The boundaries and twist defects of the color code and their
applications to topological quantum computation
Markus S. Kesselring
1
, Fernando Pastawski
1
, Jens Eisert
1
, and Benjamin J. Brown
2,3
1
Dahlem Center for Complex Quantum Systems, Freie Universit
¨
at Berlin, 14195 Berlin, Germany
2
Centre for Engineered Quantum Systems, School of Physics, University of Sydney, Sydney, New South Wales 2006, Australia
3
Niels Bohr International Academy, Niels Bohr Institute, Blegdamsvej 17, 2100 Copenhagen, Denