved that the condensed theory Lc has the higher form gauge

symmetry is equivalent to that the condensed operator is a boson (i.e., Eq. (5.17)).

We can directly generalize this statement from one condensed operator to multiple

condensed operators. As a result, the condensation conditions require that each con-

83

densate is a boson and they are mutual bosons. One way to prove such condensation

conditions is to generalize Eq. (5.18):

Lc =KIJ

4πaI ∧ daJ −

liI2πaI ∧ dci ⇒ Leff

c (c) = − li ·K−1 · lj

4πci ∧ dcj (5.19)

where i, j indices label different condensates, U{li}, and Leffc (c) is obtained from Lc by

formally integrating out aI fields. In order to make Leffc (c) in Eq. (5.19) respect the

higher form gauge symmetry:

ci 7→ ci + 2πλi, ∀ i (5.20)

we need the following conditions:

li ·K−1 · lj ∈ Z, ∀ i 6= j; li ·K−1 · li ∈ 2Z, ∀ i. (5.21)

The first condition is the same as the statement that different condensates are mutual

bosons, and the second one states that each condensate U{li} is a boson.

5.2.2 Confinement/Deconfinement

In this part, we examine the confinement/deconfinement after condensing U{l}, using

the higher form gauge symmetry Eq. (5.15). The deconfined operators are invariant

under the higher form gauge symmetry in Eq. (5.15), while the confined ones are not

invariant under the higher form gauge symmetry.

Under the higher form gauge transformation Eq. (5.15), a general loop operator

U{m}(S1) transforms as:

U{m}(S1) 7→U{m}(S1) exp i

(2πm ·K−1 · l

˛S1

λ

)(5.22)

84

We expect that the deconfined operators are invariant under the higher form gauge

symmetry. Because¸S1 λ is generally quantized to an integer, the deconfined operator

U{m}(S1) needs to satisfy:

m ·K−1 · l ∈ Z, (5.23)

in order to stay invariant under the higher form gauge symmetry. This means that

the deconfined U{m} braids trivially with the condensate U{l}.

The generalization of deconfinement to the situation of multiple condensed opera-

tors U{li} is straightforward where the index i denotes different condensed operators.

The deconfined operators should braid trivially with each condensed operator:

m ·K−1 · li ∈ Z, ∀ i . (5.24)

As we promised in Eq. (5.2), we explain that the introduced gauge field “c” must

be a flat gauge field satisfying Eq. (5.2). The reason is that we need to make sure that

attaching a Wilson loop of gauge field c to any deconfined loop operators U{m}(S1)

will not change the expectation value, i.e.:

exp i

˛S1

mIaI + c

= exp i

˛S1

mIaI

, ∀ S1 ⇒ exp i

˛S1

c

= 1, ∀ S1.

(5.25)

This means the introduced gauge field c has to be integer as in Eq. (5.2). On the other

hand, if the gauge field c is a U(1) gauge field without the integer condition Eq. (5.2),

it will also change the chirality when condensation. In this scenario, the condensed

Lagrangian Eq. (5.13) can be written in terms of the K-matrix Chern-Simons theories:

Kc =

K −(l1, l2, . . .)T

−(l1, l2, . . .) 0

(5.26)

85

The basis of this Kc matrix is (a1, a2, . . . , c). The central charge of Kc is generally

different from that ofK. However, according to previous studies based on the modular

tensor categories[92, 120], the boson condensation will not change central charge

modulo 24. This contradiction shows that, Lc, if we treat gauge field c as a U(1)

gauge field without the flatness condition, does not describe the boson condensation.

Therefore, we have justified that the Lagrangian Eq. (5.13) can describe condensed

phase when c is a flat gauge field satisfying Eq. (5.2).

86

Chapter 6

Fracton Models, Tensor Network

States and Their Entanglement

Entropies

TNS have been heavily used in condensed matter physics in the past decade, especially

in the study of 1D and 2D topological phases[121]. Amongst many examples,

1. Numerical simulations of the 1D Haldane chain led to the discovery of symmetry

protected topological phases (SPT)[122].

2. Fractional quantum Hall states can be exactly written as MPS[123, 124, 125,

126, 127, 128, 129, 130, 131, 132, 133, 134] which allows performing numerical

calculations not accessible by exact diagonalization techniques.

3. A large class of spin liquids wave functions can be constructed using TNS with

global spin rotation symmetries and lattice symmetries[135, 136, 137, 138, 139,

140, 141].

In higher spatial dimensions than 2D, more exotic gapped states of matter exist,

beyond the paradigm of topological phases.[19, 121]. Recently, 3D so-called fracton

87

models[142, 143, 144, 145, 146, 147, 148, 149, 150, 151, 152, 153, 154, 155, 156, 157,

158, 159, 160, 161, 162, 163, 164, 165, 166, 167, 168] represented by Haah code[142]

and X-cube model have been proposed, attracting the attention of both quantum

information[169] and condensed matter community[170, 171, 172, 173, 174]. They can

be realized by stabilizer code Hamiltonians, whose fundamental property is that they

consist solely of sums of terms that commute with each other. They are hence exactly

solvable. The defining features of fracton models include (but are not restricted to)

that:

1. Fracton models are gapped, since they can be realized by commuting Hamilto-

nian terms.

2. The ground state degeneracy on the torus changes as the system size changes.

Hence, fracton models seem not to have thermodynamic limits.

3. The low energy excitations can have fractal shapes, other than only points and

loops available in conventional topological phases.

(a)

(b)

Figure 6.1: Examples of TNS lattice wave functions in 1D and 2D. Each node is atensor whose indices are the lines connecting to it. The physical indices - of the quan-tum Hilbert space - are the lines with arrows, while the lines without any arrows arethe virtual indices. Connected lines means the corresponding indices are contracted.Panel (a) is an MPS for 1D systems. Panel (b) is a PEPS on a 2D square lattice.

88

4. The excitations of fracton models are not fully mobile: they can only move

either along submanifold of the 3D lattice (Type I fracton model), or completely

immobile without energy dissipation (Type II fracton model).

In this chapter, we obtain a TNS representation for some of the ground states of

three stabilizer codes in 3D: the 3D toric code model[19], the X-cube model and the

Haah code. The two latter ones belong to the catalog of fracton models, while the first

one belongs to the conventional topological phases. For instance, the ground state

degeneracies on the torus of the X-cube model and the Haah code do not converge to a

single number, as the system size increases. In contrast, the ground state degeneracy

(GSD) on the torus for 3D toric code model is 8 for all system sizes. Ref. [175]

treated the X-cube and Haah code models using the idea of lattice gauge theory. The

gauge symmetry is generally generated by part of the commuting Hamiltonian terms;

the rest of the Hamiltonian terms are interpreted as enforcing flat flux conditions.

More explicitly, the authors treated the terms only made of Pauli Z operators as the

gauge symmetry generators, and the terms only made of Pauli X operators as the

flux operators. The gauge symmetries in the X-cube and the Haah code models are

not the conventional Z2 gauge symmetry such as that in the 3D toric code model,

since the gauge symmetry generators, the Pauli Z terms of the X-cube and the Haah

code models, are different from those in the 3D toric code model. Refs. [176, 177]

derived the X-cube model from “isotropically” layered 2D toric code models and

condensations. The caveat is that this condensation is weaker than the conventional

boson condensation in modular tensor category or field theory[21, 62, 92, 120]. The

authors condense “composite flux loop” of coupled layers of 2D toric code model.

The “composite flux loop” refers to a composite of four flux excitations near a bond

of the lattice. See Ref. [177] for explicit explanations.

Using the TNS representations of some of the ground states, we obtain the en-

tanglement entropy upper bounds for all three models. We then derive the reduced

89

density matrix cuts for which the TNS represents the singular value decomposition

(SVD) of the state. For these types of cuts, the entanglement entropy of the three

stabilizer codes can be computed exactly. We find that for the fracton models, the

entanglement entropy has linear corrections to the area law, corresponding to an

exponential degeneracy in the TNS transfer matrix.

The transfer matrices of TNS of 2D toric code[19], whose eigenvalues and eigen-

states dominate the correlation functions, have been studied in Ref. [178, 179]. The

flat entanglement spectra[27] of the 2D toric code were studied in Refs. [180, 181].

Our TNS construction, when restricted to 2D toric code model, gives the exact results

of transfer matrices and entanglement spectra. See Chapter 2 for explicit calculations

and explanations. Beyond the 2D toric code, Refs. [182, 158] prove that the reduced

density matrix of any stabilizer code is a projector. Hence, the corresponding entan-

glement spectrum is flat, a property that we will rederive from our TNS.

We will not discuss cocycle twisted topological phases, including Dijkgraaf-Witten

theories[183, 184, 185] or (generalized) Walker-Wang models[186, 187, 188, 189, 190],

even though they can still be realized by commuting Hamiltonians on lattice[184, 183,

186, 82]. However, the presence of nontrivial cocycles will make the TNS construction

very different, based on the experiences in 2D TNS. Our construction will not work

for these twisted models. For instance, in 2D, the virtual index dimension using

our construction is the same as the physical index dimension. However, when we

consider cocycle twisted topological phases, the “minimal” virtual bond dimension is

generally larger than the physical index dimension[191, 38, 192, 193]. More explicitly,

the minimal virtual bond dimension for 2D toric code model is 2, while the minimal

virtual bond dimension for 2D double semion model (twisted toric code) is 4[191,

38, 192, 193]. The 2D cocycle twisted TNS has been systematically explored in the

literature for bosonic[191] and fermionic[194, 195] systems respectively.

90

The organization of this chapter is as follows: In Sec. 6.1, we set the notations

and provide an overview and the general idea of the TNS construction In Sec. 6.2,

we present the calculation of the entanglement properties using the developed TNS

construction In Sec. 6.3, we present the TNS construction for the toric code model

in 3D. The entanglement entropy is calculated from the obtained TNS. The transfer

matrix is constructed afterwards and is proven to be a projector of rank 2. In Sec. 6.4,

we present the TNS construction for X-cube model. The same calculations for the en-

tanglement entropy and the transfer matrix are presented. They are quickly shown to

be very different from the toric code model. Indeed, the entanglement entropies have

linear corrections to the area law, and the transfer matrix is exponentially degenerate.

In Sec. 6.5, we present the TNS construction for Haah code. The entanglement en-

tropies are calculated for several types of cuts. In Sec. 6.6, we summarize the chapter

and discuss future directions.

6.1 Stabilizer Code Tensor Network States

In this section, we provide an overview of the stabilizer codes and the tensor network

state description of their ground states. In this article, we focus on a few “main”

stabilizer codes in three dimensions : the toric code[19], the X-cube model[175] and

the Haah code[142]. The TNSs for these models have similarities in their derivation

and they share several (but importantly not all!) common features. Both aspects are

presented in this section. For pedagogical purposes, we discuss the 2D toric code in

Chapter 2.

6.1.1 Notations

We first fix some of the notations used in the chapter, to which we will refer throughout

the manuscript:

91

1. The Pauli matrices X and Z are defined as:

X =

0 1

1 0

, Z =

1 0

0 −1

. (6.1)

2. We introduce a g tensor, which denotes the projector from a physical index to

virtual indices. g tensors are essentially the same (up to the number of indices)

for all stabilizer codes. g tensors have two virtual indices and one physical

index for the 3D toric code model and the X-cube model, while g tensors for

the Haah code have four virtual indices and one physical index. They are

depicted in Eq. (6.5), (6.116) and (6.117).

3. We introduce the T tensor, which denotes the local tensor for each model.

It only has virtual indices and thus no physical indices. The specific tensor

elements are determined by the Hamiltonian terms.

4. Since we consider mostly models on cubic lattices, the indices of T tensors will

be denoted as x, x, y, y, z and z in the 3 directions (forward and backward)

respectively. The indices will be collectively denoted using curly brackets. For

instance, the physical indices are collectively denoted as {s}, while the virtual

indices are denoted as {t}. The virtual indices which are not contracted over

are called “open indices”. Both the physical indices and the virtual indices are

non-negative integer values.

5. Graphically, the physical indices are denoted by arrows, while the virtual indices

are not associated with any arrows. See Fig. 6.1.

6. The contraction of a network of tensors over the virtual indices is denoted as

CM ( ) whereM is the spatial manifold that the TNS lives on. The correspond-

ing wave function that arises from the contraction is denoted as |TNS〉M. When

92

U-1U

A1 A2

A1 A2

(a)

(b)

U-1UA1 A2

A1 A2

(c)

(d)

Figure 6.2: An illustration of the TNS gauge in MPS. (a) A part of an MPS. A1

and A2 are two local tensors contracted together. (b) We insert the identity operatorI = UU−1 at the virtual level - it acts on the virtual bonds. The tensor contractionof A1 and A2 does not change. (c) We further multiply U with A1 and U−1 with A2,resulting in A1 and A2 respectively in Panel (d). The tensor contraction of A1 andA2 is the same as the tensor contraction of A1 and A2. The TNS does not change aswell. Similar TNS gauges also appear in other TNS such as PEPS.

evaluating the TNS norms or any other physical quantities, we contract over

the virtual indices from both the bra and the ket layer. This contraction is still

denoted by CM ( ).

7. Lx, Ly and Lz refer to the system sizes in the three directions (the bound-

ary conditions will be specified), while lx, ly and lz refer to the sizes of the

entanglement cut. Both are measured in units of vertices.

8. The TNS gauge is defined as the gauge degrees of freedom of TNS such that

the wave function stays invariant while the local tensors change. One can insert

identity operators I = UU−1 on the virtual bonds, where U is any invertible

matrix acting on the virtual index, multiplying U and U−1 to nearby local

tensors respectively. The local tensors then change but the wave function stays

invariant. We refer to this gauge degree of freedom as the TNS gauge. The

TNS gauge exists in MPS, PEPS etc. See Fig. 6.2 for an illustration. In our

calculations, we only fix the tensor elements up to the TNS gauge.

93

Tg

T

g

(a) (b)

xzy

xzy

Figure 6.3: (a) A plane of TNS on a cubic lattice. (b) TNS on a cube. The lineswith arrows are the physical indices. The connected lines are the contracted virtualindices, while the open lines are not contracted. On each vertex, there lives a Ttensor, and on each bond, we have a projector g tensor.

6.1.2 Stabilizer Code and TNS Construction

We now summarize the general idea of constructing TNSs for stabilizer codes. In

Chapter 2, we provide the construction of the TNS for the 2D toric code model on a

square lattice. In the following, we assume that the physical spins are defined on the

bonds of the cubic lattice (such as the 3D toric code and the X-cube models). The

cases where the physical spins are defined on vertices can be analyzed similarly. The

generic philosophy of any stabilizer code model is captured by the following exactly

solvable Hamiltonian:

H = −∑v

Av −∑p

Bp (6.2)

where the Hamiltonian is the sum of the Av terms which are products of only Pauli Z

operators, and the Bp terms which are products of only Pauli X operators. v and p

denotes the positions of the Av and Bp operators on the lattice. In the 3D toric code,

v is the vertex of the cubic lattice while p is the plaquette. In the X-cube model, v

is the vertex while p is the cube. In the Haah code, both v and p are cubes. See

Sec. 6.3.1, 6.4.1 and 6.5.1 for the definitions of Hamiltonians of these three models.

94

All these local operators commute with each other:

[Av, Av′ ] = 0, ∀ v, v′

[Bp, Bp′ ] = 0, ∀ p, p′

[Av, Bp] = 0, ∀ v, p.

(6.3)

The Hamiltonian eigenstates are the eigenstates of these local terms individually. In

particular, any ground state |GS〉 should satisfy:

Av|GS〉 = |GS〉, ∀ v

Bp|GS〉 = |GS〉, ∀ p(6.4)

for all positions labeled by v and p. In this chapter, we only consider Hamiltonians

being of a sum of local terms that are either a product of Pauli Z operators, or a

product of Pauli X operators. Thus, we do not include the case of mixed products of

Pauli Z and X operators.

The ground states for the stabilizer codes with Hamiltonian as in Eq. (6.2) can be

written exactly in terms of TNS. Our construction, when restricted to the 2D toric

code model, is the same as in the literature[192, 193]. In the following, we provide

one possible general construction for such TNSs. We introduce a projector g tensor

with one physical index s and two virtual indices i, j:

gsij =

i j

s=

1 s = i = j

0 otherwise

(6.5)

where the line with an arrow represents the physical index, and the lines without

arrows correspond to the virtual indices. The physical index s = 0, 1 represents

the Pauli Z eigenstates of |↑〉, |↓〉 respectively where Z|↑〉 = |↑〉, and Z|↓〉 = −|↓〉.

The projector g tensor maps the physical spin into the virtual spins exactly. As a

95

result, the virtual index has a bond dimension 2. When a Pauli operator acts on the

physical index of a projector g tensor, its action transfers to the virtual indices of g.

For instance, a Pauli operator X acting on the physical index of a g tensor amounts

to two Pauli operators X acting on both virtual indices of the same g tensor, and

a Pauli operator Z acting on the physical index of a g tensor amounts to a Pauli

operator Z acting on either virtual index of the same g tensor.

To each vertex, we associate a local tensor T which only has virtual indices. To

each bond, we associate a projector g tensor. The TNS is obtained by contracting

the g and T tensors as depicted in Fig. 6.3 (a) and (b). We define the TNS as:

|TNS〉 =∑{s}

CR3

(gs1gs2gs3 . . . TTT . . .) |{s}〉 (6.6)

where CR3denotes the contraction over all virtual indices on R3 as illustrated in

Fig. 6.3 (b); |{s}〉 is a wave function basis for spin configurations on the cubic lattice

in Pauli Z basis. The TNS can be put on other spatial manifolds such as T 3 and

T 2 ×R. In our notation, they are denoted by changing CR3to CT 3

and CT 2×R. The

TNS for the ground states satisfies:

Av|TNS〉 = |TNS〉, ∀ v

Bp|TNS〉 = |TNS〉, ∀ p(6.7)

for all positions labeled by v and p.

The actions of Av and Bp operators on the TNS can be transferred to the virtual

indices, using the definition of the g tensor. Since the virtual indices of projector g

tensors are contracted with the virtual indices of T tensors, the actions of Av and

Bp on the physical indices will be transferred to actions on the local tensors T . By

enforcing the local tensors T to be invariant under Av and Bp actions, we obtain

Eq. (6.7), and |TNS〉 belongs to the ground state manifold. For the three models

96

analyzed in this chapter, we have found that up to TNS gauge, the elements of the

local tensor T can be reduced to two values, either 1 or 0. The first equation of

Eq. (6.7) restricts the local T tensor to be:

Txxy...

6= 0 if the indices xxy . . . satisfy

some constraints

= 0 otherwise

. (6.8)

Applying the second equation of Eq. (6.7) will further restrict the local T tensor to

be:

Txxy... =

1 if the indices xxy . . . satisfy

some constraints

0 otherwise

. (6.9)

For simplicity, we calculate the entanglement entropies of the wave function on R3

with respect to some specific entanglement cuts, and compute the ground state de-

generacy (GSD) of the 3D toric code and X-cube model on T 3.

We emphasize that in this chapter, we are only concerned with the bulk wave

functions and their entanglement entropies. In principle, the TNS of Eq. (6.6) re-

quires boundary conditions, i.e. the virtual indices at infinity on R3. The boundary

conditions are assumed not to make a difference to the reduced density matrices in

the bulk. (Notice that this is true as long as the region considered for the reduced

density matrices does not contain any boundary virtual index.) Hence, we do not

need to specify the boundary conditions for the TNS in the following calculations of

entanglement entropies.

97

6.1.3 TNS Norm

Evaluating the norm of the TNS given by Eq. (6.6) (or any scalar product between

two TNS) is straightforward. Indeed the g tensors are projectors, and hence greatly

simplify the expression of the tensor network norm when we contract over the physical

indices.

Given the wave function of Eq. (6.6), we can compute its norm as follows:

〈TNS|TNS〉 =

∑{s}

CR3

(gs1gs2gs3 . . . TTT . . .) 〈{s}|

?∑{s}

CR3

(gs1gs2gs3 . . . TTT . . .) |{s}〉

=∑{s}

CR3

(gs1?gs2?gs3? . . . T ?T ?T ? . . .) CR3

(gs1gs2gs3 . . . TTT . . .)

=∑{s}

CR3

(gs1gs2gs3 . . . T ?T ?T ? . . .) CR3

(gs1gs2gs3 . . . TTT . . .) ,

(6.10)

where ? is the complex conjugation, and we have used the fact that the g tensors

are real for our models. Now we specify a contraction order in Eq. (6.10): we first

contract over the physical indices and then we contract over the virtual indices. If the

physical indices of two projector g tensors are contracted over, the four virtual indices

will be enforced to be the same following the definition of the projector g tensor:

∑s

gsijgsmn =

i j

m n

=

1 i = j = m = n

0 otherwise

.

(6.11)

Thus when computing wave function overlap 〈TNS|TNS〉, the virtual indices in the

bra layer and the ket layer at the same place are enforced to be same. As a result,

98

we have:

〈TNS|TNS〉 = CR3

(. . .TTT . . .) , (6.12)

where CR3stands for the contraction of a network of tensors T over the virtual indices

on R3. In a slight abuse of notation, CR3in Eq. (6.12) stands for the contraction

taken over the virtual indices of both the bra and the ket layer, while the contraction

in Eq. (6.6) is taken over the virtual indices in only the ket layer. The double tensor

T is defined as

Txxy...,x′x′y′... = T ∗xxy...Tx′x′y′...δxx′δxx′δyy′ . . .

= |Txxy...|2δxx′δxx′δyy′ . . . ,(6.13)

for all the elements of T and T. The indices are not summed over in the above

equation. The indices xxy . . . come from the bra layer while the indices x′x′y′ . . .

come from the ket layer. In a 2D square lattice, a T tensor usually has 4 virtual

indices x, x, y, y, while in a 3D cubic lattice, a T tensor usually has 6 virtual indices

x, x, y, y, z, z. If the elements of the T tensor are only either 0 or 1, we get,

Txxy...,x′x′y′... = |Txxy...|2δxx′δxx′δyy′ . . .

= Txxy...δxx′δxx′δyy′ . . . .

(6.14)

Then,

〈TNS|TNS〉 =CR3

(. . .TTT . . .)

=CR3

(. . . TTT . . .) .

(6.15)

This result will be frequently used in the following discussions, especially when we

compute wave function overlaps or transfer matrices. Eqs. (6.12) and (6.15) hold true

on other manifolds as well, such as T 3 and T 2 ×R.

99

Figure 6.4: Transfer matrix (red dashed square) of a 1D MPS. The connected linesare the contracted virtual indices. The connected arrow lines are the contractedphysical indices. The MPS norm (or any other quantities) can be built using thetransfer matrix. Higher dimensional transfer matrices are similarly defined for TNSon a cylinder or a torus, by contracting in all directions except one. This leads to a1D MPS with a bond dimension exponentially larger than the TNS one.

6.1.4 Transfer Matrix

The transfer matrix method is ubiquitous when using MPS (see Fig. 6.4 for an il-

lustration of the transfer matrix). It can be generalized to TNS on a 2D cylinder

by contracting tensors along the periodic direction of the cylinder. This implies that

the bond dimension of the transfer matrix is exponentially large with respect to the

cylinder perimeter. In 3D, the TNS norm on T 3 of size Lx × Ly × Lz can be written

as an MPS using transfer matrices TMxy in each xy-plane:

〈TNS|TNS〉 = Tr (TMxy,z=1TMxy,z=2 . . .)

= Tr(

(TMxy,z=1)Lz) (6.16)

where we have assumed that all transfer matrices in each plane are the same:

TMxy,z=1 = TMxy,z=2 = . . . (6.17)

Eq. (6.16) is an alternative way of writing the wave function norm and specifies a

contraction order of the tensors in Eq. (6.12): we first contract the virtual indices

along xy-plane which defines the transfer matrix TMxy, and then contract the vir-

tual indices in z-direction which leads to the multiplication and the trace of transfer

100

matrices. The transfer matrix TMxy is defined as:

TMxy =

∑{s}

CT 2xy (gs1?gs2?gs3? . . . T ?T ?T ? . . .) 〈{s}|

∑{s}

CT 2xy (gs1gs2gs3 . . . TTT . . .) |{s}〉

,

(6.18)

where the TNS contraction is performed along the xy-plane with periodic boundary

conditions, i.e., the 2D torus T 2xy. Denoting Txy as the TNS in its ket layer, it can

depicted as:T

g

x (pbc)z

y (pbc)

(6.19)

TMxy is the overlap of the bra and ket layer of this TNS over the plane with periodic

boundary conditions.

Furthermore, applying Eq. (6.11) to Eq. (6.18), the virtual indices in the bra layer

and the ket layer are identified after the physical indices are contracted in Eq. (6.18).

Hence, we have:

TMxy = CT 2xy (. . .TTT . . .) , (6.20)

where the tensors T, defined in Eq. (6.13), are in the xy-plane with periodic boundary

conditions. The indices in the z-direction are open. By Eq. (6.14) - which is true

when the elements of the T tensor are either 0 or 1 - the transfer matrices is further

simplified to:

TMxy = CT 2xy (. . . TTT . . .) . (6.21)

101

Graphically:

TMxy =

T

x (pbc)z

y (pbc)

T T T T T T

T T T T T T T

T T T T T T T

T T T T T T T

. (6.22)

Suppose the virtual index is of dimension D. Then in Eq. (6.18), the transfer

matrix is of dimension D2LxLy ×D2LxLy . However, in Eq. (6.20), the transfer matrix

reduces to dimension DLxLy × DLxLy , since the indices in the bra layer and the ket

layer are identified due to the contraction over the physical indices of projector g

tensors.

6.2 Entanglement properties of the stabilizer code

TNS

The specific structure of the TNS discussed in the previous section allows us to

derive its entanglement properties. In this section, we show that for a large class

of entanglement cuts, the TNS is already in Schmidt form, i.e. is exactly a singular

value decomposition (SVD). We also summarize the main results for the entanglement

entropies and the transfer matrices that we have obtained for the three stabilizer

codes.

6.2.1 TNS as an exact SVD

We propose a general sufficient condition that the TNS is an SVD with respect to

particular entanglement cuts. Let us denote the TNS with open virtual indices {t}

102

as:

|{t}〉 =∑{s}

CM (TTT . . . gs1gs2gs3 . . .) |{s}〉, (6.23)

whereM is an open manifold which the TNS lives on, CM stands for the contraction

over the virtual indices inside M, but not over the open ones {t} that straddle the

boundary ofM. In Eq. (6.23), the T tensors and g tensors are the tensors insideM

such that the nodes of the local T tensors and the projector g tensors are inside M.

For example, when M is a cube, we have a TNS figure:

|{t}〉 =

{t}

, (6.24)

where inside the cube is a network of contracted tensors which are not explicitly

drawn, and the red lines denote the open virtual indices {t}. With this notation of

|{t}〉, the TNS can be written as:

|TNS〉 =∑{t}

|{t}〉A ⊗ |{t}〉A (6.25)

with respect to a region A and its complement A. |{t}〉A is the TNS in region A with

open indices {t}, while |{t}〉A is the TNS in region A with the same open indices {t}

due to tensor contraction. In other words, the TNS naturally induces a bipartition of

the wave functions. However, the two partitions do not need to each form orthonormal

sets.

103

We now propose a simple sufficient (but not generally necessary) condition to

determine when Eq. (6.25) is an exact SVD for the TNS constructed in this chapter.

We first have to make an assumption, satisfied by all our TNSs:

Local T tensor assumption: We assume that the indices of the nonzero elements

of the local T tensor are constrained: if all the indices of the element T...t... except for

t are fixed, then there is only one choice of t such that T...t... is nonzero.

This assumption can be easily verified when the local T tensors are obtained for

the three models studied in this chapter, such as the 3D toric code model in Eq. (6.43).

We are now ready to express our SVD condition:

SVD condition: If there are no two open virtual indices in {t} (see Eq. (6.24))

of the region A that connect to the same T tensor in the region A, then the non-

vanishing states |{t}〉A span an orthogonal basis. Similarly, if there are no two open

virtual indices in {t} of the region A that connect to the same T tensor in the region A,

then the non-vanishing states |{t}〉A form an orthogonal basis. Therefore, Eq. (6.25)

is an exact SVD.

Proof :

We first prove the statement for the region A. Suppose that |{t}〉A and |{t′}〉A

are two non-vanishing TNSs in the region A. Any open index in {t} of the region A

must connect to either a projector g tensor or a local tensor T . We discuss the two

situations respectively, and examine the overlap of two different states A〈{t′}|{t}〉A

as a function of the two indices configurations {t′} and {t}.

(1) If the open virtual index m in the ket layer (i.e. |{t}〉A) connects to a projector

g tensor, then the open virtual index m′ in the bra layer (i.e. A〈{t′}|), at the same

place as the index m, also connects to a projector g tensor. If we “zoom in” on the

104

local area of A〈{t′}|{t}〉A near the index m and m′, we have the following diagram:

A

m'

m

A

Entanglement Cut

ket layer

bra layer

(6.26)

By using Eq. (6.11), we can conclude that m = m′, otherwise A〈{t′}|{t}〉A = 0.

(2) If the open virtual index m0 in the ket layer connects to a local T tensor, we

require by the SVD condition that there are no other open virtual indices connecting

to this T tensor. Then the other indices of this T tensor are all inside the region A.

Similarly for the index m′0 in the bra layer. In terms of a diagram, A〈{t′}|{t}〉A near

the area of the index m0 and m′0 can be represented as:

m'

T

T

A

ket layer

bra layer

A

mi

m'i

m0

0

Entanglement Cut

(6.27)

where mi and m′i with i = 1, 2, 3 . . . denote the other virtual indices of the T tensor in

the bra and ket layer respectively, except m0 and m′0. Notice that in the ket layer, the

virtual indices mi (i = 1, 2, . . .) of the T tensor (all indices except the index m0) are

all connected with contracted projector g tensors inside region A. Correspondingly,

in the bra layer, the virtual indices m′i (i = 1, 2, . . .) are also all connected with the

same contracted projector g tensors. Hence, due to these projector g tensors and

Eq. (6.11), all the indices except m0 of the T tensor in the ket layer are equal to their

105

respective analogues in the bra layer:

mi = m′i, i = 1, 2, . . . (6.28)

otherwise the overlap would be A〈{t′}|{t}〉A = 0. The only remaining question is

whether the open indices m0 and m′0 should be identified in order to have a non-

vanishing overlap A〈{t′}|{t}〉A.

Using the local T tensor assumption:, mi (i = 1, 2, . . .) will uniquely determine

m0 in order to have nonzero element of the T tensor in the ket layer. Similarly,

m′i (i = 1, 2, . . .) will uniquely determine m′0 in order for the T tensor in the bra layer

to give a nonzero element. Therefore, Eq. (6.28) implies that:

m0 = m′0 (6.29)

such that the overlap A〈{t′}|{t}〉A is nonzero.

Therefore, both situations (1) and (2) lead to the conclusion that the open indices

{t} and {t′} should be identical in order to have a nonzero overlap A〈{t′}|{t}〉A. The

non-vanishing states |{t}〉A are orthogonal basis. A similar proof can be derived for

the region A. The orthogonality of each set |{t}〉A and |{t}〉A implies that Eq. (6.25)

is indeed an SVD. However, the singular values are not clear at this stage since the

basis may not be orthonormal (i.e., the states might not be normalized). 2

In the following specific discussions of the 3D toric code model, the X-cube model

and the Haah code, we will show that we can select a region A and a cut on the TNS

such that |{t}〉A and |{t}〉A are not only orthogonal, but also normalized. In particular

for the 3D toric code model and the X-cube model, we can just select the region A to

be a cube which satisfies the SVD condition directly. See respectively Sec. 6.3.4 and

6.4.4 for detailed discussion of these two models and the SVD condition. However,

the Haah code is different: a cubic region A does not fulfill the SVD condition, and

106

in Sec. 6.5.3 we generalize the SVD condition to the Generalized SVD Condition and

apply Bc operators to make the TNS an SVD.

6.2.2 Summary of the results

We now summarize the major results derived in this chapter for the three stabilizer

codes. Fundamentally, our calculations come down to the fact that the indices of the

nonzero elements of the local tensor T and g are constrained. More specifically, when

we calculate the entanglement entropies with a TNS which is an exact SVD, the only

task is to count the number of independent Schmidt states |{t}〉A. The number of

independent Schmidt states |{t}〉A is determined by the Concatenation lemma,

i.e., when a network of T tensors and g tensors are concatenated, the open indices of

the nonzero elements of the resulting tensors are constrained as well.

1. The TNS is the exact SVD for the ground states with respect to particular

entanglement cuts. The entanglement spectra are flat for models studied in

this chapter.

2. The entanglement of TNS is bounded by the area law:

S ≤ Area× log(D),

where D is the virtual index dimension and Area is measured in the units of

vertices. For the models studied in this chapter, the entanglement entropies are

strictly smaller than the area law when one is computing in terms of vertices.

For the toric code, the correction is a negative constant, − log(2). For the X-

cube model and Haah code, the correction includes a negative term linear with

the system size, presented in Sec. 6.4.4 and 6.5.4.

107

3. The transfer matrices in Eq. (6.18) of the 3D toric code model and the X-cube

model are shown to be a projector whose eigenvalues are either 0 or 1. For the

3D toric code, the transfer matrix in the xy-plane is a projector of rank 2. For

the X-cube model, the transfer matrix is a projector of rank 2Lx+Ly−1 where Lx

and Ly are the lattice sizes in x- and y- directions respectively.

4. We prove that the TNS ground states obtained on the torus using our construc-

tion are the +1 eigenstates of loop X operators. Hence, our TNS construction

does not include all ground states on the torus. The degeneracy of the corre-

sponding transfer matrix is smaller than the GSD on the torus. We can obtain

all the ground states with loop/surface Z operators on the TNS, which gener-

ate all the wave functions on the torus. We call the TNS with Z operators,

“twisted TNS”. Correspondingly, we also obtain more transfer matrices in the

xy-plane built from the twisted TNS, and these transfer matrices are all the

same projectors. The same TNS phenomenon in the 2D toric code model has

been studied in Ref. [178].

5. In our calculations, both the transfer matrix eigenvalue degeneracies, and the

corrections to the area law of entanglement entropies are rooted in the Con-

catenation lemma. Hence, we believe that the two contributions are related.

Specifically, suppose we consider our TNS on a 3D cylinder T 2xy ×Rz, and the

entanglement cut splits the system in two halves z > 0 and z < 0. Then,

for the toric code model, the transfer matrix TMxy has the degeneracy 2, and

the entanglement entropy correction to the area law is − log(2). For the X-

cube model, the transfer matrix TMxy has the degeneracy 2Lx+Ly−1, and the

entanglement entropy correction to the area law is −(Lx + Ly − 1) log(2) (See

Eq. (6.101)). Moreover, the GSD on T 3 is generally larger than the transfer

matrix degeneracy. Therefore, given these calculations, we conjecture that the

108

negative linear correction to the area law is a signature of the extensive ground

state degeneracy.

6.3 3D Toric Code

In this section, we construct the TNS for the 3D toric code model and then calculate

the entanglement entropy and GSD, both deriving from the Concatenation lemma.

The results are the immediate generalizations of those in the 2D toric code model. We

find a topological entanglement entropy in accordance to that obtained by Ref. [190]

using a field theoretic approach.

This section is organized as follows: In Sec. 6.3.1, we briefly review the toric code

model in a cubic lattice. In Sec. 6.3.2, we construct the TNS for the toric code model.

In Sec. 6.3.3, we prove a Concatenation lemma for toric code TNS, which is useful

in the following calculations. In Sec. 6.3.4, we calculate the entanglement entropies

on R3. In Sec. 6.3.5, we construct the transfer matrix and prove it is a projector of

rank 2. In Sec. 6.3.6, we show how to construct 8 ground states on torus by twisting

the TNS.

6.3.1 Hamiltonian of 3D Toric Code Model

The 3D toric code model can be defined on any random lattice. However, for sim-

plicity, we only work on the cubic lattice. On a cubic lattice, the physical spins are

defined on the bonds of the lattice, and the Hamiltonian is built from two types of

terms:

H = −∑v

Av −∑p

Bp. (6.30)

109

Z ZZ

Z

x

x

xx

x

z

y

Z

Z

(a) (b)

Figure 6.5: The Hamiltonian terms of the 3D toric code model. Panel (a) is Av whichis a product of 6 Z operators, and Panel (b) is Bp which is a product of 4 X operators.The circled X and Z represent the Pauli matrices acting on the spin-1/2’s. The toriccode Hamiltonian includes Av terms on all vertices v and Bp terms on all plaquettesp.

where Av is defined around a vertex v, and Bp is defined on a plaquette p:

Av =∏i∈v

Zi, Bp =∏i∈p

Xi, (6.31)

where Zi and Xi are Pauli matrices for the i-th spin. On a cubic lattice, Av is

composed of 6 Pauli Z operators while Bp is composed of 4 Pauli X operators. These

two terms are depicted in Fig. 6.5. In the 2D toric code, Av is composed of 4 Pauli

Z operators on a square lattice. The Hamiltonian is the sum of Av operators on all

vertices v and Bp operators on all plaquettes p.

It is easy to verify that all the Hamiltonian terms commute:

[Av, Av′ ] = 0, ∀ v, v′

[Bp, Bp′ ] = 0, ∀ p, p′

[Av, Bp] = 0, ∀ v, p,

(6.32)

and their eigenvalues are ±1:

A2v = 1, B2

p = 1. (6.33)

110

The ground states |GS〉 should satisfy:

Av|GS〉 = |GS〉, ∀ v

Bp|GS〉 = |GS〉. ∀ p(6.34)

These two sets of equations are enough to derive the local T tensor and to construct

TNS for the toric code model. In particular, one of the ground states on the torus

that we will find is

|ψ〉 =∏v

1 + Av2|0x〉, (6.35)

where |0x〉 is the tensor product of all X = 1 eigenstates defined on each link.

6.3.2 TNS for 3D Toric Code

We first introduce a projector g tensor Eq. (6.5) on each bond of the lattice. Both

the virtual indices and the physical indices take two values, 0 and 1. The projector g

tensor satisfies:

Z = =

x x=x

. (6.36)

In terms of algebraic equations, these diagrams correspond to:

gsi,j(−1)s = gsi,j(−1)i = gsi,j(−1)j

g1−si,j = gs1−i,1−j.

(6.37)

111

These two sets of equations are true, because (1) the indices s, i and j are identified

for nonzero gsi,j, (2) the nonzero gsi,j are always 1 according to Eq. (6.5). We can

use these conditions to transfer the action of the physical operators to the virtual

operators. Now we introduce an additional T tensor on each vertex of the cubic

lattice, which has six virtual indices. Graphically, we represent such a T tensor as:

Txy

z

z

y

x . (6.38)

Next we need to fix the elements of the T tensor, up to the TNS gauge freedom.

The method to fix the T tensor is to make it invariant under the actions of Av and

Bp operators, in order to implement the local conditions for the ground states in

Eq. (6.34). The actions of Av and Bp operators on the local tensors are:

TZ

ZZ

Z

Z

Z

Z

Z Z

ZZ

Z

g

gg

g

g

g

=

Tg

gg g

x

x

x x

x xx

xx x

x

x=

. (6.39)

where we have used Eq. (6.36) to transfer the physical operators to the virtual ones.

We require a strong version of the solution to the above equations. We want the

tensors in the dashed red rectangles to be invariant under the actions of any of the

Av and Bp (this is a sufficient constraint which guarantees that the tensors form the

112

ground state), which leads to the following equations:

T TZ Z

Z

Z=

T T

x x

T

x

xT

x

x

T

x

xT

xx

Tx

x

Tx

Tx

xx

= = =

=

=

==

Z

Z

=

Tx

x T

xx=== =

Tx

x

Tx

T

x

x

x

===T

xxTx

x

(6.40)

In the second set of equations, the first 12 equalities are obvious from the red dashed

squares, and the last 3 equalities can be derived from the first 12 ones. Expanding

the first set of conditions by using Zij = δij(−1)i, we have:

Txx,yy,zz = (−1)x+x+y+y+z+zTxx,yy,zz

Txx,yy,zz

= 0, if x+ x+ y + y + z + z = 1 mod 2

6= 0, if x+ x+ y + y + z + z = 0 mod 2,

(6.41)

where x, x, y, y, z, z are the six indices of T in the three directions respectively. We

emphasize for clarity that x is not −x; these are notations for different indices. The

second set of conditions in Eq. (6.40) further enforces that an even number of index

flipping of the virtual indices of a tensor does not change the value of the tensor

113

elements. For instance, in terms of components, we have:

Txx,yy,zz =T(1−x)(1−x),yy,zz

=T(1−x)x,(1−y)y,zz

=Txx,yy,(1−z)(1−z)

= . . .

. (6.42)

Hence, the nonzero elements of the T tensor are all equal. Up to an overall normal-

ization, we have the unique solution:

Txx,yy,zz =

0, if x+ x+ y + y + z + z = 1 mod 2

1, if x+ x+ y + y + z + z = 0 mod 2.

(6.43)

The TNS is then Eq. (6.6) with the local T being Eq. (6.43). The local T tensors are

the same on other spatial manifolds, such as T 3.

A similar set of conditions as the first equality in Eq. (6.40) have been intro-

duced by several other names in tensor network literature: Z2-injectivity[30], MPO-

injectivity[196], Z2 gauge symmetry[38] etc. The previous studies were in 2D, and

our condition is the 3D generalization. Notice that the first equation in Eq. (6.40)

alone will not necessarily lead to topological order. It only implies that the ground

state is Z2 symmetric. The state which only satisfies the first condition in Eq. (6.40)

could also be a topological trivial state by tuning the relative strength of the nonzero

elements of T tensor. This can be interpreted as a condensation transition from topo-

logical phases to trivial phases. See Refs. [38, 197, 198, 199, 200] for explanations and

examples in the case of 2D TNS.

114

T

T

x

z

y

Figure 6.6: Contraction of two local T tensors in the z-direction. We emphasize thatthere is no projector g tensor in this figure.

6.3.3 Concatenation Lemma

In this section, we consider the contraction of a network of local T tensors with

open virtual indices. One example of such a contraction is the tensor network norm

Eq. (6.15) or the transfer matrix Eq. (6.21). Since the elements of a local T tensor

are 0 for the odd sector and 1 for the even sector (see Eq. (6.43)), we will show that,

in general, a network of contracted T tensors obeys a similar rule: some elements

are zeros while the others are nonzero and identical. A Concatenation lemma is

proposed to derive the rule for the contraction of several tensors in general and will

be frequently used in the following discussions. For example, we will use this lemma

to show in Sec. 6.3.5 that the transfer matrix TMxy for the 3D toric code model is a

projector of rank 2.

Concatenation Lemma: For a network of contracted T tensors Eq. (6.43)

with open indices, the open indices need to sum to 0 mod 2, otherwise the element

of the network tensor is zero. Moreover, if nonzero, the elements of the network

tensor are constants, independent of open indices.

This lemma can be easily proved by using Z2 symmetry Eq. (6.43) and induction.

The proof is in App. C.1. We explain this lemma by a simple example. Suppose we

115

have two T tensors contracted over a pair of indices:

Tx1,x1,y1,y1,z1,x2,x2,y2,y2,z2 =∑z1,z2

Tx1x1,y1y1,z1z1Tx2x2,y2y2,z2z2δz1z2 . (6.44)

Graphically, the tensor T is represented by Fig. 6.6. The open indices of the tensor

T need to sum to an even number in order for the elements of the T tensor to be

nonzero. This comes out of writing the constraints of each of the T tensors:

x1 + x1 + y1 + y1 + z1 + z1 = 0, mod 2

x2 + x2 + y2 + y2 + z2 + z2 = 0, mod 2

z1 = z2

⇒ x1 + x1 + y1 + y1 + z1 + x2 + x2 + y2 + y2 + z2

=0, mod 2.

(6.45)

Otherwise, the tensor element of T is zero. Moreover, the elements of the contracted

tensor are 1, if nonzero:

Tx1,x1,y1,y1,z1,x2,x2,y2,y2,z2 =

0 if x1 + x1 + y1 + y1 + z1 + x2 + x2 + y2 + y2 + z2 = 1, mod 2

1 if x1 + x1 + y1 + y1 + z1 + x2 + x2 + y2 + y2 + z2 = 0, mod 2.

(6.46)

For a more complicated contraction of T tensors, we have:

T{t} =

0 if

∑i ti = 1, mod 2

Const if∑

i ti = 0, mod 2

(6.47)

where {t} denotes all the indices of the tensor T. We emphasize that the nonzero

constant does not depend on {t} .

116

6.3.4 Entanglement

We now show that Eq. (6.6) is exactly an SVD for the wave function with respect to

the entanglement cut illustrated in Fig. 6.7. For simplicity, suppose that the TNS is

defined on infinite R3. As we have emphasized at the end of Sec. 6.1.2, we do not

specify the boundary conditions of the TNS, since we are only concerned with the

bulk wave functions whose reduced density matrices are assumed not to be influenced

by the boundary conditions. If we put the wave function on a large but finite R3,

we have to specify the boundary conditions of the TNS by fixing the indices on the

boundary. Suppose the open indices on the boundary are denoted as {tb}. The norm

of the TNS on open R3, which can be expressed as a network of contracted T tensors

with open virtual indices {tb}, is zero when∑

i tbi = 1 mod 2 and nonzero when∑

i tbi = 0 mod 2, according to the Concatenation lemma of the 3D toric code

model in Sec. 6.3.3. Hence, we can only fix the boundary indices {tb} to be∑

i tbi = 0

mod 2. Calculating the entanglement on a nontrivial manifold is ambiguous since

multiple degenerate ground states, which cannot be distinguished locally, appear.

Their superpositions have different entanglement entropies.

We rewrite Eq. (6.6) by separating the tensor contractions to a spatial region A

and its complement region A. Region A contains the g tensors near the entanglement

cut as illustrated in Fig. 6.7:

|TNS〉R3 =∑{t}

|{t}〉A ⊗ |{t}〉A (6.48)

where

|{t}〉A =∑{s}∈A

∑{i}∈A

CA(gs1t1i1gs2t2i2

. . . gs3i3i4gs4i5i6Ti7...Ti8... . . .)|{s}〉. (6.49)

117

Indices denoted by s are the physical indices; indices denoted by t are the open virtual

indices straddling the entanglement cut from the region A; indices denoted by i are

the contracted virtual indices inside the region A. The tensors gs1t1i1 and gs2t1i2 etc are

the projector g tensors near the entanglement cut on the region A side as illustrated

in Fig. 6.7; gs3i3i4 and gs4i5i6 are the projector g tensors inside the region A; for this cut,

all the T tensors are inside the region A. The summation is over all physical indices

{s} inside the region A.

Thereby, |{t}〉 is the TNS for region A with open virtual indices {t}. We choose

a convention of splitting tensors whereby g tensors near the entanglement cut belong

to the region A, as illustrated in Fig. 6.7. For instance, when the region A is a cube,

we can graphically denote the basis |{t}〉 as Eq. (6.24), where in the bulk of this

cube is a TNS, and the red lines are the outgoing virtual indices {t}. The g tensors

connecting with these red lines are inside the cube. Similarly for the region A:

|{t}〉A =∑{s}∈A

∑{i}∈A

CA(gs1i1i2gs2i3i4Tt1i5...Tt2i6... . . .)|{s}〉. (6.50)

Since the TNSs for region A and A share the same boundary virtual indices {t}, then

in Eq. (6.48) the two basis for region A and A have the same label {t}. For the TNS

of Eq. (6.6), the boundary virtual indices {t} of the regions A and A are contracted

over, and thus in Eq. (6.48) {t} are summed over.

Tg

A Acut

Figure 6.7: The splitting of tensors near the entanglement cut.

118

We now show that |{t}〉A and |{t}〉A are an orthonormal basis (normalized up to

constant) for the region A and the region A respectively. Therefore, Eq. (6.48) is

exactly the SVD for the ground state wave function, i.e.,

A〈{t′}|{t}〉A ∝ δ{t′},{t}δ(∑i

ti = 0 mod 2). (6.51)

Proof:

Applying the SVD condition to the toric code TNS, we can immediately conclude

that the |{t}〉A span an orthogonal basis, and the TNS is exactly an SVD. However,

the SVD condition does not tell us whether the basis is orthonormal. In the follow-

ing, we show that |{t}〉A is not only orthogonal, but also orthonormal with a norm

independent on t, which leads to the flat singular values. Following the definition of

our basis:

A〈{t′}|{t}〉A =

∑{s′}∈A

∑{j}∈A

CA(gs′1?

t′1j1gs′2?

t′2j2. . . g

s′3?j3j4

gs′4?j5j6

T ?j7...T?j8...

. . .)〈{s′}|

∑{s}∈A

∑{i}∈A

CA(gs1t1i1gs2t2i2

. . . gs3i3i4gs4i5i6Ti7...Ti8... . . .)|{s}〉

.

(6.52)

When the open virtual indices {t′} 6= {t}, the overlap is clearly zero, as the spin

configurations on the boundary are different due to the projector g tensors. Hence,

the basis |{t}〉A are orthogonal.

Next we show that A〈{t}|{t}〉A is zero when(∑

ti∈{t} ti

)is odd. Following the

same derivations in Sec. 6.1.3, we have:

A〈{t}|{t}〉A = CA (. . . TTT . . .) (6.53)

with the open virtual indices {t}. The contraction CA is over the T tensors in the

region A. Applying the Concatenation lemma in Sec. 6.3.3, A〈{t}|{t}〉A is zero if

119

the open indices {t} are summed to be 1 mod 2:

∑i

ti = 1 mod 2 ⇒ A〈{t}|{t}〉A = 0. (6.54)

Moreover,

A〈{t}|{t}〉A = Const, when∑i

ti = 0 mod 2. (6.55)

Hence |{t}〉 is orthonormal basis up to an overall normalization factor that can be

obtained by the normalization of |TNS〉. 2

The same proof works for the region A and |{t}〉A. Therefore, we can conclude

that Eq. (6.48) is indeed an SVD, and the singular values are all identical. Hence,

for a entanglement cut, we only need to count the number of singular vectors in

Eq. (6.48). For a connected entanglement surface with N open virtual indices, the

number of singular vectors in Eq. (6.48) is 2N−1, because the open virtual indices

need to sum to be 0 mod 2. Hence, the entanglement entropy for a region whose

entanglement surface is singly connected is:

S = N log(2)− log(2). (6.56)

If the entanglement surface still has N open virtual indices but is separated into n

disconnected surfaces, then the entanglement entropy is:

S = N log(2)− n log(2) = Area× log(2)− n log(2). (6.57)

The above is true because the condition that the open indices need to have an even

summation holds true for each component of the entanglement cut. Furthermore, if

we place our TNS ground state on a 3D cylinder T 2xy ×Rz, and the entanglement cut

splits the cylinder into two halves z > 0 and z < 0, then the entanglement entropy of

120

either side is also S = Area × log(2) − log(2). The results can be easily generalized

to ZK lattice gauge models on R3:

S = Area× log(K)− n log(K) (6.58)

with the same equation holding on a cylinder T 2xy ×Rz. The entanglement spectrum

is also flat. The area is measured by the number of open virtual indices straddling

the entanglement cut.

Following the same logic, for the toric code in (d + 1) dimensions, all the open

virtual indices of region A, {ti}, have to satisfy a single constraint∑

i ti = 0 mod 2,

because they have to obey the Concatenation lemma in Sec. 6.3.3. If there are N

open virtual indices on the surface of region A, there are N − 1 independent open

virtual indices. Hence the rank of the reduced density matrix is still 2N−1, because

each independent open index can take 2 values. The entanglement entropy is

S = N log(2)− log(2). (6.59)

The topological entanglement entropy Stopo[T d−1] is independent of the dimensional-

ity, and it obeys the conjecture presented in Ref. [190]:

exp(−dStopo[T d−1]) = GSD[T d] (6.60)

where GSD[T d] = 2d.

6.3.5 Transfer Matrix as a Projector

The z-direction transfer matrix TMxy in 3D is defined as a tensor network overlap in

the xy-plane, with periodic boundary conditions. The indices in the z direction are

open and not contracted over (see Eq. (6.18) to Eq. (6.22)). In this section, we will

121

show that TMxy for the 3D toric code model is a projector of rank 2. Let us denote

the indices of the transfer matrix as

(TMxy){z},{z} =

x (pbc)z

y (pbc)

z

z

i,j

i,j

, (6.61)

where zi,j and zi,j are the indices at the position (i, j) on the xy-plane. The vector

space of this transfer matrix is of dimension 2LxLy . Suppose the vector space is

spanned by the basis e{z}, where

e{z} =Lx⊗i=1

Ly⊗j=1

ezi,j = ez1,1 ⊗ ez1,2 ⊗ . . . ezLx,Ly . (6.62)

ezi,j is the local “virtual bond Hilbert space” spin |0〉 = |↑〉, |1〉 = |↓〉 basis for

the index zi,j, where i and j are the coordinates of zi,j in the x- and y-directions

respectively. We can consider the matrix multiplication of the transfer matrix TMxy

with an element of the basis e{z}:

TMxy · e{z} =∑{z}

(TMxy){z},{z} (6.63)

where {z} is fixed for the both LHS and RHS. Applying the Concatenation lemma

in Sec. 6.3.3 to (TMxy){z},{z}, we conclude that the terms satisfying

Lx∑i=1

Ly∑j=1

zi,j + zi,j = 0 mod 2 (6.64)

122

contribute equally to the RHS of Eq. (6.63), while the terms satisfying

Lx∑i=1

Ly∑j=1

zi,j + zi,j = 1 mod 2 (6.65)

do not contribute. Therefore, we can rewrite the summation more precisely:

TMxy · e{z} =∑

{z} with∑i zi+zi even

(TMxy){z},{z} ∝∑

{z} with∑i zi+zi even

e{z}. (6.66)

When the {z} satisfies∑

i,j zi,j = 0 mod 2, we have

TMxy · e{z} ∝∑

{z} with∑i zi even

e{z}, (6.67)

while when the {z} satisfies∑

i,j zi,j = 1 mod 2, we have

TMxy · e{z} ∝∑

{z} with∑i zi odd

e{z}. (6.68)

Therefore, TMxy is a projector of rank 2. Hence, it has only two eigenvalues of 1, and

the corresponding unnormalized eigenvectors are:

∑{z} with

∑i zi even

e{z}

∑{z} with

∑i zi odd

e{z}.

(6.69)

123

6.3.6 GSD and Transfer Matrix

We know that the 3D toric code model has three pairs of nonlocal operators along

the cycles of 3D torus:

WX [Cx] =∏i∈Cx

Xi, WZ [Cyz] =∏i∈Cyz

Zi;

WX [Cy] =∏i∈Cy

Xi, WZ [Cxz] =∏i∈Cxz

Zi;

WX [Cz] =∏i∈Cz

Xi, WZ [Cxy] =∏i∈Cxy

Zi.

(6.70)

where Cx is a loop along the cycle of x direction on lattice, Cyz is a closed surface

along the cycles of yz directions on dual lattice, and similarly for the other directions.

The figures for these operators are:

Z

x

x x

ZZ

x

z

y

(6.71)

The commutation relations include:

WX [Cx]WZ [Cyz] = −WZ [Cyz]WX [Cx],

WX [Cy]WZ [Cxz] = −WZ [Cxz]WX [Cy],

WX [Cz]WZ [Cxy] = −WZ [Cxy]WX [Cz].

(6.72)

124

All other combinations of operators commute. Hence, there are 8 degenerate ground

states in total on the torus, assuming that Eq. (6.70) has exhausted all nonlocal

operators.

We can also put our TNS on the 3-torus, i.e., |TNS〉T 3 . It is not hard to verify

using Eq. (6.36) and (6.40) that:

WX [Cx]|TNS〉T 3 = |TNS〉T 3 ,

WX [Cy]|TNS〉T 3 = |TNS〉T 3 ,

WX [Cz]|TNS〉T 3 = |TNS〉T 3 .

(6.73)

As already mentioned in Sec. 6.3.2, |TNS〉T 3 = |ψ〉 where |ψ〉 is defined in Eq. (6.35).

Both |TNS〉T 3 and |ψ〉 are +1 eigenstates of WX operators. However, the transfer

matrix defined by |TNS〉T 3 does not provide 8 fold degenerate eigenvalues, but only

2, as shown in Sec. 6.3.5.

We can act with the WZ [Cyz] and WZ [Cxz] on the TNS by using Eq. (6.36) and

(6.40) to generate all the ground states. The TNSs obtained by this action in terms

125

of a xy-plane of tensors are depicted as below:

Z

Z

Z

Z

Z Z Z Z Z Z Z

T

T

Z Z Z Z Z Z Z

TZ

Z

Z

Z

x (pbc)

y (pbc)

g

g

g

z

(6.74)

The intersection of WZ [Cyz] and WZ [Cxz] with the xy-plane is the line Z operators,

illustrated by the blue circle Z in Eq. (6.74). WZ [Cxy] acts on the xy-plane on the

dual lattice, and thus does not change the transfer matrix at all. We denote these xy-

planes of TNSs in Eq. (6.74) as Tα,βxy (with open indices along the z direction), where

α, β ∈ {0, 1} label whether we have inserted the Z operators in the x and y direction

respectively. The subindex xy in Tα,βxy means that the TNS is on a xy-plane. Clearly,

Tα,βxy are different since they support different holonomies of WX [Cx] operators and

WX [Cy] operators. After obtaining Tα,βxy , we can define four twisted transfer matrices

correspondingly by contracting the physical indices between bra Tα,βxy and ket Tα,β

xy .

The twisted transfer matrices are denoted as TMα,βxy . For instance, TM0,0

xy is the

untwisted transfer matrix in Eq. (6.61).

126

Each of these transfer matrices TMα,βxy is also a projector of rank 2. The reasons

are that (1) the contraction of the projector g tensors between the bra Tα,βxy and ket

Tα,βxy makes the indices in the bra layer and ket layer identical; (2) the Z operators in

the bra and ket layer will cancel each other and produce an identity operator. Hence,

the transfer matrices built from the twisted TNS TM1,0xy , TM0,1

xy and TM1,1xy are equal

to that built from the untwisted TNS TM0,0xy :

TMα,βxy = TM0,0

xy , ∀ α, β ∈ {0, 1} (6.75)

The transfer matrix has degeneracy 2, and it is the same for each of 4 TNSs which

are different. We have 4 × 2 = 8 degenerate eigenstates. We have to emphasize

that Eq. (6.75) holds true, when there are no physical operators in between the bra

and ket layer of Tα,βxy , and the physical indices of projector g tensors are directly

contracted. If a physical operator is inserted, for instance WX [Cx], then the twisted

transfer matrices in the presence of WX [Cx] are NOT the same as the untwisted one

sandwiching the same WX [Cx].

In constructing the TNS on torus, we need to choose the same Tα,βxy for each xy-

plane. If we use different Tα,βxy in each xy-plane to construct a TNS on torus, the

corresponding wave function is no longer a ground state for the 3D toric code model.

More specifically, in the 3D toric code model, if we contract Lz − 1 T0,0xy ’s with one

twisted Tα,βxy Eq. (6.74), then the corresponding wave function will support a pair of

loop excitations because the Bp operators up and below Z operators are violated, and

is no longer a ground state. Graphically, the contraction of these Lz − 1 number of

T0,0xy and one T1,0

xy is:

Z Z Z Z

z

. (6.76)

127

Z ZZ

ZZ Z

Z

Z

Z

Z

Z

Z

x

x

x

x

xx

x xx x

xxx

z

y

(a) (b) (c) (d)

Figure 6.8: The Hamiltonian terms of the X-cube model: (a) Av,yz, (b) Av,xy, (c)Av,xz and (d) Bc. The circled X and Z represent the Pauli matrices acting on thephysical spin-1/2’s.

The reason for this energy costing is that there is a line of Z operators. However,

the nonlocal operators in 3D toric code do not have a line Z operator, but only have

surface Z operators. See Eq. (6.71). On the other hand, the TNS made of all the

same Tα,βxy , for instance T1,0

xy :

Z Z Z Z

z

Z Z Z Z

Z Z Z Z

Z Z Z Z(6.77)

is allowed, because it corresponds to a closed surface operator included in Eq. (6.71).

We will come back to this point in Sec. 6.4.6 where this issue is subtle and makes a

difference when we count GSD from transfer matrices.

6.4 X-cube Model

In this section, we construct the TNS for the X-cube model, one of the simplest frac-

ton models. Using it, we then calculate the entanglement entropy and the GSD of

this model. The results are generally different from those in conventional topologi-

cal phases. The entanglement entropy has a linear correction to the area law, and

the GSD grows exponentially with the size of the system. This section is organized

as follows: In Sec. 6.4.1, we briefly review the X-cube model in a cubic lattice. In

Sec. 6.4.2, we construct the TNS for the X-cube model. In Sec. 6.4.3, we prove a

128

Concatenation lemma for the X-cube TNS. In Sec. 6.4.4, we calculate the entan-

glement entropies on R3. In Sec. 6.4.5, we construct the transfer matrix and prove

that it is a projector of rank 2Lx+Ly−1. In Sec. 6.4.6, we show how the transfer matrix

gives rise to an extensive GSD on torus.

6.4.1 Hamiltonian of X-cube Model

The model is defined on a cubic lattice. All the spin 1/2’s are associated with the

bonds of the cubic lattice. The Hamiltonian is:

H = −∑v

(Av,xy + Av,yz + Av,xz)−∑c

Bc (6.78)

where the summation is taken over all vertices and cubes respectively. Each term is

depicted in Fig. 6.8. More specifically, Av,xy is the product of four Z operators around

the vertex v in the xy plane. Similarly for Av,yz and Av,xz. Bc flips the 12 spins of a

cube c. It is trivial to show that all the Hamiltonian terms commute:

[Av,xy, Av′,xy] = [Av,yz, Av′,yz] = [Av,xz, Av′,xz] = 0,

[Av,xy, Av′,yz] = [Av,xy, Av′,xz] = [Av,yz, Av′,xz] = 0,

[Bc, Av′,xy] = [Bc, Av′,yz] = [Bc, Av′,xz] = 0

[Bc, Bc′ ] =0, ∀ v, v′, c, c′.

(6.79)

Hence, this model can be exactly solved. The ground state |GS〉 (on R3) needs to

satisfy that:

Av,xy|GS〉 =|GS〉,

Av,yz|GS〉 =|GS〉,

Av,xz|GS〉 =|GS〉,

Bc|GS〉 =|GS〉, ∀ v, c.

(6.80)

129

This set of equations will be used to derive the local T tensor for the X-cube model.

The nonlocal operators of the X-cube model which are required to commute with

all Hamiltonian terms on torus include 9 loop operators[175]:

WX [Cx] =∏i∈Cx

Xi,WX [Cy] =∏i∈Cy

Xi,WX [Cz] =∏i∈Cz

Xi,

WZ [Cx,z] =∏i∈Cx

Zi,WZ [Cy,z] =∏i∈Cy

Zi,WZ [Cz,x] =∏i∈Cz

Zi,

WZ [Cx,y] =∏i∈Cx

Zi,WZ [Cy,x] =∏i∈Cy

Zi,WZ [Cz,y] =∏i∈Cz

Zi,

(6.81)

where Cx is a straight line along the cycle of the x direction on lattice, and Cx,z is a

line along the cycle of the x direction on dual lattice while the lattice bonds of Cx,z

are in the z-direction, and similarly for the other directions. The figures for them are:

Z

x

x x

Z

Z

Z

Z

Z

x

z

y

(a) (b) (c)

(e) (d) (f)

(g) (h) (i)

. (6.82)

The algebras of these loop operators are grouped into three independent sets:

1. The operator (a) in Eq. (6.82) anti-commutes with the operator (f) and the

operator (h) when they overlap at one spin. Thus, WX [Cx] anti-commutes with

WZ [Cz,x] and WZ [Cy,x] when they overlap at one spin.

130

2. The operator (b) in Eq. (6.82) anti-commutes with the operator (g) and the

operator (i) when they overlap at one spin. Thus, WX [Cy] anti-commutes with

WZ [Cx,y] and WZ [Cz,y] when they overlap at one spin.

3. The operator (c) in Eq. (6.82) anti-commutes with the operator (d) and the

operator (e) when they overlap at one spin. Thus, WX [Cz] anti-commutes with

WZ [Cx,z] and WZ [Cy,z] when they overlap at one spin.

All other combinations of operators commute, because they do not overlap at the same

physical spin. Each of the three algebras gives a representation of dimension[155]

2Ly+Lz−1, 2Lz+Lx−1, 2Lx+Ly−1. (6.83)

respectively. Hence the total dimension of the ground state space is 22Lx+2Ly+2Lz−3.

The derivations of the GSD in terms of these operator algebras are explained in

App. C.3.

6.4.2 TNS for X-cube Model

Following the same prescription in Sec. 6.3, we can write down the TNS for the X-

cube model. First, we introduce a projector g tensor on the bonds of the lattice (see

Eq. (6.5)). The virtual index is either 0 or 1. The g tensor is a projector which

projects the physical spin to the virtual indices. The tensor g satisfies Eq. (6.36).

The TNS is not only composed of the g tensor on the bonds of the lattice, but also

the T tensors on the vertices. The T tensor has six virtual indices and no physical

index. Unlike the g tensor, the T tensor will be specified by the Hamiltonian terms.

We now implement Eq. (6.80) on the TNS. Using the condition Eq. (6.36), we can

131

transfer the operators in Hamiltonian terms from physical indices to virtual indices:

TZ

Z

Z

Z

Z

Z

ZZ

g

gg

g

g

g

=

ggg

g

gg

g

g

gg

g

T

T T

T

TT

TT

xx

x

xx

x

xxx

xx

xx

xx

xx x

xx

xxx

x xx

x xx

xxx

=

TZ

Z

Z

Z

ZZ

ZZ

g

gg

g

g

g

=

T

ZZ

Z

Z Z

Z Z

Zg

gg

g

g

g

=

x

x x

x

. (6.84)

132

Requiring that the tensors in the dashed red rectangles are invariant will lead to the

following (strong) conditions on the T tensor:

T T T TZ Z

Z

ZZ Z

Z

Z

Z

Z

Z

Z= = =

T T

x xx

T

x xx

Tx

x

x

T

xx

xT

x

xx Tx

x

T

x

xT

xx

x

xx

= = =

= = =

= =

. (6.85)

The first set of conditions is required by the operators Av,xy, Av,yz and Av,xz around

the vertex v. The second set of equations is required by the operators Bc around the

cube c. The X operators acting on the 12 spins of the cube c will be transferred to

the virtual spins of the eight T tensors around the cube c, using Eq. (6.36). The X

operators act on the eight quadrants of a T tensor. Clearly, these two sets of the

conditions are not independent. The solution to these conditions is:

Txx,yy,zz =

1 if

x+ x+ y + y = 0 mod 2,

x+ x+ z + z = 0 mod 2,

y + y + z + z = 0 mod 2.

0 otherwise.

(6.86)

133

A useful consequence of Eq. (6.85) is:

T Tx

T

x

= = =T

xx

xx . (6.87)

We now have fixed the TNS for the X-cube model using the local conditions Eq. (6.80).

The wave function on R3 can also be represented as a tensor contraction of Eq. (6.6).

6.4.3 Concatenation Lemma

In this section, we consider the contraction of a network of local T tensors with open

virtual indices for the X-cube model, similar to the idea developed in Sec. 6.3.3.

However, here the situation is more complicated than the 3D toric code model. The

elements of a local T tensor are either 0 or 1 in Eq. (6.85), depending on the even/odd

sector in three directions. The elements of the contracted T tensors will also be either

0 or 1 with a similar criterion.

Concatenation Lemma: For a network of the contracted T tensors in

Eq. (6.86), the sums of the open indices along each xy, yz and xz planes have

to be even. Otherwise, the tensor element of this network is zero. The nonzero

elements are constants independent of the virtual indices.

This lemma is a consequence of Eq. (6.86). See App. C.2 for an induction proof.

In this section, we explain this result by considering a simple example. Suppose that

we have two T tensors contracted along the z direction:

Tx1,x1,y1,y1,z1,x2,x2,y2,y2,z2 =∑z1,z2

Tx1x1,y1y1,z1z1Tx2x2,y2y2,z2z2δz1z2 . (6.88)

134

Graphically, T is the same as depicted in Fig. 6.6. For each of the two T tensors, we

have x1 + x1 + y1 + y1 = 0 mod 2

x1 + x1 + z1 + z1 = 0 mod 2

y1 + y1 + z1 + z1 = 0 mod 2

(6.89)

and x2 + x2 + y2 + y2 = 0 mod 2

x2 + x2 + z2 + z2 = 0 mod 2

y2 + y2 + z2 + z2 = 0 mod 2.

(6.90)

Therefore, setting z1 = z2 due to the tensor contraction, the open indices of T need

to satisfy:

x1 + x1 + y1 + y1 = 0, mod 2

x2 + x2 + y2 + y2 = 0, mod 2

z1 + z2 + x1 + x1 + x2 + x2 = 0, mod 2

z1 + z2 + y1 + y1 + y2 + y2 = 0, mod 2

(6.91)

in order for the elements of the tensor T to be nonzero. The last set of equations

intuitively means that the open indices of the tensor T (Fig. 6.6) in each xy, yz and

xz plane need to have an even summation. Moreover, the elements of the contracted

135

A A A A

(1) (2) (3) (4)

Figure 6.9: Figures for several regions A for which we calculate the entanglemententropies. (1) Region A is a cube of size lx × ly × lz. (2) Region A is a cube oflx× ly× lz with a hole of size l′x× l′y× l′z in it. (3) Region A is a cube of size lx× ly× lzand a small cube of height l′z on top of it. (4) Region A is a cube of size lx × ly × lzcarved on top of it a small cube of height l′z.

T tensor are 1 independent of indices:

Tx1,x1,y1,y1,z1,x2,x2,y2,y2,z2 =

1 if

x1 + x1 + y1 + y1 = 0, mod 2 and

x2 + x2 + y2 + y2 = 0, mod 2 and

z1 + z2 + x1 + x1 + x2 + x2 = 0, mod 2 and

z1 + z2 + y1 + y1 + y2 + y2 = 0, mod 2

0 otherwise.

(6.92)

Generally for a complicated contraction of local T tensors denoted by T, we have:

T{t} =

Const 6= 0 Concatenation lemma is satisfied

0 else.

(6.93)

This is the notation that we will use when computing the entanglement entropies or

the transfer matrix degeneracies.

6.4.4 Entanglement

We can show that a cubic region A and its deformations are the exact SVD of |TNS〉.

See Fig. 6.9 for some examples of deformed A regions. We can read out the singular

136

values of |TNS〉 for these entanglement cut. Suppose we consider the wave function

on R3 for simplicity (i.e., without dealing with the multiple ground states on T 3).

We rewrite the wave function:

|TNS〉 =∑{t}

|{t}〉A ⊗ |{t}〉A. (6.94)

where {t} represent the tensor virtual indices connecting the region A and A. |{t}〉A

is the TNS for the region A with open virtual indices {t}. Similarly for |{t}〉A for

the complement region A. Because every virtual bond is contracted over for |TNS〉,

|{t}〉A and |{t}〉A have to share the same {t}, and {t} is summed over in Eq. (6.94).

Next we show that this set of basis |{t}〉A is orthonormal:

A〈{t′}|{t}〉A ∝ δ{t′},{t}, (6.95)

when the Concatenation lemma is satisfied for both of {t′}, {t}, and the basis

|{t}〉A are not null vectors.

Proof:

The open virtual indices {t} satisfy the SVD condition in Sec. 6.2.1. Hence, we

can conclude that |{t}〉A and |{t}〉A are orthogonal basis for the region A and A, and

thus Eq. (6.94) is exactly SVD. In order to calculate the entanglement entropies, we

need to show that |{t}〉A and |{t}〉A are not only orthogonal, but also orthonormal.

The proof is essentially the same as in Sec. 6.3.4. Here we briefly repeat it. We

use the same conventions of SVD basis |{t}〉A and |{t′}〉A as in Sec. 6.3.4. Suppose

{t′} and {t} both satisfy Concatenation lemma in Sec. 6.4.3. Hence, |{t}〉A and

|{t′}〉A are not null vectors. More specifically, the wave function |{t}〉A is the same

as in Eq. (6.49). All the virtual indices except {t} are contracted over.

137

When {t′} 6= {t}, the basis overlap is zero, because the spins on the boundary

of A are different. When {t′} = {t}, the overlap is nonzero. Moreover, the overlaps

A〈{t}|{t}〉A are constants independent of {t}, due to the Concatenation lemma

in Sec. 6.4.3. Hence, |{t}〉A is an orthonormal basis for the region A, up an overall

normalization factor. 2

Therefore, using the orthonormal basis |{t}〉A and |{t}〉A, Eq. (6.94) is indeed the

SVD. Furthermore, the singular values are all identical. As a result, the entanglement

of the region A is determined by the number of basis states |{t}〉A which are involved

in Eq. (6.94), i.e., the rank of the contracted tensor for the region A. The rank of the

contracted tensor can be counted by the Concatenation lemma in Sec. 6.4.3. We

only need to count the number of indices that satisfy the Concatenation lemma.

We now list a few simple examples of entanglement entropies. The entanglement cuts

are displayed in Fig. 6.9. Correspondingly, their entanglement entropies are:

1. When region A is a cube of size lx × ly × lz:

SAlog 2

= 2lxly + 2lylz + 2lxlz − lx − ly − lz + 1 = Area− lx − ly − lz + 1.

(6.96)

The calculation details are the following:

The number of indices straddling this entanglement cut is 2lxly + 2lylz + 2lxlz.

This is the maximum possible number of basis states in the SVD of Eq. (6.94).

However, these indices are not free. They are subject to certain constraints, in

order for the singular vectors to have non-vanishing norms. Using the Con-

catenation lemma in Sec. 6.4.3, we know that the open indices in each xy,

yz, and xz plane must have even summations. We denote the indices to be ti,j,k

138

where i, j, k are the coordinates of such a index. Then, we have;

∑i,j

ti,j,k = 0 mod 2, ∀ k

∑i,k

ti,j,k = 0 mod 2, ∀ j

∑j,k

ti,j,k = 0 mod 2, ∀ i

(6.97)

where the summation is only taken over open virtual indices near the entangle-

ment cut in each xy, yz and xz plane. Therefore, we have lx, ly, lz number of

constraints respectively. However, these constraints are not independent. Only

lx + ly + lz − 1 of them are independent, because the three sets of constraints

sum to be an even number. Hence, the number of free indices is

2lxly + 2lylz + 2lxlz − lx − ly − lz + 1.

The number of singular vectors in Eq. (6.94) is:

22lxly+2lylz+2lxlz−lx−ly−lz+1,

which leads to the entropy written in Eq. (6.96). 2

2. When the region A is a cube of size lx × ly × lz with an empty hole of size

l′x × l′y × l′z:

SAlog 2

=2lxly + 2lylz + 2lxlz + 2l′xl′y + 2l′yl

′z + 2l′xl

′z

− l′x − l′y − l′z − lx − ly − lz + 2

=Area− l′x − l′y − l′z − lx − ly − lz + 2.

(6.98)

139

3. When the region A is a cube of size lx × ly × lz with an additional convex cube

of height l′z on top (Fig. 6.9 (3)):

SAlog 2

=Area− lx − ly − lz − l′z + 1. (6.99)

4. When the region A is a cube of size lx× ly× lz with an additional concave cube

of height l′z on top (Fig. 6.9 (4)):

SAlog 2

=Area− lx − ly − lz − l′z + 1. (6.100)

The area part of the entanglement is measured by the number of indices straddling

the entanglement cut. The constant contribution to the entanglement entropy is

universal[201], as it counts the number of connected components of the entanglement

surface. As opposed to the toric code case, the constants are positive numbers. We

emphasize that the linear corrections to the area law in the entanglement entropies

states have not been observed in quantum field theories.

Furthermore, if we put the TNS on a cylinder T 2xy×Rz and the entanglement cut

splits the cylinder into two halves z > 0 and z < 0, then the entanglement entropy

of either side is:

SAlog 2

= Area− Lx − Ly + 1. (6.101)

We emphasize that the entanglement spectrum is flat, because all singular values are

identical in Eq. (6.94).

140

Tg

x x x x x x x

Tg

x x x x x x x x x x x x

Tg

x x x x x x x x x x x x

Tg

x (pbc)

z

y (pbc)

Figure 6.10: Derivations for the first equation in Eq. (6.107). The rest two equationscan be proved similarly. In the first step, the physical X operators can be transferredto the virtual level by using Eq. (6.36), and in the third step, all the virtual Xoperators are exactly canceled in pairs (dashed red rectangles in the third figure) dueto Eq. (6.87).

6.4.5 Transfer Matrix as a Projector

Following the same reasoning explained in Sec. 6.1.4, the transfer matrix of the X-cube

model in the xy-plane is:

TMxy = CT 2xy (. . . TTT . . .) (6.102)

with open virtual indices in the z-direction. Graphically, see Eq. (6.22) and Eq. (6.61).

In this section, we will show that the TMxy for the X-cube model is also a projector.

However, the projection is more complicated than in the 3D toric code example.

Using the transfer matrix basis e{z} defined in Eq. (6.62), we show that:

TMxy · e{z} =∑

Concatenation Lemma

(TMxy){z},{z} ∝∑

Concatenation Lemma

e{z} (6.103)

141

where the notation {z} and {z} collectively denote all z indices perpendicular to the

xy-plane. The summation with the Concatenation lemma in Sec. 6.4.3 means that:

∑i

zi,j + zi,j = 0 mod 2, ∀ j

∑j

zi,j + zi,j = 0 mod 2, ∀ i(6.104)

where the subindex i, j of zi,j denote the positions of zi,j in the x- and y-direction

respectively. These two equations mean that in each xz and yz plane, the indices

have an even summation. For instance, the red dashed rectangles below:

x (pbc)

y (pbc)

TT

. (6.105)

Among the Lx+Ly equations in Eq. (6.104), only Lx+Ly−1 are linearly independent,

because the summations of the two sets of constraints are the same:

∑j

(∑i

zi,j + zi,j

)= 0 mod 2⇔

∑i

(∑j

zi,j + zi,j

)= 0 mod 2. (6.106)

The summation in Eq. (6.103) can be separated into Lx +Ly − 1 number of different

“parity” sectors, similar to the 3D toric code case Eq. (6.69). Hence, TMxy is a

projector of rank 2Lx+Ly−1. It has 2Lx+Ly−1 degenerate nonzero eigenvalues.

6.4.6 GSD and Transfer Matrix

The TNS, which gives us the single ground state on R3, has the minimum energy by

construction. If we contract these tensors on the torus T 3 with periodic boundary

142

conditions, then we still yield only one ground state. Moreover, this ground state is

the +1 eigenstate of all WX operators in Eq. (6.81):

WX [Cx]|TNS〉T 3 = |TNS〉T 3 ,

WX [Cy]|TNS〉T 3 = |TNS〉T 3 ,

WX [Cz]|TNS〉T 3 = |TNS〉T 3 .

(6.107)

These three equations can be proved by using Eq. (6.36) and Eq. (6.85); the deriva-

tions are summarized in Fig. 6.10.

Other ground states on the torus can be obtained by acting with the nonlocal

operators WZ [Cz,x] and WZ [Cz,y] in Eq. (6.81) on the TNS |TNS〉T 3 . The physi-

cal operator WZ [Cz,x] and WZ [Cz,y] can be transferred to the virtual indices using

Eq. (6.36).

After applying WZ [Cz,x] and WZ [Cz,y] in Eq. (6.81) on TNS, we can generate

2Lx+Ly TNSs exemplified by Fig. 6.11. The intersections of WZ [Cz,x] and WZ [Cz,y]

with the xy-plane are the blue circled Z operators in Fig. 6.11. We denote these

planes of tensors in Fig. 6.11 as T~α,~βxy where ~α and ~β are binary vectors (values in

{0, 1}) of length Lx and Ly representing the absence or presence of Pauli Z operators.

For instance, the untwisted plane of TNS is T~0,~0xy using this convention. This notation

T~0,~0xy is similar to that in Sec. 6.3.6. Inserting a Z operator at the virtual level will

change the holonomy of WX in the xy-plane. For instance, for Panel (a) in Fig. 6.11,

the WX operator along the first row has a −1 eigenvalue, while the WX operators

along the second, the third and the fourth row have a +1 eigenvalue.

Each of these T~α,~βxy will generate a transfer matrix TM~α,~β

xy which has 2Lx+Ly−1

degenerate eigenvalues. The reasons are that (1) when building the transfer matrices

from the twisted T~α,~βxy , the contraction over the physical indices of the projector g

tensors makes the virtual indices from the bra layer and ket layer identical; (2) the Z

operators in the bra layer cancel their analogues respectively in the ket layer. Hence,

143

Z

Z

x (pbc)

y (pbc)

T

T

Z

Z

T

T

Z

Z

T

T

(a) (b) (c)

(d) (e) (f)

Z

Z

Z

Z

Z

Z Z

Figure 6.11: Examples for the X-cube TNS in a xy-plane, obtained by acting WZ [Cz,x]and WZ [Cz,y] on the constructed TNS. The intersection of one WZ [Cz,x] operator andone WZ [Cz,y] operator with the xy-plane is only one Pauli Z operator, i.e., the circledblue Z in this figure.

Z

x (pbc)

y (pbc)

T

ZT

T

T

Z Z Z Z Z Z Z

(a) (b)

(c) (d)

Z

Z

Z

Z

Z

Z

Z Z Z Z Z Z Z

Figure 6.12: We act WZ [Cx,y] and WZ [Cy,x] operators on Panel (a) in Fig. 6.11. Hence,we have four TNSs in this xy-plane that can be related to each other. See the textfor the explanations.

the transfer matrices TM~α,~βxy obtained from the twisted T~α,~β

xy are equal to that obtained

from the untwisted one T~0,~0xy .

TM~α,~βxy = TM

~0,~0xy , ∀ ~α, ~β (6.108)

Thus, the transfer matrices TM~α,~βxy built from the twisted T~α,~β

xy are also projectors of

dimension 2Lx+Ly−1.

144

The TNS on torus is then constructed as the contraction of T~α,~βxy on each xy-

plane. However, the subtlety of the X-cube model is that in constructing TNS on

the torus, there are still degree of freedom we can play with. In the 3D toric code

case Sec. 6.3.6, as we evolve the state in the z direction, we have to use the same

Tα,βxy in each plane, otherwise the corresponding wave function is no longer a ground

state. However, we have more choices for the X-cube model. In each plane of T~α,~βxy ,

we have four choices that do not change the energy: we can still act WZ [Cx,y] and

WZ [Cy,x] on TNS in the xy-plane without affecting other xy-planes. These choices

do not raise the energy because WZ [Cx,y] and WZ [Cy,x] are the nonlocal operators of

the X-cube model Eq. (6.81): they do not cost any energy. For each of the T~α,~βxy built

from Fig. 6.11, we can find 3 others: all 4 planes of tensors T~α,~βxy can be used in the z

direction. Take Panel (a) of Fig. 6.11 as an example at one point in the z direction.

The 4 T~α,~βxy are depicted in Fig. 6.12. Their expressions are:

1. For Panel (a), we do not apply any operators.

2. For Panel (b), we apply WZ [Cy,x] on TNS.

3. For Panel (c), we apply WZ [Cx,y] on TNS.

4. For Panel (d), we apply both WZ [Cx,y] and WZ [Cy,x] on TNS.

Due to this choice, four twisted T~α,~βxy will be grouped together, and there are

2Lx+Ly

4(6.109)

number of groups of twisted T~α,~βxy . Hence, the total GSD that we can obtain from the

transfer matrices built from T~α,~βxy is:

2Lx+Ly−1 × 2Lx+Ly

4× 4Lz = 22Lx+2Ly+2Lz−3. (6.110)

145

Each factor in the above formula has an explanation:

1. 2Lx+Ly−1 is the degeneracy of each transfer matrix.

2. 2Lx+Ly

4is the number of groups of the twisted T~α,~β

xy due to the “ambiguity”

explained in the paragraph before Eq. (6.109).

3. 4Lz is the number of combinations for T~α,~βxy , since for each xy plane we can pick

any of the 4 T~α,~βxy belonging to the same group.

This is the total GSD of X-cube model on torus.

6.5 Haah Code

In this section, we derive the TNS for the Haah code following a similar prescription

as that in Sec. 6.3.2 and Sec. 6.4.2. We then compute the entanglement entropies

using the TNS for several types of entanglement cuts. This section is organized as

follows. In Sec. 6.5.1, we review the Hamiltonian of the Haah code. In Sec. 6.5.2,

we present the construction of TNS for the Haah code. In Sec. 6.5.3, we discuss the

entanglement cuts for which the TNS is an exact SVD. In Sec. 6.5.4, we discuss the

cubic entanglement cut, where the TNS is not an exact SVD. The calculation of the

entanglement entropies proceeds in the same way as that of the 3D toric code model

or X-cube model: one counts the number of constraints for open indices.

6.5.1 Hamiltonian of Haah code

The Haah code is defined on a cubic lattice. As opposed to the 3D toric code and the

X-cube model discussed in Sec. 6.3 and 6.4, there are two spin-1/2’s defined on each

vertex of a cubic lattice. The Hamiltonian of the Haah code is a sum of commuting

operators where each term is the product of Pauli X or Z operators. Specifically,

146

there are two types of the Hamiltonian terms:

H = −∑a,b,c

Aabc −∑a,b,c

Babc. (6.111)

The A and B operators are defined on each cube in the cubic lattice, and the indices

a, b, c represent the vertex coordinates. If we choose the space to be R3, then a, b, c ∈

Z. If we choose the space to be a 3D torus of the size Lx × Ly × Lz with periodic

boundary condition on each side, then a ∈ ZLx , b ∈ ZLy and c ∈ ZLz . The operators

defined on a = 0, b = 0, c = 0 are

A000 = ZL110Z

L101Z

L011Z

L111Z

R100Z

R010Z

R001Z

R111

B000 = XL000X

L110X

L101X

L011X

R000X

R100X

R010X

R001.

(6.112)

The superscripts L/R represent the left or the right spin on a vertex where the Pauli

operators act on. The subscripts (ijk) ∈ Z2 × Z2 × Z2 represent the coordinate of

vertices (on a cube). The operators Aabc and Babc on all other cubes can be obtained

by translation. Pictorially the two types of terms are:

x

z

y

(6.113)

It is straightforward to check that all the Hamiltonian terms commute.

147

T T

T T

(a)

x

z

y

T T

T T

T T

T T

(b)

Figure 6.13: Tensor contraction for the Haah code TNS. (a) The lattice size is 2×3×3.(b) The lattice size is 3× 3× 3

6.5.2 TNS for Haah Code

The ground state |GS〉 is obtained by requiring

Aabc|GS〉 = |GS〉 (6.114)

Babc|GS〉 = |GS〉 (6.115)

148

for every a, b, c. We can solve these two equations similarly to the 3D toric code model

in Sec. 6.3.2 and the X-cube model in Sec. 6.4.2 to obtain a TNS representation. We

now specify the projector g tensor and the local T tensor.

There are 2 types of g tensors gL and gR associated with the left and right physical

spins on each vertex. Each g tensor has 1 physical index s and 4 virtual indices i, j, k, l.

The reason for these 4 virtual indices (rather than 2 virtual indices as in the toric

code and the X-cube examples) is that, for each vertex, the virtual indices from T

tensors (to be defined below) in the neighboring 8 octants need to be fully contracted;

this requires the g tensor to have 4 virtual indices. The index assignment of the left

and right g tensors, gLsijkl and gRsijkl, are:

gLsijkl =

s

i

j

k l

I

IIIII

IV

V

V IV II

V III

(6.116)

and

gRsijkl =

l

j

ki

s

I

IIIII

IV

V I

V

V II

V III

(6.117)

where s is the physical index in {|0〉 = |↑〉, |1〉 = |↓〉}, and ijkl are virtual indices.

We use a blue dot for the g tensor on the right spin and a red dot for the g tensor on

the left spin. The green dots at the center of each cube represent T tensors (which we

define below). Similar to the toric code model and the X-cube model, we require that

149

the g tensor acts as a projector from the physical index to the four virtual indices:

gLsijkl =

1 i = j = k = l = s

0 otherwise

,

gRsijkl =

1 i = j = k = l = s

0 otherwise

.

(6.118)

The four virtual indices of gLsijkl extend along the III, VIII, VII, VI octants (as shown

in Eq. (6.116)), and the four virtual indices of gRsijkl extend along the II, VII, IV, V

octants (as shown in Eq. (6.117)).

The tensor T{i} is defined at the center of each cube, and every T tensor has 8

virtual indices. Graphically, the T tensor is:

Ti1i2i3i4i5i6i7i8 = T . (6.119)

The T tensor is contracted to 8 of the total 16 (8 vertices times 2 degrees of freedom

per vertex) g tensors located at the cube corners via the virtual indices. The reason

for only 8 virtual indices (instead of 16 virtual indices) in the T tensor is that among

16 spins around the cube (a, b, c) only eight of them are addressed by the Pauli Z

operators in the Aabc term of the Hamiltonian. The elements of the T tensor for

a given set of virtual indices i1i2i3i4i5i6i7i8 are determined by solving Eq. (6.114)

and Eq. (6.115). Imposing the condition Eq. (6.114) and transferring the physical Z

150

operators to the virtual level, we find that:

TZ

Z Z

Z Z

Z

ZZ

= T (6.120)

which amounts to

Ti1i2i3i4i5i6i7i8 = (−1)∑8n=1 inTi1i2i3i4i5i6i7i8 , (6.121)

where i1, · · · , i8 are the eight virtual indices of the T tensor defined in Eq. (6.119).

Hence,

Ti1i2i3i4i5i6i7i8 = 0, if8∑

n=1

in = 1 mod 2. (6.122)

151

Imposing the condition Eq. (6.115) and transferring the physical X operators to the

virtual level, we find that

T = Tx x

= T

x

x

= T

x

x

= T

x

x = Tx x

= T

x x

= T

x x

= T

x

x

= T

x

x = T

x

x

= Tx

x

= Tx

x

= Tx

x

= T

x x

x

x

xx

.

(6.123)

In terms of components, Eq. (6.123) means that Ti1i2i3i4i5i6i7i8 = Ti′1i′2i′3i′4i′5i′6i′7i′8 where

i′1i′2i′3i′4i′5i′6i′7i′8 are obtained by flipping arbitrary pairs of indices from i1i2i3i4i5i6i7i8.

152

For example,

Ti1i2i3i4i5i6i7i8 = T(1−i1)(1−i2)i3i4i5i6i7i8

= Ti1(1−i2)(1−i3)i4i5i6i7i8

= Ti1i2(1−i3)(1−i4)i5i6i7i8

= ...

(6.124)

Combining Eq. (6.121) and (6.124), we find that any configuration of Ti1i2i3i4i5i6i7i8

satisfying the condition∑8

k=1 ik = 0 mod 2 are equal. We can rescale the T tensor

such that Ti1i2i3i4i5i6i7i8 = 1 for∑8

k=1 ik = 0 mod 2, i.e.,

Ti1i2i3i4i5i6i7i8 =

1∑8

n=1 in = 0 mod 2

0∑8

n=1 in = 1 mod 2.

(6.125)

For simplicity, we consider the space to be R3 where the Haah code has a unique

ground state.

|TNS〉 =∑{s}

CR3 (gL,s1gR,s2gL,s3gR,s4 . . . TTT . . .

)|{s}〉. (6.126)

We emphasize that the contraction of the Haah code TNS is quite different from

that of the 3D toric code model and the X-cube model. The main difference is that

the g tensor has 4 virtual indices for the Haah code, while it has only 2 virtual indices

for the 3D toric code and the X-cube model. As an example of contraction, we take

two blocks of size 2×2×1 and 2×2×2 in Fig. 6.13. The T tensors with their virtual

indices are drawn explicitly. Each red or blue node in the two figures is a projector

g tensor, whose physical index is not drawn; we only draw the virtual legs that are

connected to the T tensors inside the blocks. In the block 2× 2× 2, all the 8 virtual

indices of the two g tensors (4 per each g tensor) in the middle of all the cubes are

153

contracted with T tensors, while other g tensors have open virtual indices (which are

not explicitly drawn).

6.5.3 Entanglement Entropy for SVD Cuts

In this section, we compute the entanglement entropies for two types of cuts for which

the TNS is an SVD.

Two types of SVD Cuts

To compute the entanglement entropy, we use the same convention adopted in the

discussion of the 3D toric code (in Sec. 6.3) and the X-cube model (in Sec. 6.4): the

open virtual indices of the region A connect directly to the g tensors inside A while

the open virtual indices of the region A connect with T tensors inside A. We further

choose a region A such that the TNS is an SVD, and compute the entanglement

entropy. We find two types of entanglement cuts for which the Haah code TNS is

an exact SVD. For the general cubic region A, we need an extra step to perform the

SVD of the TNS. This derivation will be presented in Sec. 6.5.4. We now specify the

two types of regions that the Haah code TNS is an exact SVD.

1. Region A only consists of the spins connecting to a set of (l−1) T tensors which

are contracted along a certain direction. Figure 6.14 shows an example with

l− 1 = 3 contracted along the z direction. (Since in Sec. 6.3 and 6.4, we used l

as the number of vertices along each side of region A, there are l− 1 bonds (or

cubes) along each side.)

2. Region A contains all the spins connecting with T tensors which are contracted

in a “tripod-like” shape, where three legs extend along x, y, z directions. If there

are lx− 1 cubes in the x leg, ly − 1 cubes in the y leg, and lz − 1 cubes in the z

154

leg, then there are 1 + (lx − 2) + (ly − 2) + (lz − 2) = lx + ly + lz − 5 cubes (or

T tensors) region A. Figure 6.15 shows an example with lx = ly = lz = 3.

In the first case and for l = 2, 3, we use brute-force numerics to find that the reduced

density matrix is diagonal, which shows that the TNS is an exact SVD.

T

T

T

Figure 6.14: Region A contains all the spins connecting with l − 1 T tensors which

are contracted along z direction. The figure shows an example with l = 4.

T T

T

T

Figure 6.15: Region A contains all the spins connecting with T tensors which are

contracted in a “tripod-like” shape, where three legs extend along x, y, z directions.

There are three legs extending along x, y, z directions respectively. In general, three

legs can have different length, each with lx−1, ly−1, lz−1 cubes along three directions.

This figure shows an example where lx = ly = lz = 3.

In order to show that the above cuts correspond to an SVD, we follow the argu-

ments developed in Sec. 6.2.1. In Sec. 6.2.1, we proposed a SVD Condition. However,

we find that the region A of both types, shown in Fig. 6.14 and 6.15, do not sat-

155

isfy the SVD Condition: Two open virtual indices in region A connects with the

same T tensor, which violates the SVD Condition. For instance, the g1 and g2 in

Fig. 6.15 connects to the same T tensor in their upper-left cube which is in the region

A. Here, we propose a Generalized SVD Condition which suffices to prove that the

entanglement cut corresponding to Fig. 6.14 and 6.15 are SVD.

Generalized SVD condition: Let {t} be the set of open virtual indices. Given

a set of physical indices {s} inside region A, if {t} can be uniquely determined

by the {s} inside region A via the g tensor projection condition Eq. (6.118) and

T tensor constraints Eq. (6.125), then |{t}〉A is orthogonal. Since |{t}〉A is or-

thogonal because all the open virtual indices are connected with g tensors, the TNS

|TNS〉 =∑{t} |{t}〉A ⊗ |{t}〉A is SVD.

To prove the Generalized SVD Condition, we notice that if we have two different

sets of open virtual indices {t}A and {t′}A, the physical indices {s}A and {s′}A which

connect (via g tensors) to the T tensors on the boundary of region A cannot be the

same. Otherwise, if {s}A = {s′}A, since the physical indices {s}A and {s′}A in the

region A uniquely determine the open virtual indices {t}A and {t′}A, {t}A = {t′}A,

hence it is in contradiction with our assumption {t}A 6= {t′}A. Therefore, {t}A 6=

{t′}A implies {s}A 6= {s′}A, and hence A〈{t}|{t′}〉A = 0. This is in the same spirit of

the proof in Sec. 6.2.1. The proof of normalization of the wave function is independent

of {t} is also the same as in Sec. 6.2.1. Furthermore, A〈{t}|{t′}〉A = 0 for {t} 6= {t′}

is the straightforward because {t}A are connected with g tensors. In summary, if the

entanglement cut satisfies the Generalized SVD Condition, we have

1. A〈{t}|{t′}〉A ∝ δ{t},{t′} when |{t}〉A and |{t′}〉A are not null vectors;

2. A〈{t}|{t′}〉A ∝ δ{t},{t′} when |{t}〉A and |{t′}〉A are not null vectors.

This shows that the TNS is an SVD.

156

We explain the Generalized SVD Condition in the simplest example, i.e., l = 2 in

case 1. There is only one T tensor, and the region A contains 8 physical spins.

T

All other spins apart from the eight connecting with the T tensor belong to region

A. Because the virtual indices and physical indices are related by the g tensor which

is a projector, we use i1 to denote the values of both virtual indices and physical

indices connecting with left g tensor located at (x, y, z) = (0, 0, 1). Here, we use

the coordinate convention where the (x, y, z) = (0, 0, 0) is located at the left down

frontmost corner as in Fig. 6.13. Similarly we use i2, i3, i4, i5, i6, i7, i8 to label the

values of the virtual/physical indices on the remaining seven nodes connecting with

the same T tensor. Hence the set of open indices is effectively {i1, i2, i3, i4, i5, i6, i7, i8}

(after identified by the g tensors). We further consider how the physical indices from

the region A constrain the open indices. Consider the T tensor in the region A (which

we denote by T ′) which shares two spins i7, i8 with the region A (The T ′ tensor lives

157

in the lower right corner):

T

T'

(6.127)

Since six among the eight virtual indices of T ′ are contracted with g tensors inside

region A, the remaining two open virtual indices, i.e., i7 and i8 are subject to one

constraint from the T ′ tensor:

i7 + i8 = fixed, (6.128)

where “fixed” means that the sum is fixed by the physical indices inside the region

A. We can similarly consider the constraints coming from other T tensors in region

158

A. The whole set of constraints are listed as follows:

i7 + i8 = fixed

i1 + i2 = fixed

i5 = fixed

i6 = fixed

i6 + i7 = fixed

i2 + i3 = fixed

i5 = fixed

i8 = fixed

i1 = fixed

i4 = fixed

i3 = fixed

i7 = fixed. (6.129)

The “fixed” on the right hand side of the equations means that the virtual indices

or the sum of the virtual indices are fixed by the physical indices in the region A.

All variables and equations are defined module 2. The above equations uniquely

determine all the open virtual indices i1...i8. Therefore, such a choice of region A of

the entanglement cut satisfies the Generalized SVD Condition.

For the first type of the region A with general l, and the second type of region A

with general lx, ly, lz, we can similarly check that the TNS satisfies the Generalized

SVD Condition. Numerically, we checked that the Haah code TNS indeed satisfies

the Generalized SVD Condition for 2 ≤ l ≤ 9 for the first type, and 3 ≤ lx ≤ 8, 3 ≤

ly ≤ 8, 3 ≤ lz ≤ 8 for the second type. The numerical procedure for this check is to

159

list all the constraints for the indices in the region A and find how many solutions

exist for these constraints.

Entanglement entropy

We now compute the entanglement entropy for the exact SVD TNSs. We first consider

the case 1 with general l, such as in Fig. 6.14. All the spins connecting with l − 1

contracted T tensors along the z directions are in region A, and the remaining belong

to region A. The number of open virtual indices, after identified by the local g

tensors, is 8 + 7(l − 2) = 7l − 6. The number of constraints from the local T tensors

is simply the number of T tensors l− 1, because they are all independent. Hence the

number of independent open virtual indices is 7l − 6 − (l − 1) = 6l − 5. Therefore,

the entanglement entropy is

S(A)

log 2= 6l − 5. (6.130)

We further consider the case 2 — the region A of tripod shape. The legs in the

x, y, z direction contains lx− 1, ly− 1, lz− 1 T tensors respectively. We first count the

total number of open virtual indices. When lx = ly = lz = 3 as shown in Fig. 6.15,

there are 26 physical spins (or g tensors) in total. However, there is one g tensor (at

the left spin of (x, y, z) = (1, 1, 1)) whose four virtual indices are all contracted by the

T tensors within region A. Hence the number of open virtual indices, after identified

by the local g tensor, is 25. Moreover, we notice that adding one T tensor in one of the

three legs of region A brings 7 extra spins. Therefore the total number of open virtual

indices (after identified by the g tensor) is (26−1) + 7(lx−3) + 7(ly−3) + 7(lz−3) =

7lx+7ly+7lz−38. We further numerically count the number of constraints that these

open virtual indices satisfy. We find the number of constraints is the number of cubes

minus 1, i.e., (lx+ly+lz−5)−1 = lx+ly+lz−6. Therefore the number of independent

160

open virtual indices is (7lx + 7ly + 7lz − 38)− (lx + ly + lz − 6) = 6lx + 6ly + 6lz − 32.

The entanglement entropy is

S(A)

log 2= 6lx + 6ly + 6lz − 32. (6.131)

6.5.4 Entanglement Entropy for Cubic Cuts

In this section, we consider the case where the region A is a cube of size l × l × l,

where l is the number of vertices in each direction of the cube. The cut is chosen such

that all the open virtual indices straddling the region A are connected to g tensors in

the region A (i.e., all the physical spins near the boundary belong to the region A).

For example, for l = 2 as shown in (6.119), all 16 physical spins belong to the region

A. For l = 3 as shown in Fig. 6.13 (b), all 54 physical spins belong to the region A.

For the simplicity of notations, in this section, we denote the Hamiltonian terms as

Ac and Bc where the subscript refers to a cube c.

SVD for TNS

For the cubic region A, we find that the TNS for the Haah code is different from that

for the toric code and X-cube model: the TNS for the Haah code is not an exact

SVD. The TNS basis in the region A, |{t}〉A, are orthonormal, since the open virtual

indices are connected with g tensors. However, the TNS basis |{t}〉A in the region A

are not orthogonal. In other words, the basis |{t}〉A is over complete.

The subtlety that the TNS bipartition is not an exact SVD manifests as follows:

the singular vectors in the region A for the ground states of the Haah code have to be

the eigenvectors of all Ac and Bc operators that actually lie in the region A, and the

corresponding eigenvalues should all be 1. Notice that our TNS basis state |{t}〉A, if

not null, are the eigenvectors of all Ac operators inside the region A with eigenvalues

1, and are also the eigenvectors of Bc operators with eigenvalues 1 when Bc operators

161

T T

Cut

x

x

x

xx

T T

Cut

x

x

(a)

(b)

Figure 6.16: Transferring the Pauli X operators of the Bc operator from the regionA (a) to the region A (b).

are deep inside the region A, i.e., when they do not act on any spin at the boundary

of A. However, |{t}〉A are not the eigenvectors of Bc operators, when Bc operators

are inside the region A but also adjacent to the region A’s boundary. The reason

is that the Bc operators adjacent to the region A’s boundary, when acting on the

TNS basis |{t}〉A, will flip the physical spins on the boundary, and thus flip the open

virtual indices {t} due to the projector g tensors. Therefore, the basis |{t}〉A is no

longer the singular vectors for the Haah code. This is not an a priori problem, but

a result of the geometry of the Haah code, whose spins cannot be written on bonds

but have to be written on sites. A similar situation would occur if the 2D toric code

model would be re-written to have its spins on sites.

The method to find the correct SVD for the TNS is to use the |{t}〉A to construct

the eigenvectors of Bc operators by projection. We prove the following statement:

If |{t′}〉A = Bc|{t}〉A when Bc is inside the region A and also adjacent to the

region A’s boundary, then A〈{t′}|{t}〉A = 0 and |{t′}〉A = |{t}〉A.

The proof is as follows. The first part of the statement is a consequence of the

|{t}〉A basis state orthogonality. Indeed, Bc flips physical spins located at the region

162

A’s boundary. Thus the two sets {t} and {t′} are distinct. The second part of the

statement is more involved. Suppose for simplicity that we consider two nearest

neighbor T tensors for the region A and A in Fig. 6.16. The Bc operator acts on the

right cube Fig. 6.16 (a). The physical spins on the boundary of the region A which

are flipped by Bc are those covered by circled X in Fig. 6.16 (a). Then these Pauli X

operators can be transferred to the virtual indices due to projector g tensors, and the

virtual indices of the T tensor in the region A obtain two X operators as in Fig. 6.16

(b). Notice that the T tensor for the Haah code is invariant under this action (see

the 12th cube in Eq. (6.123)). This is also true for other T tensors in the region A

that are affected by Bc. The transfer of X operators from the region A to the region

A gives exactly the same equations in Eq. (6.123) when we solve for the T tensors.

Hence, the X operators transferred to the open virtual indices in the region A do not

change the state at all, i.e., |{t′}〉A = |{t}〉A. As a consequence, we can perform the

following factorization

|{t}〉A ⊗ |{t}〉A + |{t′}〉A ⊗ |{t′}〉A

=[(1 +Bc)|{t}〉A

]⊗ |{t}〉A.

(6.132)

The left part of the tensor product is an eigenstate of Bc with eigenvalue 1.

Therefore, in the TNS decomposition Eq. (6.25), we can group the basis state |{t}〉A

which are connected by this Bc operator. This factorization can be extended to

any product of Bc operators inside the region A and also adjacent to the region A’s

boundary. Notice that any such product has at least one X operator belonging to

only one Bc and so is different from the identity. When acting with all the possible

products of these Bc operator (including the identity) on a given |{t}〉A will generate

as many unique states as there are Bc’s. The TNS can be brought to the following

163

form

|TNS〉 =∑{t}′

[∏c

(1 +Bc

2

)|{t}〉A

]⊗ |{t}〉A, (6.133)

where the product over c only involves the Bc operators inside the region A and also

adjacent to the region A’s boundary and the sum over {t}′ is over the open virtual

index configurations that are not related by the action of these Bc operators.

Counting the number of TNS basis in region A: Notations

To compute the upper bound of the entanglement entropy, we need to find the number

of singular vectors in the region A that are also eigenstates of any Bc operators fully

lying in the region A. This number that we denote as basis(TNS(A)) is

basis(TNS(A)) = 2N−NB , (6.134)

where N is the number of independent open virtual indices and NB is the number

of Bc operators inside the region A and also adjacent to the region A’s boundary.

Every open virtual index connected to a g tensor located in A and at the boundary

of this region. Since each g tensor has a unique independent virtual index, we have

N = Ng −Nc where Ng is the number of g tensors in A and at the boundary of this

region and Nc is the number of constraints on the open indices coming from the T

tensors within the region A. We thus get

log2(basis(TNS(A))) = Ng −Nc −NB (6.135)

and the upper bound on the entanglement entropy reads

S(A) = (Ng −Nc −NB) log 2. (6.136)

164

Counting Ng and NB

We first count Ng. The number of g tensors can be computed by looking at Fig. 6.13

(b). We consider the region A with size lx × ly × lz (Notice that lx, ly, lz are the

number of vertices in each direction). In eight corners, there are 8× 2 = 16 vertices.

On the four hinges along x direction, there are 2 × 4 × (lx − 2) vertices, where 2

means there are two spins on each vertex, and 4 means four hinges. And similar for

2 × 4 × (ly − 2) and 2 × 4 × (lz − 2) in the y and z directions respectively. For the

xy-plane, there are 2× 2× (lx− 2)(ly− 2), where the first 2 comes from two spins per

vertex, and the second 2 comes from two xy-planes. Similarly 2× 2× (lx− 2)(lz − 2)

and 2 × 2 × (ly − 2)(lz − 2) for xz and yz plane respectively. Therefore, the total

number of g tensors is

Ng =16 + 8(lx − 2) + 8(ly − 2) + 8(lz − 2)

+ 4(lx − 2)(ly − 2) + 4(lx − 2)(lz − 2)

+ 4(ly − 2)(lz − 2)

=4lxly + 4lxlz + 4lylz − 8lx − 8ly − 8lz + 16.

(6.137)

We further count NB. As explained in Sec. 6.5.4, NB is the number of Bc operators

inside the region A and adjacent to the boundary of the region A. For a cubic region

A with size l × l × l (which is the case we consider below), the number of such Bc

operators are

NB = (l − 1)3 − (l − 3)3 = 6l2 − 24l + 26,∀l ≥ 3. (6.138)

For l = 2, we just have one Bc operator. Hence we have

NB = 6l2 − 24l + 26− δl,2,∀l ≥ 2. (6.139)

165

Counting Nc: Contribution from the T tensors

The open indices may be constrained by the T tensors fully inside the region A. In

the following, we will discuss the specific entanglement cuts where lx = ly = lz = l.

We rely on numerical calculations to evaluate Nc. We first consider the examples

l = 2 and l = 3 in detail, and then we describe our algorithm to search the number

of linearly independent constraints.

For l = 2, as shown in Eq. (6.119), no g tensor has all virtual indices contracted.

There is only one T tensor within region A. The element of the T tensor is

Ti1i2i3i4i5i6i7i8 (6.140)

where i1, i2, i3, i4, i5, i6, i7, i8 are all contracted virtual indices. Because they are con-

tracted with g tensors where at least one virtual index is open, all the contracted

virtual indices i1, i2, i3, i4, i5, i6, i7, i8 are equal to some open indices, and we denote

them as

i1 = t1, i2 = t2, i3 = t3, i4 = t4,

i5 = t5, i6 = t6, i7 = t7, i8 = t8.

(6.141)

The constraints on {i}’s are hence equivalent to the constraints on {t}’s, i.e.,

t1 + t2 + t3 + t4 + t5 + t6 + t7 + t8 = 0 mod 2. (6.142)

There is only one constraint from the T tensor. Hence Nc = 1 for l = 2.

For l = 3, as shown in the Fig. 6.13 (b), we have eight constraints from eight T

tensors which involve the open indices via the g tensors. The eight equations are

8∑n=1

i(x,y,z)n = 0 mod 2, x, y, z ∈ {0, 1} (6.143)

166

where the up-index (x, y, z) represents the position of the T tensor, and n counts the

eight indices of each cube in the 2 × 2 × 2 cut. All the i’s are contracted virtual

indices. However, except the virtual indices that are connected with the central two

g tensors (which are defined on the two spins at the vertex (x, y, z) = (1, 1, 1)), all

other indices (which are defined on two spins at vertices (x, y, z), x, y, z ∈ {0, 1, 2}

except (x, y, z) = (1, 1, 1)) are equal to some open indices via g tensors. Specifically,

the virtual indices that are connected with the two center g tensors are

i0004 = i100

3 = i0101 = i001

7 mod 2

i0005 = i110

2 = i1018 = i011

6 mod 2.

(6.144)

Since we only count the number of constraints for the open indices, we need to Gauss-

eliminate all these eight virtual indices i0004 , i100

3 , i0101 , i001

7 , i0005 , i110

2 , i1018 , i011

6 from the

above 8 equations. Therefore, we obtain 8− 2 = 6 independent equations in terms of

open indices only. Hence there are 6 constraints for the open indices.

For the general l, we apply the same principle. We first enumerate all possible

constraints from the T tensors, and then we Gauss-eliminate all the virtual indices

that are contracted within the region A. Hence we obtain a set of equations purely

in terms of the open indices. The number of constraints is the rank of this set of

equations. We list the number of linear independent constraints for the open indices

as follows:

l(≥ 3) 3 4 5 6 7 8 9 10 11 12 13 14 15 16

Nc 6 12 18 24 30 36 42 48 54 60 66 72 78 84(6.145)

Hence, for l ≥ 3, there are

6l − 12 (6.146)

167

linearly independent constraints for the open indices. Taking into account the fact

that when l = 2 the number of constraints is 1, we infer that the number of constraints

for a generic l is:

6l − 12 + δl,2. (6.147)

Entanglement entropy

We are ready to collect all the data we have obtained and compute the entanglement

entropy for the cubic cut. For the entanglement cut of size l× l× l, the total number

of g tensors is

Ng = 12l2 − 24l + 16. (6.148)

The number of of T tensor constraints is

Nc = 6l − 12 + δl,2,∀l ≥ 2. (6.149)

The number of Bc operators is

NB = 6l2 − 24l + 26− δl,2,∀l ≥ 2. (6.150)

Therefore the upper bound of the entanglement entropy reads

S

log 2=Ng −Nc −NB

=6l2 − 6l + 2,∀l ≥ 2.

(6.151)

The entanglement entropies also have negative linear corrections.

168

If the region A is much larger than the region A, we conjecture that the region A

will not impose any additional constraint. In that case, the upper bound would be

saturated.

6.6 Conclusion and Discussion

In this chapter, we present our TNS construction for three stabilizer models in 3D.

The ground states of these stabilizer codes are the eigenstates of all local Hamiltonian

terms with +1 eigenvalues. The constructions of these TNSs share the same general

idea and work in other dimensions as well:

1. We introduce a projector g tensor for each physical spin which identifies the

physical index with the virtual indices.

2. The physical operators acting on the TNS can be transferred to the virtual

indices using Eq. (6.36).

3. The local T tensors contracted with the projector g tensors are specified by the

local Hamiltonian terms.

After we obtain the TNS for the ground state, we can prove that the TNS is

an exact SVD for the ground state with some specific entanglement cuts. The en-

tanglement spectra are completely flat for the models studied in this chapter. The

entanglement entropies can be computed by counting the number of singular vectors.

For the 3D toric code model, the entanglement entropies have a constant correction

to the area law, − log(2). For the X-cube model and the Haah code, the entanglement

entropies have linear corrections to the area law as shown in Sec. 6.4.4 and 6.5.4. The

analytical calculation of the entanglement entropies is rooted in the Concatenation

lemma, since the Concatenation lemma is introduced to count the number of

169

singular vectors. The Concatenation lemmas are rooted the symmetry properties

of the local tensors. For instance, Eq. (6.40) and (6.43) for the 3D toric code model.

The transfer matrices can also be constructed. For the 3D toric code and the

X-cube models, we prove that the transfer matrix is a projector whose dimension

is counted by the Concatenation lemma as well. For the 3D toric code model,

the transfer matrix is of dimension 2. For the X-cube model, the transfer matrix

is of dimension 2Lx+Ly−1 where Lx and Ly are the sizes of the torus in the x and y

directions respectively. The GSD on the torus is generally larger than the degeneracy

of the transfer matrix.

Since both the entanglement entropies and the transfer matrix degeneracies are

rooted in the Concatenation lemma (or more fundamentally the symmetry prop-

erties of the local T tensors), we believe that these two phenomena are related. More-

over, we conjecture that the negative linear correction to the area law is a signature

of fracton models. This is similar to the negative constant correction (i.e., the topo-

logical entanglement entropy[22, 23]) in 2D.

In this chapter, the TNSs are all the ground states of some exactly solvable local

models. If we move away from these fine-tuned points without going through phase

transitions, we expect the transfer matrix degeneracies to be still robust, since these

degeneracies give rise to the GSD. In Ref. [179], this statement has been numerically

verified in the 2D toric code model and its phase transitions to the trivial phases.

If we move away from the fine-tuned points, we also expect that the linear term

of the entanglement entropies for the fracton models does not vanish, although the

specific coefficients of the linear terms might change. An important result is about

the topological entanglement entropy. The topological entanglement entropy for the

fracton models was first introduced in Ref. [158], and is defined as the linear combi-

nations of the entanglement entropies of different regions, in order to exactly cancel

the area law. See Ref. [158] for the definition details. Importantly, the topological

170

entanglement entropies of fracton models are linear with respect to the sizes of the

entanglement cuts. Furthermore, Ref. [158] argues using perturbation theories that

the topological entanglement entropies, of the same three models as in our chapter

are robust to adiabatic perturbations. Hence, Ref. [158] indicates that there should

be a linear correction to the area law which does not vanish, even when moving away

from the fine-tuned wave functions.

However, we also have to admit that the rigorous statements, about the entan-

glement spectra, entropies and the transfer matrix degeneracies of a generic fracton

model ground state, need to be verified by the numerical studies for the 3D fracton

models in the future.

171

Appendix A

Appendix for Boson Condensation

A.1 Quantum dimensions of A and T

A.1.1 Proof of da =∑

r∈T nradr

From the Eq. (3.5) we obtain, by multiplying both sides by the quantum dimension

dt of the particle t in the T theory and summing over t

∑r,s,t∈T

nransbN

trsdt =

∑c∈A,t∈T

N cabn

tcdt =

∑r,s∈A

nransbdrds, (A.1)

where we are only considering T theories which are also fusion categories (not braided

ones) and hence satisfy the equivalent Eq. (1.31) for the T theories

∑t∈T

N trsdt = drds. (A.2)

We then have the trivial re-writing

∑c∈A

N cab

(∑t∈T

ntcdt

)=

(∑r∈T

nradr

)(∑s∈T

nsbds

), (A.3)

172

which means that(∑

t∈T ntcdt)

is an eigenvalue of the matrix (Na)bc with eigenvalues

and eigenvector(∑

r∈T nradr)

and(∑

s∈T nsbds), respectively. Since the eigenvector

has positive entries, by the Perron-Frobenius theorem, the eigenvalue is the largest

eigenvalue of the Na matrix, and hence it is indeed da

da =∑r∈T

nradr. (A.4)

A.1.2 Proof of dr = 1q

∑a∈A n

rada

We start with Eq. (3.5), multiply by da and sum over a ∈ A. Using Eq. (1.31), it

follows

∑r,s∈T

∑a∈A

nradaNtrsn

sb =

∑a,c∈A

ntcNcabda

= db∑c∈A

ntcdc.

(A.5)

For the simplicity of notations, let αt ≡∑

c∈A dcntc, which satisfies the eigenvalue

equation

∑r∈T

(∑s∈T

nsbNs

)tr

αr = dbαt. (A.6)

Notice the unorthodox use of the matrix (Ns)tr = N tsr, unlike in the line following

Eq. (1.31). We define the matrix this way in order not to use the equation nta = nta.

This matrix has the vector of quantum dimensions (d1, . . . , dN)T, where N are the

number of particles in the fusion category T , as an eigenvector ∀s in (Ns)tr. Since

we are using the Ns matrix in an unorthodox fashion (it is the transpose of the usual

Ns matrix), we prove the statement

∑t

N rstdr = dsdt =

∑t

N tsrdr = dsdt. (A.7)

173

It follows from the above that∑

r(Ns)trdr = dsdt. Hence (d1, . . . , dN)T is a common

eigenvector of all the Ns, even as defined in the unusual way above.

We now sum Eq. (A.6) over b to get

∑r∈T

(∑b∈A

∑s∈T

nsbNs

)tr

αr =

(∑b∈A

db

)αt. (A.8)

The matrix (∑

b∈A∑

s∈T nsbNs)tr is a completely positive matrix with integer strictly

positive coefficients: for any t, r, there exists s such that N tsr > 0 and for every s

there exists an nsb > 0. As such, it satisfies a stronger version of the Perron-Frobenius

theorem which says that there is a unique eigenvector with all elements positive, and

all other eigenvectors have at least one negative element. As such, since αt is all

positive, we identify it as the unique largest eigenvector. But since the Ns have a

common eigenvector, the quantum dimensions of the condensed theory, we then can

identify this eigenvector with

αt =∑c∈A

dcntc = qdt, (A.9)

where q is a proportionality constant. We now find two expressions for it. First,

multiplying Eq. (A.9) by dt and summing over t gives

q∑t∈T

d2t =

∑c∈A

dc∑t∈T

ntcdt =∑c∈A

d2c , (A.10)

where the last equality follows from Eq. (A.4). This implies

q = D2A/D

2T . (A.11)

174

Furthermore, multiplying Eq. (3.5) by dadb for t = ϕ and summing over a, b reads

∑c∈A

N cabn

ϕc =

∑t∈T

ntantb, (A.12)

which implies

∑a,b,c∈A

dadbNcabn

ϕc =

∑t∈T

∑a∈A

danta

∑b∈A

dbntb

= q2D2T .

(A.13)

On the other hand,

q2D2T =

∑a,b,c∈A

dadbNcabn

ϕc

=∑b,c∈A

d2bdcn

ϕc

= D2A

∑c∈A

dcnϕc ,

(A.14)

hence

q =∑c∈A

dcnϕc . (A.15)

A.2 Chiral algebra

In this section, we review the connection between the above formalism and CFT. As

pointed out by Bais and Slingerland [42], the mathematics of boson condensation has a

parallel in conformal field theories. First, for at least some MTCsA, the particle labels

are in one-to-one correspondence with the conformal families in some (not necessarily

unique) CFT. (The MTC-conformal block correspondence generalizes Witten’s work

[80] on the relationship between Chern-Simons theory and chiral Wess-Zumino-Witten

175

models.) Second, when this correspondence holds, the process of condensation in the

TQFT is closely related to extending the chiral algebra in the CFT [115].

Let us consider a CFT with a chiral algebra A which contains the stress-tensor

T (z) and all locally commuting holomorphic operators in the theory such as currents

Ja(z) associated to Lie groups, etc. The mode expansions of these operators give

rise to infinite dimensional algebras, like Virasoro, Kac-Moody or W -algebras. The

irreducible representation spaces of the chiral algebra A, denoted by Ha, are labelled

by the primary fields a, whose number is finite in a RCFT. The primary fields are

in one-to-one correspondence with the anyons of a TQFT. The TQFT is nothing but

the CFT reduced to its basic topological data like braiding and fusion matrices, etc.

(However, due to this reduction, several distinct CFTs may correspond to the same

TQFT.)

For each representation space Ha there is a character

χa(τ) = TrHae2πiτ(L0−c/24), (A.16)

given by the partition function of the states in Ha propagating along a torus with

modular parameter τ (with Im τ > 0). The constant c is the central charge of the

CFT and L0 is the zero element of the Virasoro algebra. The modular transformations

act on the characters as

χa(τ + 1) = θae− iπc

12 χa(τ), (A.17)

χa

(−1

τ

)=

∑b

Sabχb(τ),

where θa = e2πiha is the topological spin, ha the conformal weight of the primary

field a. The full CFT also contains an anti-chiral algebra, A, which for simplicity we

assume to be isomorphic to A. Correspondingly, the complete Hilbert space is the

176

tensor product H = ⊕aHa ⊗ Ha and the total partition function is

Zdiag(τ, τ) =∑a

χa(τ)χa(τ), (A.18)

which is modular invariant thanks to the S, T unitarity: SS† = TT † = 1. The

pairing between the left and right states of a non-chiral CFT can be more general

than (A.18)

Z(τ, τ) =∑a,b

χa(τ)Mab χb(τ), (A.19)

where M is called the mass matrix which must satisfy [S,M ] = [T,M ] = 0 to guaran-

tee the modular invariance of the partition function (A.19). A fundamental problem

in RCFT is to classify all possible modular invariant partition functions, that is,

mass matrices M . This program has been achieved for theories with simple currents

[100, 101, 102], but it is far from being solved in general.

There are three types of mass matrices: i) Those associated to automorphisms of

the fusion rule algebra, ii) those corresponding to a chiral extension of A, and iii) a

combination of i) and ii). This result is related to the naturality theorem due to Moore

and Seiberg: In a CFT when the left and right chiral algebras are maximally extended

the field content matrix defines an automorphism ω of the fusion rule algebra, i.e.:

Ma,b = δa,ω(b) [107]. A chiral algebra A ⊗ A is called maximally extended when it

includes all the holomorphic and antiholomorphic fields in H (i.e., those with integer

conformal weights). [106]

The mass matrices and the associated naturality theorem have a precise correspon-

dence within the boson condensation encountered in the main text. Let us explain it

in more detail.

An extension of the chiral algebra A can arise if there exists a subset {γi} of pri-

mary fields with integer conformal weights that are mutually local. One can therefore

177

add these holomorphic fields to those already included in A to obtain an extended

chiral algebra U . It is then clear that the representation spaces of the new algebra

U should be a combination of those of the original algebra A. In particular, the

(irreducible) conformal family vector space Hϕ corresponding to the new identity

representation ϕ will be the direct sum of the old identity conformal family H1 plus

the conformal families corresponding to the old primaries γi, that is Hϕ = H1⊕iHγi .

The fields γi correspond to the bosons that condense in the TQFT. The space Hϕ is

the CFT version of the vacuum after condensation.

The irreducible representation spaces of the extended chiral algebra U , denoted by

Hu, break down into the direct sum of irreducible representations Ha of the smaller

algebra A. Such decompositions are called branching rules and are noted as

Hu → ⊕a∈AnuaHa. (A.20)

The branching coefficient nua gives the multiplicity of the irreducible representation

a of A in the decomposition of the irreducible representation u of U . The fields

appearing in the decomposition (A.20) have to be mutually local with respect to the

fields in the chiral algebra U . From Eqs. (A.20) and. (A.16) follows the expression for

the character of the representation u in terms of the characters of the representations

a [recall Eq.(C1)] [115]

χu(τ) =∑a∈A

nua χa(τ), u ∈ U . (A.21)

The primary field u corresponds to a deconfined anyon in the TQFT. The TQFT

Eq. (3.4) means in CFT that the primary fields that built up a representation of

the extended algebra must have the same conformal weights modulo integers. On

the other hand, if a field a is such that the orbit γi × a, ∀i contains fields with

different conformal weights, then they disappear from the representation theory of

178

U . These fields are associated to the confined anyons defined in Eq. (3.3). Given the

characters (A.21) of the extended chiral algebra U , one can construct the diagonal

partition function

Z(τ, τ) =∑u∈U

|χu(τ)|2, (A.22)

which when written in terms of the characters (A.16) of A reads like Eq. (A.19) with

Mab =∑u∈U

nua nub . (A.23)

This equation shows that an extension of the chiral algebra gives rise to an off-diagonal

partition function and in turn to a boson condensation in the TQFT.

The original and chirally extended CFTs are both assumed to be modular theories,

with their characters transforming under modular transformation S and T of the torus

parameter τ as

χs

(−1

τ

)=∑t

Sstχt(τ) =∑t,a

Sstntaχa(τ)

=∑b

nsbχb

(−1

τ

)=∑a,b

nsbSbaχa(τ),

(A.24)

i.e.,

nS = Sn. (A.25a)

Similarly

nT = Tn, (A.25b)

where S and T are modular matrices for the U algebra. Equation (A.25a) and

Eq. (A.25b) also appear as matching conditions in the study of gapped domain walls

between two topological phases [64].

One can easily deduce that [M,S] = [M,T ] = 0. Moreover, through Eq. (A.25a)

and Eq. (A.25b), we can show that

179

1. c = c (mod 24),

2. θs = θa, if nsa 6= 0,

3. nC = Cn,

4. q ≡∑

a nϕada = DA/DU ,

5. dt = 1q

∑a∈A n

tada,

where C and C are the charge conjugation matrices for the U and A theories re-

spectively, and DU and DA are total quantum dimension of the U and A theory,

respectively.

So far we have discussed the mass matrices that correspond to extensions of the

chiral algebra. The other possibility is that the mass matrix is a permutation P of the

irreducible representations of A that corresponds to an automorphism of the fusion

rules [115]. This case does not describe boson condensation. The third possibility is

that the mass matrix describes an off diagonal partition function of the chiral algebra

U , namely M = nPnT, with P a permutation automorphism of the fusion rules of U .

These possibilities were mentioned before in connection with the Moore and Seiberg

naturally theorem [107].

The conclusions we obtain above, including Eq. (A.25a) and Eq. (A.25b), can

be viewed as necessary conditions for boson condensation. So, a solution of the

above consistency equations does not guarantee the existence of a boson condensation

A → U . It could still happen, for example, that the fusion coefficients derived from

such a solution via the Verlinde formula are not integers (see Appendix A.3.2 for an

example). Then, the solution has to be discarded. However, the absence of a solution

does imply that there is no boson condensation A → U .

180

A.3 Condensations in SU(2) CFTs

To illustrate some properties of the condensation transition we consider the family

of CFTs that correspond to SU(2) at level k. These theories have (k + 1) primary

fields in corresponding conformal blocks labelled by integers a = 0, . . . , k that are

denoted as φa, and the corresponding conformal characters are denoted as χa. (In

the corresponding TQFT, the anyon with a = 0 is the vacuum.) The matrix elements

of the modular S and T matrices are given by

Sab =

√2

2 + ksin

π(a+ 1)(b+ 1)

k + 2, (A.26)

and

Tab = e2πia(a+2)4(k+2) δab, c =

3k

k + 2. (A.27)

All the modular invariant partition functions of this CFT were obtained by Cappelli,

Itzykson and Zuber who found a surprising correspondence with the ADE classifica-

tion of Lie groups [114]. The complete list is

ZAk+1=

k∑n=0,n∈Z

|χn|2, (A.28a)

ZD2`+2=

2`−2∑n=0,n∈2Z

|χn + χ4`−n|2 + 2|χ2`|2,

ZD2`+1=

4`−2∑n=0,n∈2Z

|χn|2 + |χ2`−1|2

+2`−3∑

n=1,n∈2Z+1

(χnχ4`−2−n + χ4`−2−nχn),

181

ZE6 = |χ0 + χ6|2 + |χ3 + χ7|2 + |χ4 + χ10|2,

ZE7 = |χ0 + χ16|2 + |χ4 + χ12|2 + |χ6 + χ10|2

+|χ8|2 + χ8(χ2 + χ14) + (χ2 + χ14)χ8,

ZE8 = |χ0 + χ10 + χ18 + χ28|2

+|χ6 + χ12 + χ16 + χ22|2 ,

where k = 4` in ZD2`+2, k = 4` − 2 and in ZD2`+1

, while k = 10 in ZE6 , k = 16 in

ZE7 , and k = 28 in ZE8 . Here, χa are the characters of the irreducible representation

spaces of the chiral algebra of SU(2)k.

The origin of these off-diagonal partition functions is the following:

• D2`+2: J = φ4`, is a bosonic simple current with integer conformal weight

hJ = `. For ` = 1, φ4 is a current that yields a chiral extension corresponding

to the conformal embedding SU(2)4 ⊂ SU(3)1. (Notice that the central charge

of the two CFTs is the same cSU(2)4 = cSU(3)1 .)

• D2`+1: the simple current J = φ4` has half-odd conformal weights, hJ = `−1/2,

so it does not yield an extension of the chiral algebra, i.e., it does not correspond

to condensation. The partition function can be written as ZD2`+1=∑

a χa χω(a),

where ω is the unique automorphism of the fusion rules, namely ω(a) = a for a

even and ω(a) = k − a for a odd.

• E6: chiral extension with the field φ6 with h6 = 1. This is not a sim-

ple current. The chiral extension corresponds to the conformal embedding

SU(2)10 ⊂ SO(5)1, both CFT’s have the same central charge, namely c = 5/2.

The SO(5)1 algebra can be constructed with 5 Majorana fermions (i.e. Ising

models). In the SU(2)10 theory one has h4 = 1/2, h10 = 5/2, h3 = 5/16,

182

h7 = 21/16, so that h10 − h4 = 2 and h7 − h3 = 1. The field φ3 can be built

from the product of 5 spin fields of the Ising model which have hσ = 1/16.

• E7: explained by an exceptional automorphism of the D10 chiral algebra

[107, 115] [see Eq.(A.34)].

• E8: chiral extension with three operators with h10 = 1, h18 = 3, and h28 = 7.

The remaining fields in ZE8 have weights: h6 = 2/5, h12 = 7/5, h16 = 12/5,

h22 = 22/5. The central charge is c = 14/5, which coincides with that of G2 at

level k = 1 [115].

The results explained above can be summarized in the following table:

Type k Z Comments

Ak+1 k - -

D2`+2 4` EXT SU(2)4 ⊂ SU(3)1

D2`+1 4`− 2 AUT -

E6 10 EXT SU(2)10 ⊂ SO(5)1

E7 16 AUT -

E8 28 EXT SU(2)28 ⊂ (G2)1

(A.29)

where E6, E7, E8 and G2 are the exceptional Lie groups, while EXT and AUT stand

for an extension of the chiral algebra and an automorphism of the theory, respectively.

Note that some theories, e.g., k = 16 have a D as well as a E invariant, as case that

we will now discuss in detail.

183

A.3.1 SU(2)16

The SU(2)16 CFT is special in that it has two different off-diagonal partition functions,

given by [recall Eq. (A.28a)]

ZD10 = |χ0 + χ16|2 + |χ2 + χ14|2 + |χ4 + χ12|2

+ |χ6 + χ10|2 + 2 |χ8|2(A.30)

and

ZE7 = |χ0 + χ16|2 + (χ2 + χ14)χ8 + χ8(χ2 + χ14)

+ |χ4 + χ12|2 + |χ6 + χ10|2 + |χ8|2 .(A.31)

Both of these theories correspond to a condensation of the boson a = 16. There are

exactly two distinct solutions to the equation [M,S] = [M,T ] = 0, given by

M (1) = nnT, M (2) = nPnT, (A.32)

where

nT =

1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1

0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0

0 0 0 0 1 0 0 0 0 0 0 0 1 0 0 0 0

0 0 0 0 0 0 1 0 0 0 1 0 0 0 0 0 0

0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0

(A.33)

184

and

P =

1 0 0 0 0 0

0 0 0 0 0 1

0 0 1 0 0 0

0 0 0 1 0 0

0 0 0 0 1 0

0 1 0 0 0 0

(A.34)

is an automorphism of the theory U . These two solutions for M are in one-to-one cor-

respondence with the two off-diagonal partition functions above. Here, M (1) encodes

the condensation transition itself, while the existence of the additional matrix M (2) is

related to the “naturality theorem” discussed in the main text and in Appendix A.2.

Interestingly, the equation Sn = nS, that yields the S matrix of the theory after

condensation, has two distinct solutions S and S ′, where

S =2

3

sin(π18

)12

cos(

2π9

)cos(π9

)12

12

12

1 12

−12

−12−1

2

cos(

2π9

)12− cos

(π9

)− sin

(π18

)12

12

cos(π9

)−1

2− sin

(π18

)cos(

2π9

)−1

2−1

2

12

−12

12

−12

−12

1

12

−12

12

−12

1 −12

(A.35)

185

and S ′ is obtained by exchanging the last two rows of S, a so-called Galois symme-

try [202]. Both matrices S and S ′ yield the same fusion rules Nt, e.g.,

N3 =

0 0 0 1 0 0

0 0 1 1 1 1

0 1 1 2 1 1

1 1 2 2 1 1

0 1 1 1 0 1

0 1 1 1 1 0

. (A.36)

It is worth noting that this condensation of a theory without multiplicities (all N cab in

SU(2)16 are 0 or 1) yields a theory with multiplicities: some of the N str in Eq. (A.36)

are larger than 1.

A.3.2 SU(2)28

The SU(2)28 CFT is special in that it also has two different off-diagonal partition

functions, given by [recall Eq. (A.28a)]

ZD16 = |χ0 + χ28|2 + |χ2 + χ26|2 + |χ4 + χ24|2

+ |χ6 + χ22|2 + |χ8 + χ20|2 + |χ10 + χ18|2

+ |χ12 + χ16|2 + 2 |χ14|2

(A.37)

and

ZE8 = |χ0 + χ10 + χ18 + χ28|2

+ |χ6 + χ12 + χ16 + χ22|2 .(A.38)

As is clear from ZE8 , the particles 10, 18, and 28 are bosons, besides the vacuum

0. The two partition functions correspond to two distinct condensations possible in

186

SU(2)28. These are the only condensations possible. Their corresponding n matrices

can be read off directly from these partition functions.

The partition function ZE8 stands for a condensation of all bosons with nϕa = 1

each. The resulting theory is the Fibonacci TQFT with particles 6, 12, 16, and 22

each restricting to the τ particle.

The partition function ZD16 corresponds to the condensation of the top-level boson

28 only, which results in a 9-particle non-Abelian TQFT with some multiplicities

larger than 1. For example, the restriction of 10 and 18, which we call 5, obeys the

fusion rule [c.f. Eq. (A.36)]

5× 5 = 0 + 1 + 2 + 3 + 2 · 4 + 2 · 5 + 2 · 6 + 7 + 8. (A.39)

Condensing four layers of Ising TQFT

Here we give some detail on one particular condensation in a theory composed of a

tensor product of 4 layers of the Ising TQFT. We show that, generically, in our algo-

rithm in the main text Sec. 3.6, we do need to check that the resulting S matrix gives

integer fusion matrices using Verlinde’s formula. This is not a complete discussion of

all possible condensations in the 4 layer Ising theory.

We focus on the condensate containing 1111, 11ψψ, 1ψ1ψ, and 1ψψ1, but not

ψψψψ, ψψ11, ψ1ψ1, and ψ11ψ. (Including these other bosons in the condensate

would yield the ν = 4 theory from Kitaev’s 16-fold way.) The corresponding M matrix

has only one nonzero entry on the column (and row) that corresponds to the σσσσ

particle, namely Mσσσσ,σσσσ = 4. Since the quantum dimension dσσσσ = 4, this allows

for two distinct solutions n: one in which σσσσ particle restricts to twice some particle

a and one in which it splits into 4 Abelian particles a1, a2, a3, a4. In both cases, we

can find solutions to Eq. (3.64) that are unitary and satisfy S2 = Θ(ST )3 = C.

187

However, for the theory in which the σσσσ particle splits into 4 particles, the

fusion coefficients Nt obtained from Eq. (3.65) are not all nonnegative integers (some

of them are ±1/2). This establishes by example that we have to impose that Nt is

nonnegative integer valued, in addition to the other conditions in Sec. 3.6. It also

shows that in the example at hand, despite the ambiguity in the possible solutions

for n, the particle content (up to possible automorphisms) of the final theory is fixed

by the choice of condensate. Whether this statement is true in general is presently

not known to us.

The allowed solution, in which σσσσ particle restricts to twice the same particle,

is the one we naively expect, upon inspection of the anyons in the condensates by the

following argument: all condensed anyons have a vacuum particle 1 in the first layer.

Hence the Ising theory in the first layer will be preserved under condensation and the

result is a direct product of the ν = 1 Ising theory and the ν = 3 Ising theory from

Kitaev’s 16-fold way. The particle that σσσσ twice restricts to is the direct product

of a ν = 1 Ising σ and ν = 3 Ising σ, which we have already proved in Sec. 3.7.1 has

nσ′

(σσσ) = 2.

188

Appendix B

Appendix for Uncondensable

Bosons

B.1 No-go theorem with Abelian sector

We have seen from the examples discussed in the main text, that the no-go theorem

can often be used to not only show that individual bosons in a TQFT cannot con-

dense, but that an entire TQFT is not condensable. Here, we extend this discussion

to examples of TQFTs that have noncondensable sub-structures. This problem is

motived by physical examples: in the fractional quantum Hall effect, for instance,

one frequently discusses phases that are described by a direct (or semi-direct) prod-

uct of an Abelian and a non-Abelian TQFT. A simple example is the Z3 Read-Rezayi

state of bosons, which is described by the TQFT AFib × Z2. While such a theory

admits condensations, already in the Z⊗N2 sector, when enough layers are considered,

one has the intuition that the noncondensability of Fibonacci should still constrain

the possible condensations.

Lemma 1. Consider a TQFT

A×X , (B.1)

189

where X is an Abelian TQFT (i.e., all its anyons have quantum dimension 1). Fur-

ther, for all particles b ∈ A (not only for the bosons), except for the vacuum, let there

exist a set Ib = {a1, . . . , am} of zero modes of b, containing anyons from A, such that

the quantum dimensions satisfy

db ≤m∑i=1

dai . (B.2)

Then, any possible condensation transition will lead to a theory of the form

A× Y , (B.3)

where the Abelian TQFT Y can be obtained from X through a condensation.

Proof. This Lemma follows almost directly from the no-go theorem. Let us denote a

particle from A×X by the pair (b, x) where b ∈ A and x ∈ X . If (b, x) is boson, we

can show that it has to be an uncondensable one, except if b = 1. The set

I(b,x) = {(a1, x), · · · , (am, x)} , (B.4)

(where a1, · · · , am form a set Ib of zero modes of b whose existence is guaranteed by

assumption) satisfies all the conditions 1–3 form the definition of a set of (b, x) zero

modes. Since x is an Abelian particle, dx = 1 and Eq. (B.2) directly implies that the

sum of the quantum dimensions of the particles in I(b,x) satisfies the inequality (4) from

the main text. Hence, (b, x) cannot condense. In turn, this implies any condensable

boson in A × X is of the form (1, x). A condensate of this form is transparent to

the anyons in A and will thus leave this sub-TQFT unaffected. It will only induce a

condensation X → Y , so that the final theory is of the from (B.3).

We return to the example of AFib × Z2. Consider N layers of this theory, i.e.,

A⊗NFib × Z⊗N2 . This multi-layer TQFT satisfies all assumptions of Lemma 1: for each

190

anyon b ∈ A⊗NFib , a choice for the set Ib is given by Ib = {1, b}. This is so because

all possible bosons appearing in the fusion product of b× b are uncondensable by the

no-go theorem and the sum of the quantum dimensions of Ib, given by 1 +db is larger

than db. We conclude that the A⊗NFib structure is preserved under any condensation

transition in such a theory.

B.2 Proof for Example (iii), Multiple layers of

SO(3)k

In this section, we show that no condensation is possible in the TQFT SO(3)⊗Nk

comprised of N layers of SO(3)k for any odd k and any integer N . The proof goes

by induction. We denote the particles in SO(3)⊗Nk with a shorthand notation. An

anyon that has the identity particle from SO(3)k in all layers, except for the k0 layers

i1, i2, · · · , ik0 , is denoted by {ji1ji2 · · · jik0}. Here 1 ≤ jil ≤ (k − 1)/2 can stand for

any anyon from SO(3)k (except the identity 0), for all l = 1, · · · , k0.

Induction base

First, consider particles {ji} with just one nontrivial anyon in some layer i. This will

serve as the induction base. By the no-go theorem and our proof in Example (ii),

we know that no bosons of form {ji} can condense. [Use the particles with only one

nontrivial in that same layer i to build the set Iji as elaborated for Example (ii).] As

a corollary, the anyons {ji} do not split: when fused with themselves no condensable

boson appears in the fusion product, which prevents splitting by Eq. (2) from the

main text for t = ϕ.

Induction step

We assume that for any 1 ≤ l ≤ k0 all {ji1ji2 · · · jil}

191

1. do not condense and

2. do not split.

We now show the induction step, namely that all particles with nontrivial anyons in

(k0 + 1) layers {ji1ji2 · · · jik0+1} neither condense nor split.

We begin by showing that {ji1ji2 · · · jik0+1} cannot condense. The particles

{ji1ji2 · · · jik0+1} can be obtained by fusing a {ji1ji2 · · · jik0} with a {jik0+1

}, where

ik0+1 /∈ {i1, · · · , ik0}. In this case, Eq. (2) from the main text reads for t = ϕ

Nϕ{ji1 ···jik0 }

↓,{jik0+1}↓ = nϕ{ji1ji2 ···jik0 jik0+1

}. (B.5)

Now, because of the uniqueness of the antiparticle, Nϕ{ji1 ···jik0 }

↓,{jik0+1}↓ can be either

0 or 1. If it was 1, {ji1ji2 · · · jik0}↓ would be the antiparticle of {jik0+1

}↓. Because

all particles are their own antiparticles, this would imply {ji1ji2 · · · jik0}↓ = {jik0+1

}↓.

However, this is not possible for k0 > 1, because the associativity of fusion would

then also imply that {ji1}↓ is the antiparticle (and coinciding with) {ji2 · · · jik0 jik0+1}↓,

i.e., {ji1}↓ = {ji2 · · · jik0 jik0+1}↓. Remembering that {ji1ji2 · · · jik0}, {ji2ji2 · · · jik0+1}

do not split, and equating the quantum dimensions of the particles for these two

identifications we have

dji1dji2 · · · djik0 = djik0+1,

dji1 = dji2 · · · djik0 djik0+1.

(B.6)

For k0 > 1, this contradicts the fact that all nontrivial particles in this theory have

quantum dimensions d > 1. This rules out the possibility Nϕ{ji1 ···jik0 }

↓,{jik0+1}↓ = 1 and

shows that {ji1ji2 · · · jik0+1} does not condense for k0 > 1.

The case k0 = 1 needs to be considered separately, as both lines in Eq. (B.6)

are identical in this case, and therefore do not lead to a contradiction. Assume that

Nϕ{ji1}↓,{ji2}↓

= 1. In the case ji1 6= ji2 , we can rely on the following argument to

192

disprove this assumption: As all anyons in SO(3)k with k odd have distinct quantum

dimension, it follows that the two anyons {ji1} and {ji2} restrict to distinct particles

and in particular ϕ /∈ j↓i1 × j↓i2

– with Eq. (2) from the main text this implies that

{ji1ji2} neither splits nor condenses. In the case ji1 = ji2 ≡ j, define j ≡ {ji1ji2}.

We want to show that j does not condense. As there are no fermions in SO(3)k with

k odd, j can only be a boson if θj = 1, i.e., if {ji1} and {ji2} are bosons. Our no-go

theorem applies to all bosons {ji1} and {ji2} with zero mode sets I{ji1} and I{ji2}. We

can then use the set Ij = I{ji1}×I{ji2}, containing the fusion product of any particle

in I{ji1} with any particle in I{ji2}, to prove that j cannot condense. To show that

Ij is a set of zero modes of j, the main challenge is to show that the product of any

two elements from Ij cannot condense. The product of any two elements from Ij is

always of the form {ji1ji2}. We have shown that when ji1 6= ji2 such particles cannot

condense. We therefore need only show that non-trivial particles of form {ji1ji2} with

ji1 = ji2 both bosons cannot condense. In order to show they are not condensable, we

can use the proof given for Example (ii). For that, observe that the anyons j have the

same fusion coefficients among themselves as the j anyons in SO(3)k in Example (ii)

have, i.e., N j′′

j,j′= N j′′

j,j′ , where j, j′, j′′ ∈ SO(3)k. Recall that conditions 1–3 from

the definition of a set of zero modes only depend on the fusion coefficients N j′′

j,j′ and

the information, which particles are bosons. Hence, conditions 1–3 are satisfied for

Ij whenever they are satisfied for Ij in Example (ii). It remains to show that Ij is of

large enough quantum dimension to satisfy the fundamental inequality Eq. (4) from

the main text. For j, Eq. (4) from the main text takes the form

dj = d2j <

(∑a∈Ij

da

)2

=∑a∈Ij

da. (B.7)

193

Upon taking the square root, this is equivalent to Eqs. (14) and (15) from the main

text, which were shown to hold in Example (ii). Therefore the j = {ji1ji2} anyons do

not condense and all {ji} have distinct restrictions.

We conclude that for any k0 ≥ 1 only Nϕ{ji1 ···jik0 }

↓,{jik0+1}↓ = 0 is permitted and

hence Eq. (B.5) implies that {ji1ji2 · · · jik0+1} does not restrict to the identity ϕ, i.e.,

it does not condense. This proves the assumption 1 of the induction step for k0 + 1.

To complete the induction step, we need to show that {ji1ji2 · · · jik0+1} does not

split. For that, consider Eq. (2) from the main text for {ji1 · · · jik0+1} with itself and

t = ϕ

∑r

(nr{ji1 ···jik0+1

}

)2

=∑c

N c{ji1 ···jik0+1

},{ji1 ···jik0+1} n

ϕc

=nϕ1 = 1. (B.8)

We have used that none of the {ji1ji2 · · · jil} with 1 ≤ l ≤ k0 + 1 can restrict to

the identity ϕ since they cannot condense. This implies none of {ji1 · · · jik0+1} splits,

which proves the assumption 2 of the induction step for k0 + 1.

We have thus shown inductively that none of the particles (except for the vacuum)

restricts to the vacuum in the N -layer theory SO(3)⊗Nk . Thus, there is no condensate

and with it no condensation in any number N of layers of SO(3)k with k odd.

B.3 General constraints on boson condensation

In this section, we list lemmas that pose other general constraints on condensation

transitions in TQFTs.

Lemma 2. Suppose S = {a1, · · · , am} is a collection of particles in a TQFT A with

ai × ai containing no bosons other than the identity – i.e., ntai = δta↓i

and ai does not

194

split. Moreover assume a↓i 6= a↓j for i 6= j. Then if a boson B appears in the fusion of

ai and aj, ai × aj = B + · · · for any i 6= j, then B is not condensable.

Proof. Using Eq. (2) from the main text for a = ai, b = aj and t = ϕ, we have

δij =∑

t ntaintaj =

∑c n

ϕcN

caiaj

. For i 6= j we get∑

B nϕBN

Baiaj

= 0. So if boson B

appears in ai × aj, we must have nϕB = 0, so that B is not condensable.

Lemma 3. Consider a TQFT A with no fusion multiplicity and just one boson B.

If B is condensed then either B is abelian or B↓ = ϕ+ r where r is a single anyon.

Proof. As there is just a single boson, B = B. Equation (2) from the main text

implies ∑t

ntBntB =

∑c

nϕcNcBB = 1 + nϕBN

BBB. (B.9)

Notice, however, that the left-hand side is greater or equal to nϕBnϕB. For condensation,

this implies nϕB = 1, and tells us that∑

t ntBn

tB = 1 or 2. In the former case, B↓ = ϕ.

This implies dB = 1, and so B is a quantum dimension 1 boson hence must have

NaBB = δa,1. In the latter case, B restricts to just two particles with multiplicity 1

each, so that B↓ = ϕ+ r.

Lemma 4. With the conditions of Lemma 3, and assuming B has dB > 1, conden-

sation of B can only occur if NaBBN

bBB ≤ NB

ab for all anyons a and b of A.

Proof. Lemma 3 shows B↓ = ϕ + r, where r a simple object. Consider a ∈ A where

a 6= 1, B. Equation (2) from the main text for b = B and t = ϕ reads

nranrB = nra = NB

aBnϕB = NB

aB(B.10)

Consider now Eq. (2) from the main text for b 6= 1, B and for a 6= b, which gives

∑t

ntantb = NB

ab ≥ nranrb (B.11)

195

Combining the a,B and b, B and a, b equations gives the inequality

N bBBN

aBB ≤ NB

ab. (B.12)

196

Appendix C

Appendix for Tensor Network

States and Fracton Models

C.1 Proof for the Concatenation Lemma for the

3D Toric Code Model

In this section, we prove the Concatenation lemma in Sec. 6.3.3 for the 3D toric

code model by induction. First of all, we propose and prove two lemmas:

(A) Let Tt1t2t3... be a contraction of a network of local T tensors, whose (i.e. open)

indices {t1t2t3...} are un-contracted virtual indices. If Tt1t2t3... satisfies the Con-

catenation lemma of the 3D toric code model in Sec. 6.3.3, then the contrac-

tion of Tt1t2t3... over a subset of its open virtual indices, say contracting over t1

and t2, i.e.,∑

t1t2Tt1t2t3...δt1t2 still satisfies the Concatenation lemma of the

3D toric code model in Sec. 6.3.3.

(B) If Tt1t2t3... and Tt1 t2 t3... are two networks of contracted local T tensors both

of which satisfying the Concatenation lemma of the 3D toric code

model in Sec. 6.3.3, then the contraction over one pair of indices, say

197

∑t1 t1

Tt1t2t3...Tt1 t2 t3...δt1 t1 , still satisfies the Concatenation lemma of the

3D toric code model in Sec. 6.3.3.

Proof:

(A): Since T satisfies the Concatenation lemma, its elements T{t} are:

T{t} =

0 if

∑i ti = 1 mod 2

N if∑

i ti = 0 mod 2,

(C.1)

where N is a constant independent of the open virtual indices in the Concatenation

lemma. Suppose that we contract two indices of T, tm, tn ∈ {t}, and we denote the

contraction as T′ and the remaining open virtual indices after the contraction {t′}.

Then we have:

T′{t′} =∑tm,tn

T{t}δtm,tn

=∑tm,tn

T...tm...tn...δtm,tn

=∑tm

T...tm...tm...

=T...0...0... + T...1...1...

(C.2)

Hence, the contraction still satisfies the Concatenation lemma:

T′{t′} =

0 if

∑i t′i = 1 mod 2

2N if∑

i t′i = 0 mod 2.

(C.3)

198

(B): Since T and T satisfy the Concatenation lemma, their elements T{t} and

T{t} are:

T{t} =

0 if

∑i ti = 1 mod 2

N if∑

i ti = 0 mod 2

T{t} =

0 if

∑i ti = 1 mod 2

N if∑

i ti = 0 mod 2,

(C.4)

where N and N are the constants independent of the indices {t} and {t} respectively.

Suppose that we contract two indices tm ∈ {t} and tn ∈ {t}, and we denote the

contraction as T′ and the remaining open virtual indices after the contraction {t′}.

Then we have:

T′{t′} =∑tm,tn

T{t}T{t}δtm,tn

=∑tm,tn

T...tm...T...tn...δtm,tn

=T...0...T...0... + T...1...T...1...,

(C.5)

The last line is nonzero if and only if∑

i 6=m ti and∑

j 6=n tn have the same parity. If

this parity is even (resp. odd), only T...0...T...0... (resp. T...1...T...1...) is nonzero and

equal to NN . Since∑

i t′i =

∑i 6=m ti +

∑j 6=n tn, we conclude that:

T′{t′} =

0 if

∑i t′i = 1 mod 2

NN if∑

i t′i = 0 mod 2.

(C.6)

T′{t′} still satisfies the Concatenation lemma. 2

Having proved Lemma (A) and (B), we can further prove that:

199

(C) If T and T are two networks of contracted local T tensors which both satisfy the

Concatenation lemma of the 3D toric code model in Sec. 6.3.3, then their

contraction over any pairs of indices still satisfies the Concatenation lemma

of the 3D toric code model in Sec. 6.3.3.

Proof:

We can decompose the contraction process into two steps: (1) contract T and T

over one pair of indices; (2) contract the rest of the indices. Lemma (B) guarantees

that the outcome tensor of the contraction (1) still satisfies the Concatenation

lemma. Lemma (A) guarantees that the outcome tensor of the contraction (2) also

satisfies the Concatenation lemma. Hence, Lemma (C) is proved. 2

Now we can complete the induction proof for the Concatenation lemma of the

3D toric code model: First of all, we point out the a single local T tensor satisfies

the Concatenation lemma. Next, we assume that two networks of contracted local

T tensors satisfy the Concatenation lemma, and prove that their contraction also

satisfies the Concatenation lemma. This induction step is, in fact, Lemma (C).

Therefore, we have completed the induction proof for the Concatenation lemma

of the 3D toric code model.

C.2 Proof for the Concatenation Lemma for the

X-cube Model

In this section, we prove the Concatenation lemma in Sec. 6.4.3 for the X-cube

model by induction. We point out that the subtlety of the Concatenation lemma

of the X-cube model is that the number of constraints for the open virtual indices is

linear with respect to the system size. More precisely, each constraint corresponds to

a set of open virtual indices along a xy, yz or xz plane to have an even summation.

See Eq. (6.86) for an example. Hence, we need to keep track of the constraints when

200

we contract more tensors. For clarity, we denote the sets of constraints of {t} as:

fxy({t}) :

{ ∑i∈xy-plane

ti, for all xy-planes

}

fyz({t}) :

{ ∑i∈yz-plane

ti, for all yz-planes

}

fxz({t}) :

{ ∑i∈xz-plane

ti, for all xz-planes

}.

(C.7)

Notice that these equations are not linearly independent. Their summation is auto-

matically true. Physically, we can view these quantities fxy({t}), fyz({t}) and fxz({t})

as the “parities” in each xy, yz and xz plane. To begin with, we propose and prove

Lemma (D) which is the induction step:

(D) If T is a network of contracted local T tensors of the X-cube model which

satisfies the Concatenation lemma of the X-cube model in Sec. 6.4.3, then

the contraction of T and a T tensor still satisfies the Concatenation lemma

of the X-cube model in Sec. 6.4.3.

Proof:

We first fix the notations: the elements of T are T{t}; {tc} = {tc1, tc2, . . .} ⊂ {t}

are the indices that contract with the indices of the local T tensor; and the outcome

tensor is denoted as T′{t′} where {t′} denotes the open virtual indices after contraction.

Since T is assumed to satisfy the Concatenation lemma, we have:

T{t} =

N if

fxy({t}) = 0 mod 2

fyz({t}) = 0 mod 2

fxz({t}) = 0 mod 2

0 otherwise,

(C.8)

201

where N is the constant independent of the open virtual indices {t}, and fxy({t}),

fyz({t}) and fxz({t}) denote the set of summations over the open virtual indices in

each xy, yz and xz plane. See Eq. (6.92) for an example when T is a contraction of two

local T tensors. Notice that the elements of Txxyyzz also satisfy the Concatenation

lemma for the X-cube model, as shown by Eq. (6.86).

Using these notations, the tensor contraction is just:

T′{t′} =∑{tc}

T...tc1...tc2...T...tc1...tc2.... (C.9)

We now discuss one particular way of contraction: contraction over one pair of indices.

Other contractions can be proved using the exact same method.

Suppose {tc} contains only one index. Without loss of generality, we assume that

this index is the x index of Txxyyzz. Then the tensor contraction is:

T′{t′} =∑x

T...x...Tx... (C.10)

Graphically,

xy

z

z

y

x . (C.11)

202

When T...x... and Tx... are both nonzero, the indices satisfy that:

fxy({t}/x, x) =0 mod 2

fxz({t}/x, x) =0 mod 2

x+ x+ y + y =0 mod 2

x+ x+ z + z =0 mod 2

fxy({t}/x, x, y, y) =0 mod 2

fxz({t}/x, x, z, z) =0 mod 2,

(C.12)

where {t}/x denotes the t indices excluding the x index. We only list the equations

whose variables include the index x. Then, we include the fyz constraints from T and

T , which can be concatenated into fyz({t′}). We find:

T′{t′} =

N if

fxy({t′}) = 0 mod 2

fyz({t′}) = 0 mod 2

fxz({t′}) = 0 mod 2

0 otherwise.

(C.13)

Hence, T′ still satisfies the Concatenation lemma of the X-cube model. We can

further contract other indices of T tensor with T. For instance, the index y of T

tensor with another index of T in the same plane. Then the outcome tensor still

satisfies the Concatenation lemma of the X-cube model, because (1) the “parities”

fxy, fyz and fxz do not change after contraction, (2) the contraction is the same for

the open indices of the same parities. Therefore, Lemma (D) is proved. 2

Having proven Lemma (D), now we can complete the induction proof for the

Concatenation lemma of the X-cube model: First of all, we point out that a single

203

local T tensor of the X-cube model satisfies the Concatenation lemma. Next, as

the induction step, we assume that one network of contracted local T tensors satisfies

the Concatenation lemma, and prove that contracting one more local T tensor also

satisfies the Concatenation lemma. This induction step is, in fact, Lemma (D).

Therefore, we have completed the induction proof for the Concatenation lemma

of the X-cube model in Sec. 6.4.3.

C.3 GSD for the X-cube Model

In this section, we work out the representation dimension of the operators in

Eq. (6.82). In particular, the first group of algebras is that WX [Cx] anti-commutes

with WZ [Cz,x] and WZ [Cy,x] when they have overlaps. In the projected yz plane, the

204

operators WX [Cx], WZ [Cz,x] and WZ [Cy,x] can be depicted as:

z (pbc)

y (pbc)

, (C.14)

where the blue dot and the blue lines denote the projected operators on the yz-plane.

There are LyLz number of WX operators and Ly +Lz number of WZ operators. The

205

anti-commutation relations happens when they have overlaps:

z (pbc)

y (pbc)

. (C.15)

Other combinations of operators commute. In a more pictorial language, the commu-

tation relations are just that the blue point at the coordinate (y, z) flips the vertical

line and the horizontal line passing the blue point (y, z). In this pictorial language,

we can find that we can flip any pair of lines independently using the points. In the

operator language, we can flip any pair of WZ operators using WX operators. For

206

instance:z (pbc)

y (pbc)

. (C.16)

Therefore, we set a “reference” line and flip all other Ly + Lz − 1 number of lines

using the dots. In the operator language, we set a “reference” WZ operator in this

projected yz-plane, and flip all other Ly +Lz − 1 number of WZ operators using WX

operators. Hence, we can generate 2Ly+Lz−1 dimensional Hilbert space using operators

WX [Cx], WZ [Cz,x] and WZ [Cy,x]. Similarly for other algebras below Eq. (6.82), we can

generate 2Lx+Lz−1 and 2Lx+Ly−1 dimensional Hilbert space respectively. The ground

state degeneracy is then their product: 22Lx+2Ly+2Lz−3.

207

Bibliography

[1] D. C. Tsui, H. L. Stormer, and A. C. Gossard. Two-dimensional magneto-transport in the extreme quantum limit. Phys. Rev. Lett., 48:1559–1562, May1982.

[2] R. B. Laughlin. Anomalous quantum hall effect: An incompressible quantumfluid with fractionally charged excitations. Phys. Rev. Lett., 50:1395–1398, May1983.

[3] X. G. Wen, Frank Wilczek, and A. Zee. Chiral spin states and superconductivity.Phys. Rev. B, 39:11413–11423, Jun 1989.

[4] X. G. Wen. Vacuum degeneracy of chiral spin states in compactified space.Phys. Rev. B, 40:7387–7390, Oct 1989.

[5] X. G. WEN. Topological orders in rigid states. International Journal of ModernPhysics B, 04(02):239–271, 1990.

[6] Xiao-Gang Wen. A theory of 2+1D bosonic topological orders. National ScienceReview, 3(1):68–106, 11 2015.

[7] Xie Chen, Zheng-Cheng Gu, and Xiao-Gang Wen. Local unitary transforma-tion, long-range quantum entanglement, wave function renormalization, andtopological order. Phys. Rev. B, 82:155138, Oct 2010.

[8] Zheng-Cheng Gu and Xiao-Gang Wen. Tensor-entanglement-filtering renor-malization approach and symmetry-protected topological order. Phys. Rev. B,80:155131, Oct 2009.

[9] Xiao-Gang Wen. Quantum field theory of many-body systems: from the origin ofsound to an origin of light and electrons. Oxford University Press on Demand,2004.

[10] F. D. M. Haldane. Model for a quantum hall effect without landau lev-els: Condensed-matter realization of the ”parity anomaly”. Phys. Rev. Lett.,61:2015–2018, Oct 1988.

[11] F. D. M. Haldane. Stability of chiral luttinger liquids and abelian quantum hallstates. Phys. Rev. Lett., 74:2090–2093, Mar 1995.

208

[12] C. L. Kane and E. J. Mele. Quantum spin hall effect in graphene. Phys. Rev.Lett., 95:226801, Nov 2005.

[13] B. Andrei Bernevig and Shou-Cheng Zhang. Quantum spin hall effect. Phys.Rev. Lett., 96:106802, Mar 2006.

[14] B. Andrei Bernevig, Taylor L. Hughes, and Shou-Cheng Zhang. Quantum spinhall effect and topological phase transition in hgte quantum wells. Science,314(5806):1757–1761, 2006.

[15] Congjun Wu, B. Andrei Bernevig, and Shou-Cheng Zhang. Helical liquid andthe edge of quantum spin hall systems. Phys. Rev. Lett., 96:106401, Mar 2006.

[16] N. Regnault and B. Andrei Bernevig. Fractional chern insulator. Phys. Rev. X,1:021014, Dec 2011.

[17] Yong-Shi Wu. Braid groups, anyons and gauge invariance. International Journalof Modern Physics B, 05(10):1649–1664, 1991.

[18] Yasuhiro Hatsugai, Mahito Kohmoto, and Yong-Shi Wu. Anyons on a torus:Braid group, aharonov-bohm period, and numerical study. Phys. Rev. B,43:10761–10768, May 1991.

[19] A.Yu. Kitaev. Fault-tolerant quantum computation by anyons. Annals ofPhysics, 303(1):2 – 30, 2003.

[20] Alexei Kitaev. Anyons in an exactly solved model and beyond. Annals ofPhysics, 321(1):2 – 111, 2006. January Special Issue.

[21] I. S. Eliens, J. C. Romers, and F. A. Bais. Diagrammatics for bose condensationin anyon theories. Phys. Rev. B, 90:195130, Nov 2014.

[22] Alexei Kitaev and John Preskill. Topological entanglement entropy. Phys. Rev.Lett., 96:110404, Mar 2006.

[23] Michael Levin and Xiao-Gang Wen. Detecting topological order in a groundstate wave function. Phys. Rev. Lett., 96:110405, Mar 2006.

[24] Wikipedia contributors. Greenberger-Horne-Zeilinger state — Wikipedia,the free encyclopedia. https://en.wikipedia.org/w/index.php?

title=Greenberger%E2%80%93Horne%E2%80%93Zeilinger_state&oldid=

884046173, 2019. [Online; accessed 21-April-2019].

[25] Mark Srednicki. Entropy and area. Phys. Rev. Lett., 71:666–669, Aug 1993.

[26] M B Hastings. An area law for one-dimensional quantum systems. Journal ofStatistical Mechanics: Theory and Experiment, 2007(08):P08024–P08024, aug2007.

209

[27] Hui Li and F. D. M. Haldane. Entanglement spectrum as a generalization of en-tanglement entropy: Identification of topological order in non-abelian fractionalquantum hall effect states. Phys. Rev. Lett., 101:010504, Jul 2008.

[28] Xiao-Liang Qi, Hosho Katsura, and Andreas W. W. Ludwig. General rela-tionship between the entanglement spectrum and the edge state spectrum oftopological quantum states. Phys. Rev. Lett., 108:196402, May 2012.

[29] David Perez-Garcia, Frank Verstraete, Michael M Wolf, and J Ignacio Cirac.Matrix product state representations. arXiv preprint quant-ph/0608197, 2006.

[30] Norbert Schuch, Ignacio Cirac, and David Perez-Garcıa. Peps as ground states:Degeneracy and topology. Annals of Physics, 325(10):2153–2192, 2010.

[31] Y.-Y. Shi, L.-M. Duan, and G. Vidal. Classical simulation of quantum many-body systems with a tree tensor network. Phys. Rev. A, 74:022320, Aug 2006.

[32] G. Vidal. Class of quantum many-body states that can be efficiently simulated.Phys. Rev. Lett., 101:110501, Sep 2008.

[33] Xiao-Gang Wen. Topological orders and edge excitations in fractional quantumhall states. Advances in Physics, 44(5):405–473, 1995.

[34] Alexander M Polyakov. Fermi-bose transmutations induced by gauge fields.Modern Physics Letters A, 3(03):325–328, 1988.

[35] Parsa Hassan Bonderson. Non-Abelian anyons and interferometry. PhD thesis,California Institute of Technology, 2007.

[36] Andrei Bernevig and Titus Neupert. Topological superconductors and categorytheory. Lecture Notes of the Les Houches Summer School: Topological Aspectsof Condensed Matter Physics, pages 63–121, 2017.

[37] F A Bais and J C Romers. The modularS-matrix as order parameter for topo-logical phase transitions. New Journal of Physics, 14(3):035024, mar 2012.

[38] Huan He, Heidar Moradi, and Xiao-Gang Wen. Modular matrices as topo-logical order parameter by a gauge-symmetry-preserved tensor renormalizationapproach. Phys. Rev. B, 90:205114, Nov 2014.

[39] Heidar Moradi and Xiao-Gang Wen. Universal wave-function overlap and uni-versal topological data from generic gapped ground states. Phys. Rev. Lett.,115:036802, Jul 2015.

[40] Fangzhou Liu, Zhenghan Wang, Yi-Zhuang You, and Xiao-Gang Wen. Modulartransformations and topological orders in two dimensions. arXiv e-prints, pagearXiv:1303.0829, Mar 2013.

[41] Mark Dirk Frederik de Wild Propitius. Topological interactions in broken gaugetheories. PhD thesis, Amsterdam U., 1995.

210

[42] F. A. Bais and J. K. Slingerland. Condensate-induced transitions between topo-logically ordered phases. Phys. Rev. B, 79:045316, Jan 2009.

[43] Yi-Zhuang You, Chao-Ming Jian, and Xiao-Gang Wen. Synthetic non-abelianstatistics by abelian anyon condensation. Phys. Rev. B, 87:045106, Jan 2013.

[44] Maissam Barkeshli and Xiao-Gang Wen. Anyon condensation and continu-ous topological phase transitions in non-abelian fractional quantum hall states.Phys. Rev. Lett., 105:216804, Nov 2010.

[45] F J Burnell, Steven H Simon, and J K Slingerland. Phase transitions in topolog-ical lattice models via topological symmetry breaking. New Journal of Physics,14(1):015004, jan 2012.

[46] F. J. Burnell, Steven H. Simon, and J. K. Slingerland. Condensation of achiralsimple currents in topological lattice models: Hamiltonian study of topologicalsymmetry breaking. Phys. Rev. B, 84:125434, Sep 2011.

[47] Ling-Yan Hung and Yidun Wan. Symmetry-enriched phases obtained viapseudo anyon condensation. International Journal of Modern Physics B,28(24):1450172, 2014.

[48] Ling-Yan Hung and Yidun Wan. Generalized ade classification of topolog-ical boundaries and anyon condensation. Journal of High Energy Physics,2015(7):120, Jul 2015.

[49] F. A. Bais and C. J. M. Mathy. Defect-mediated melting and the breaking ofquantum double symmetries. Phys. Rev. B, 73:224120, Jun 2006.

[50] C.J.M. Mathy and F.A. Bais. Nematic phases and the breaking of doublesymmetries. Annals of Physics, 322(3):709 – 735, 2007.

[51] F. A. Bais, B. J. Schroers, and J. K. Slingerland. Broken quantum symmetryand confinement phases in planar physics. Phys. Rev. Lett., 89:181601, Oct2002.

[52] F. Alexander Bais, Peter van Driel, and Mark de Wild Propitius. Quantumsymmetries in discrete gauge theories. Physics Letters B, 280(1):63 – 70, 1992.

[53] Teresia Mansson, Ville Lahtinen, Juha Suorsa, and Eddy Ardonne. Condensate-induced transitions and critical spin chains. Phys. Rev. B, 88:041403, Jul 2013.

[54] Yuxiang Gu, Ling-Yan Hung, and Yidun Wan. Unified framework of topologicalphases with symmetry. Phys. Rev. B, 90:245125, Dec 2014.

[55] Marc Daniel Schulz, Sebastien Dusuel, Gregoire Misguich, Kai Phillip Schmidt,and Julien Vidal. Ising anyons with a string tension. Phys. Rev. B, 89:201103,May 2014.

211

[56] Marc Daniel Schulz, Sebastien Dusuel, Kai Phillip Schmidt, and Julien Vidal.Topological phase transitions in the golden string-net model. Phys. Rev. Lett.,110:147203, Apr 2013.

[57] H. Bombin and M. A. Martin-Delgado. Family of non-abelian kitaev models ona lattice: Topological condensation and confinement. Phys. Rev. B, 78:115421,Sep 2008.

[58] Sergey B Bravyi and A Yu Kitaev. Quantum codes on a lattice with boundary.arXiv preprint quant-ph/9811052, 1998.

[59] Salman Beigi, Peter W. Shor, and Daniel Whalen. The quantum double modelwith boundary: Condensations and symmetries. Communications in Mathe-matical Physics, 306(3):663–694, Sep 2011.

[60] Anton Kapustin and Natalia Saulina. Topological boundary conditions inabelian chernsimons theory. Nuclear Physics B, 845(3):393 – 435, 2011.

[61] Michael Levin. Protected edge modes without symmetry. Phys. Rev. X,3:021009, May 2013.

[62] Liang Kong. Anyon condensation and tensor categories. Nuclear Physics B,886:436 – 482, 2014.

[63] F. A. Bais, J. K. Slingerland, and S. M. Haaker. Theory of topological edgesand domain walls. Phys. Rev. Lett., 102:220403, Jun 2009.

[64] Tian Lan, Juven C. Wang, and Xiao-Gang Wen. Gapped domain walls, gappedboundaries, and topological degeneracy. Phys. Rev. Lett., 114:076402, Feb 2015.

[65] Yasuyuki Kawahigashi. A remark on gapped domain walls between topologicalphases. Letters in Mathematical Physics, 105(7):893–899, Jul 2015.

[66] Ling-Yan Hung and Yidun Wan. Ground-state degeneracy of topological phaseson open surfaces. Phys. Rev. Lett., 114:076401, Feb 2015.

[67] F. A. Bais and S. M. Haaker. Topological symmetry breaking: Domain wallsand partial instability of chiral edges. Phys. Rev. B, 92:075427, Aug 2015.

[68] Juven C. Wang and Xiao-Gang Wen. Boundary degeneracy of topological order.Phys. Rev. B, 91:125124, Mar 2015.

[69] Alexei Kitaev and Liang Kong. Models for gapped boundaries and domainwalls. Communications in Mathematical Physics, 313(2):351–373, Jul 2012.

[70] E. Grosfeld and K. Schoutens. Non-abelian anyons: When ising meets fibonacci.Phys. Rev. Lett., 103:076803, Aug 2009.

[71] FA Bais and JK Slingerland. Topological entanglement entropy relations formulti phase systems with interfaces. arXiv preprint arXiv:1006.2017, 2010.

212

[72] H. Bombin. Topological order with a twist: Ising anyons from an abelian model.Phys. Rev. Lett., 105:030403, Jul 2010.

[73] Maissam Barkeshli and Xiao-Liang Qi. Topological nematic states and non-abelian lattice dislocations. Phys. Rev. X, 2:031013, Aug 2012.

[74] Maissam Barkeshli, Chao-Ming Jian, and Xiao-Liang Qi. Twist defects andprojective non-abelian braiding statistics. Phys. Rev. B, 87:045130, Jan 2013.

[75] Maissam Barkeshli, Chao-Ming Jian, and Xiao-Liang Qi. Theory of defects inabelian topological states. Phys. Rev. B, 88:235103, Dec 2013.

[76] Maissam Barkeshli, Chao-Ming Jian, and Xiao-Liang Qi. Classification of topo-logical defects in abelian topological states. Phys. Rev. B, 88:241103, Dec 2013.

[77] Yi-Zhuang You and Xiao-Gang Wen. Projective non-abelian statistics of dislo-cation defects in a zN rotor model. Phys. Rev. B, 86:161107, Oct 2012.

[78] Jeffrey C. Y. Teo, Abhishek Roy, and Xiao Chen. Unconventional fusion andbraiding of topological defects in a lattice model. Phys. Rev. B, 90:115118, Sep2014.

[79] Nicolas Tarantino, Netanel H Lindner, and Lukasz Fidkowski. Symmetry frac-tionalization and twist defects. New Journal of Physics, 18(3):035006, mar2016.

[80] Edward Witten. Quantum field theory and the jones polynomial. Comm. Math.Phys., 121(3):351–399, 1989.

[81] Jrg Frhlich, Jrgen Fuchs, Ingo Runkel, and Christoph Schweigert. Correspon-dences of ribbon categories. Advances in Mathematics, 199(1):192 – 329, 2006.

[82] Michael A. Levin and Xiao-Gang Wen. String-net condensation: A physicalmechanism for topological phases. Phys. Rev. B, 71:045110, Jan 2005.

[83] Tian Lan and Xiao-Gang Wen. Topological quasiparticles and the holographicbulk-edge relation in (2 + 1)-dimensional string-net models. Phys. Rev. B,90:115119, Sep 2014.

[84] Liang Kong and Xiao-Gang Wen. Braided fusion categories, gravitationalanomalies, and the mathematical framework for topological orders in any di-mensions. arXiv preprint arXiv:1405.5858, 2014.

[85] Liang Kong, Xiao-Gang Wen, and Hao Zheng. Boundary-bulk relation fortopological orders as the functor mapping higher categories to their centers.arXiv preprint arXiv:1502.01690, 2015.

[86] Jrgen Fuchs, Ingo Runkel, and Christoph Schweigert. Conformal correlationfunctions, frobenius algebras and triangulations. Nuclear Physics B, 624(3):452– 468, 2002.

213

[87] Jurgen Fuchs and Christoph Schweigert. Category theory for conformal bound-ary conditions. Fields Institute Commun, 39:25, 2003.

[88] J. Fuchs, I. Runkel, and C. Schweigert. Boundaries, defects and frobeniusalgebras. Fortschritte der Physik, 51(78):850–855.

[89] Jrgen Fuchs, Christoph Schweigert, and Carl Stigner. The classifying algebrafor defects. Nuclear Physics B, 843(3):673 – 723, 2011.

[90] Michael Muger. From subfactors to categories and topology i: Frobenius alge-bras in and morita equivalence of tensor categories. Journal of Pure and AppliedAlgebra, 180(1-2):81–157, 2003.

[91] Alexander Kirillov, Jr. and Viktor Ostrik. On q-analog of McKay correspon-dence and ADE classification of affine sl(2) conformal field theories. 2001.

[92] Titus Neupert, Huan He, Curt von Keyserlingk, German Sierra, and B. AndreiBernevig. Boson condensation in topologically ordered quantum liquids. Phys.Rev. B, 93:115103, Mar 2016.

[93] A. B. Zamolodchikov. Irreversibility of the Flux of the Renormalization Groupin a 2D Field Theory. JETP Lett., 43:730–732, 1986. [Pisma Zh. Eksp. Teor.Fiz.43,565(1986)].

[94] Ian Affleck and Andreas W. W. Ludwig. Universal noninteger “ground-statedegeneracy” in critical quantum systems. Phys. Rev. Lett., 67:161–164, Jul1991.

[95] Daniel Friedan and Anatoly Konechny. Boundary entropy of one-dimensionalquantum systems at low temperature. Phys. Rev. Lett., 93:030402, Jul 2004.

[96] Daniel Green, Michael Mulligan, and David Starr. Boundary entropy can in-crease under bulk rg flow. Nuclear Physics B, 798(3):491 – 504, 2008.

[97] A.N. Schellekens and S. Yankielowicz. Extended chiral algebras and modularinvariant partition functions. Nuclear Physics B, 327(3):673 – 703, 1989.

[98] A. N. Schellekens and S. Yankielowicz. Simple Currents, Modular Invariantsand Fixed Points. Int. J. Mod. Phys., A5:2903–2952, 1990.

[99] Kenneth A. Intriligator. Bonus Symmetry in Conformal Field Theory. Nucl.Phys., B332:541–565, 1990.

[100] B. Gato-Rivera and A. N. Schellekens. Complete classification of simple currentautomorphisms. Nucl. Phys., B353:519–537, 1991.

[101] Beatriz Gato-Rivera and Adrian Norbert Schellekens. Complete classificationof simple current modular invariants for rcft’s with a center (z p) k. Commu-nications in mathematical physics, 145(1):85–121, 1992.

214

[102] M Kreuzer and Adrian Norbert Schellekens. Simple currents versus orbifoldswith discrete torsiona complete classification. Nuclear Physics B, 411(1):97–121,1994.

[103] J. Fuchs, A.N. Schellekens, and C. Schweigert. A matrix s for all simple currentextensions. Nuclear Physics B, 473(1):323 – 366, 1996.

[104] Cumrun Vafa. Toward Classification of Conformal Theories. Phys. Lett.,B206:421–426, 1988.

[105] Maissam Barkeshli, Parsa Bonderson, Meng Cheng, and Zhenghan Wang. Sym-metry, Defects, and Gauging of Topological Phases. 2014.

[106] Gregory Moore and Nathan Seiberg. Classical and quantum conformal fieldtheory. Comm. Math. Phys., 123(2):177–254, 1989.

[107] Gregory W. Moore and Nathan Seiberg. Naturality in Conformal Field Theory.Nucl. Phys., B313:16–40, 1989.

[108] Chao-Ming Jian and Xiao-Liang Qi. Layer construction of 3d topological statesand string braiding statistics. Phys. Rev. X, 4:041043, Dec 2014.

[109] T.H. Hansson, Vadim Oganesyan, and S.L. Sondhi. Superconductors are topo-logically ordered. Annals of Physics, 313(2):497 – 538, 2004.

[110] Lukasz Fidkowski, Xie Chen, and Ashvin Vishwanath. Non-abelian topologicalorder on the surface of a 3d topological superconductor from an exactly solvedmodel. Phys. Rev. X, 3:041016, Nov 2013.

[111] Max A. Metlitski, Lukasz Fidkowski, Xie Chen, and Ashvin Vishwanath. Inter-action effects on 3D topological superconductors: surface topological order fromvortex condensation, the 16 fold way and fermionic Kramers doublets. 2014.

[112] Alexei Davydov and Tom Booker. Commutative Algebras in Fibonacci Cate-gories. 2011.

[113] Titus Neupert, Huan He, Curt von Keyserlingk, German Sierra, and B. AndreiBernevig. Boson condensation in topologically ordered quantum liquids. Phys.Rev. B, 93:115103, Mar 2016.

[114] Andrea Cappelli, C. Itzykson, and J. B. Zuber. Modular Invariant PartitionFunctions in Two-Dimensions. Nucl. Phys., B280:445–465, 1987.

[115] P. Di Francesco, P. Mathieu, and D. Senechal. Conformal Field Theory. Grad-uate Texts in Contemporary Physics. Springer-Verlag, New York, 1997.

[116] Michael Levin and Zheng-Cheng Gu. Braiding statistics approach to symmetry-protected topological phases. Phys. Rev. B, 86:115109, Sep 2012.

215

[117] Xie Chen, Zheng-Cheng Gu, Zheng-Xin Liu, and Xiao-Gang Wen. Symme-try protected topological orders and the group cohomology of their symmetrygroup. Phys. Rev. B, 87:155114, Apr 2013.

[118] Davide Gaiotto, Anton Kapustin, Nathan Seiberg, and Brian Willett. Gener-alized global symmetries. Journal of High Energy Physics, 2015(2):172, Feb2015.

[119] Anton Kapustin and Nathan Seiberg. Coupling a qft to a tqft and duality.Journal of High Energy Physics, 2014(4):1, Apr 2014.

[120] Titus Neupert, Huan He, Curt von Keyserlingk, Germn Sierra, and B AndreiBernevig. No-go theorem for boson condensation in topologically ordered quan-tum liquids. New Journal of Physics, 18(12):123009, 2016.

[121] Xiao-Gang Wen. Zoo of quantum-topological phases of matter. arXiv preprintarXiv:1610.03911, 2016.

[122] Frank Pollmann, Erez Berg, Ari M. Turner, and Masaki Oshikawa. Symme-try protection of topological phases in one-dimensional quantum spin systems.Phys. Rev. B, 85:075125, Feb 2012.

[123] Michael P. Zaletel and Roger S. K. Mong. Exact matrix product states forquantum hall wave functions. Phys. Rev. B, 86:245305, Dec 2012.

[124] Michael P. Zaletel, Roger S. K. Mong, and Frank Pollmann. Topological char-acterization of fractional quantum hall ground states from microscopic hamil-tonians. Phys. Rev. Lett., 110:236801, Jun 2013.

[125] B. Estienne, Z. Papic, N. Regnault, and B. A. Bernevig. Matrix product statesfor trial quantum hall states. Phys. Rev. B, 87:161112, Apr 2013.

[126] B Estienne, N Regnault, and BA Bernevig. Fractional quantum hall ma-trix product states for interacting conformal field theories. arXiv preprintarXiv:1311.2936, 2013.

[127] Yang-Le Wu, B. Estienne, N. Regnault, and B. Andrei Bernevig. Braidingnon-abelian quasiholes in fractional quantum hall states. Phys. Rev. Lett.,113:116801, Sep 2014.

[128] B. Estienne, N. Regnault, and B. A. Bernevig. Correlation lengths and topolog-ical entanglement entropies of unitary and nonunitary fractional quantum hallwave functions. Phys. Rev. Lett., 114:186801, May 2015.

[129] Yang-Le Wu, B. Estienne, N. Regnault, and B. Andrei Bernevig. Matrix productstate representation of non-abelian quasiholes. Phys. Rev. B, 92:045109, Jul2015.

216

[130] Adolfo G. Grushin, Johannes Motruk, Michael P. Zaletel, and Frank Pollmann.Characterization and stability of a fermionic ν = 1/3 fractional chern insulator.Phys. Rev. B, 91:035136, Jan 2015.

[131] Michael P. Zaletel, Roger S. K. Mong, Frank Pollmann, and Edward H. Rezayi.Infinite density matrix renormalization group for multicomponent quantum hallsystems. Phys. Rev. B, 91:045115, Jan 2015.

[132] Scott Geraedts, Michael P. Zaletel, Zlatko Papic, and Roger S. K. Mong. Com-peting abelian and non-abelian topological orders in ν = 1/3 + 1/3 quantumhall bilayers. Phys. Rev. B, 91:205139, May 2015.

[133] Nicolas Regnault. Entanglement spectroscopy and its application to the quan-tum hall effects. Topological Aspects of Condensed Matter Physics: LectureNotes of the Les Houches Summer School: Volume 103, August 2014, 103:165,2017.

[134] S. D. Geraedts, C. Repellin, Chong Wang, Roger S. K. Mong, T. Senthil, andN. Regnault. Emergent particle-hole symmetry in spinful bosonic quantum hallsystems. Phys. Rev. B, 96:075148, Aug 2017.

[135] Norbert Schuch, Didier Poilblanc, J. Ignacio Cirac, and David Perez-Garcıa.Resonating valence bond states in the peps formalism. Phys. Rev. B, 86:115108,Sep 2012.

[136] Ling Wang, Didier Poilblanc, Zheng-Cheng Gu, Xiao-Gang Wen, and FrankVerstraete. Constructing a gapless spin-liquid state for the spin-1/2 J1 − J2

heisenberg model on a square lattice. Phys. Rev. Lett., 111:037202, Jul 2013.

[137] Mohsin Iqbal, Didier Poilblanc, and Norbert Schuch. Semionic resonatingvalence-bond states. Phys. Rev. B, 90:115129, Sep 2014.

[138] Didier Poilblanc, Philippe Corboz, Norbert Schuch, and J. Ignacio Cirac.Resonating-valence-bond superconductors with fermionic projected entangledpair states. Phys. Rev. B, 89:241106, Jun 2014.

[139] Didier Poilblanc, J. Ignacio Cirac, and Norbert Schuch. Chiral topological spinliquids with projected entangled pair states. Phys. Rev. B, 91:224431, Jun 2015.

[140] Matthieu Mambrini, Roman Orus, and Didier Poilblanc. Systematic construc-tion of spin liquids on the square lattice from tensor networks with su(2) sym-metry. Phys. Rev. B, 94:205124, Nov 2016.

[141] Jia-Wei Mei, Ji-Yao Chen, Huan He, and Xiao-Gang Wen. Gapped spin liquidwith z2 topological order for the kagome heisenberg model. Phys. Rev. B,95:235107, Jun 2017.

[142] Jeongwan Haah. Local stabilizer codes in three dimensions without string logicaloperators. Phys. Rev. A, 83:042330, Apr 2011.

217

[143] Sergey Bravyi and Jeongwan Haah. Quantum self-correction in the 3d cubiccode model. Phys. Rev. Lett., 111:200501, Nov 2013.

[144] Jeongwan Haah. Commuting pauli hamiltonians as maps between free modules.Communications in Mathematical Physics, 324(2):351–399, Dec 2013.

[145] Jeongwan Haah. Lattice quantum codes and exotic topological phases of matter.arXiv preprint arXiv:1305.6973, 2013.

[146] Jeongwan Haah. Bifurcation in entanglement renormalization group flow of agapped spin model. Phys. Rev. B, 89:075119, Feb 2014.

[147] Sagar Vijay, Jeongwan Haah, and Liang Fu. A new kind of topological quantumorder: A dimensional hierarchy of quasiparticles built from stationary excita-tions. Phys. Rev. B, 92:235136, Dec 2015.

[148] Jeongwan Haah. Algebraic methods for quantum codes on lattices. RevistaColombiana de Matematicas, 50(2):299–349, 2016.

[149] T. Devakul, S. A. Parameswaran, and S. L. Sondhi. Correlation function diag-nostics for type-I fracton phases. September 2017.

[150] Gabor B. Halasz, Timothy H. Hsieh, and Leon Balents. Fracton topologicalphases from strongly coupled spin chains. 2017.

[151] Timothy H. Hsieh and Gabor B. Halasz. Fractons from Partons. 2017.

[152] Sagar Vijay and Liang Fu. A Generalization of Non-Abelian Anyons in ThreeDimensions. 2017.

[153] Dominic J. Williamson. Fractal symmetries: Ungauging the cubic code. Phys.Rev. B, 94:155128, Oct 2016.

[154] M. Pretko. Higher-Spin Witten Effect and Two-Dimensional Fracton Phases.ArXiv e-prints, July 2017.

[155] K. Slagle and Y. B. Kim. Quantum Field Theory of X-Cube Fracton TopologicalOrder and Robust Degeneracy from Geometry. ArXiv e-prints, August 2017.

[156] O. Petrova and N. Regnault. A simple anisotropic three-dimensional quantumspin liquid with fracton topological order. ArXiv e-prints, September 2017.

[157] A. Prem, M. Pretko, and R. Nandkishore. Emergent Phases of Fractonic Matter.ArXiv e-prints, September 2017.

[158] H. Ma, A. T. Schmitz, S. A. Parameswaran, M. Hermele, and R. M. Nandk-ishore. Topological Entanglement Entropy of Fracton Stabilizer Codes. ArXive-prints, October 2017.

218

[159] M. Pretko. Finite-temperature screening of U(1) fractons. Phys. Rev. B,96(11):115102, September 2017.

[160] B. Shi and Y.-M. Lu. Decipher the nonlocal entanglement entropy of fractontopological orders. ArXiv e-prints, May 2017.

[161] M. Pretko. Emergent gravity of fractons: Mach’s principle revisited. Phys. Rev.D, 96(2):024051, July 2017.

[162] M. Pretko. Generalized electromagnetism of subdimensional particles: A spinliquid story. Phys. Rev. B, 96(3):035119, July 2017.

[163] D. A. Johnston, M. Mueller, and W. Janke. Plaquette Ising models, degeneracyand scaling. European Physical Journal Special Topics, 226, March 2017.

[164] Michael Pretko and Leo Radzihovsky. Fracton-elasticity duality. arXiv preprintarXiv:1711.11044, 2017.

[165] AT Schmitz, Han Ma, Rahul M Nandkishore, and SA Parameswaran. Recov-erable information and emergent conservation laws in fracton stabilizer codes.arXiv preprint arXiv:1712.02375, 2017.

[166] Kevin Slagle and Yong Baek Kim. X-cube model on generic lattices: Newphases and geometric order. arXiv preprint arXiv:1712.04511, 2017.

[167] Wilbur Shirley, Kevin Slagle, Zhenghan Wang, and Xie Chen. Fracton modelson general three-dimensional manifolds. arXiv preprint arXiv:1712.05892, 2017.

[168] Andrey Gromov. Fractional topological elasticity and fracton order. arXivpreprint arXiv:1712.06600, 2017.

[169] Daniel Gottesman. Class of quantum error-correcting codes saturating thequantum hamming bound. Phys. Rev. A, 54:1862–1868, Sep 1996.

[170] Sergey Bravyi, Bernhard Leemhuis, and Barbara M. Terhal. Topological orderin an exactly solvable 3d spin model. Annals of Physics, 326(4):839 – 866, 2011.

[171] Claudio Chamon. Quantum glassiness in strongly correlated clean systems: Anexample of topological overprotection. Phys. Rev. Lett., 94:040402, Jan 2005.

[172] Michael Pretko. Subdimensional particle structure of higher rank u(1) spinliquids. Phys. Rev. B, 95:115139, Mar 2017.

[173] Luka Trifunovic, Daniel Loss, and Jelena Klinovaja. From coupled rashbaelectron- and hole-gas layers to three-dimensional topological insulators. Phys.Rev. B, 93:205406, May 2016.

[174] Abhinav Prem, Jeongwan Haah, and Rahul Nandkishore. Glassy quantumdynamics in translation invariant fracton models. Phys. Rev. B, 95:155133, Apr2017.

219

[175] Jeongwan and Haah and Liang Fu. Fracton topological order, generalized latticegauge theory, and duality. Physical Review B, 94(23):235157, 2016.

[176] Han Ma, Ethan Lake, Xie Chen, and Michael Hermele. Fracton topologicalorder via coupled layers. Physical Review B, 95(24):245126, 2017.

[177] Sagar Vijay. Isotropic layer construction and phase diagram for fracton topo-logical phases. arXiv preprint arXiv:1701.00762, 2017.

[178] Norbert Schuch, Didier Poilblanc, J Ignacio Cirac, and David Perez-Garcia.Topological order in the projected entangled-pair states formalism: Transferoperator and boundary hamiltonians. Physical Review Letters, 111(9):090501,2013.

[179] Jutho Haegeman, Valentin Zauner, Norbert Schuch, and Frank Verstraete.Shadows of anyons and the entanglement structure of topological phases. Naturecommunications, 6, 2015.

[180] J Ignacio Cirac, Didier Poilblanc, Norbert Schuch, and Frank Verstraete. Entan-glement spectrum and boundary theories with projected entangled-pair states.Physical Review B, 83(24):245134, 2011.

[181] Wen Wei Ho, Lukasz Cincio, Heidar Moradi, Davide Gaiotto, and Guifre Vidal.Edge-entanglement spectrum correspondence in a nonchiral topological phaseand kramers-wannier duality. Physical Review B, 91(12):125119, 2015.

[182] Liujun Zou and Jeongwan Haah. Spurious long-range entanglement and replicacorrelation length. Phys. Rev. B, 94:075151, Aug 2016.

[183] Yuting Hu, Yidun Wan, and Yong-Shi Wu. Twisted quantum double model oftopological phases in two dimensions. Phys. Rev. B, 87:125114, Mar 2013.

[184] Yidun Wan, Juven C. Wang, and Huan He. Twisted gauge theory model oftopological phases in three dimensions. Phys. Rev. B, 92:045101, Jul 2015.

[185] Huan He, Yunqin Zheng, and Curt von Keyserlingk. Field theories for gaugedsymmetry-protected topological phases: Non-abelian anyons with abelian gaugegroup z

⊗3

2 . Phys. Rev. B, 95:035131, Jan 2017.

[186] Kevin Walker and Zhenghan Wang. (3+1)-tqfts and topological insulators.Frontiers of Physics, 7(2):150–159, Apr 2012.

[187] C. W. von Keyserlingk, F. J. Burnell, and S. H. Simon. Three-dimensionaltopological lattice models with surface anyons. Phys. Rev. B, 87:045107, Jan2013.

[188] F. J. Burnell, C. W. von Keyserlingk, and S. H. Simon. Phase transitions inthree-dimensional topological lattice models with surface anyons. Phys. Rev.B, 88:235120, Dec 2013.

220

[189] C. W. von Keyserlingk and F. J. Burnell. Walker-wang models and axionelectrodynamics. Phys. Rev. B, 91:045134, Jan 2015.

[190] Y. Zheng, H. He, B. Bradlyn, J. Cano, T. Neupert, and B. A. Bernevig. Struc-ture of the Entanglement Entropy of (3+1)D Gapped Phases of Matter. ArXive-prints, October 2017.

[191] Oliver Buerschaper. Twisted injectivity in projected entangled pair states andthe classification of quantum phases. Annals of Physics, 351:447–476, 2014.

[192] Oliver Buerschaper, Miguel Aguado, and Guifre Vidal. Explicit tensor net-work representation for the ground states of string-net models. Phys. Rev. B,79:085119, Feb 2009.

[193] Zheng-Cheng Gu, Michael Levin, Brian Swingle, and Xiao-Gang Wen. Tensor-product representations for string-net condensed states. Phys. Rev. B,79:085118, Feb 2009.

[194] C Wille, O Buerschaper, and J Eisert. Fermionic topological quantum statesas tensor networks. Physical Review B, 95(24):245127, 2017.

[195] Dominic J Williamson, Nick Bultinck, Jutho Haegeman, and Frank Verstraete.Fermionic matrix product operators and topological phases of matter. arXivpreprint arXiv:1609.02897, 2016.

[196] Mehmet Burak Sahinoglu, Dominic Williamson, Nick Bultinck, Michael Marien,Jutho Haegeman, Norbert Schuch, and Frank Verstraete. Characterizing topo-logical order with matrix product operators. arXiv preprint arXiv:1409.2150,2014.

[197] Sujeet K Shukla, M Burak Sahinoglu, Frank Pollmann, and Xie Chen. Bosoncondensation and instability in the tensor network representation of string-netstates. arXiv preprint arXiv:1610.00608, 2016.

[198] Michael Marien, Jutho Haegeman, Paul Fendley, and Frank Verstraete.Condensation-driven phase transitions in perturbed string nets. arXiv preprintarXiv:1607.05296, 2016.

[199] Kasper Duivenvoorden, Mohsin Iqbal, Jutho Haegeman, Frank Verstraete, andNorbert Schuch. Entanglement phases as holographic duals of anyon conden-sates. Phys. Rev. B, 95:235119, Jun 2017.

[200] Jose Garre-Rubio, Sofyan Iblisdir, and David Perez-Garcıa. Symmetry reduc-tion induced by anyon condensation: a tensor network approach. arXiv preprintarXiv:1702.08759, 2017.

[201] H Casini and M Huerta. Entanglement entropy in free quantum field theory.Journal of Physics A: Mathematical and Theoretical, 42(50):504007, 2009.

221

[202] Jan de Boer and Jacob Goeree. Markov traces and ii1 factors in conformal fieldtheory. Communications in Mathematical Physics, 139(2):267–304, Aug 1991.

222


Recommended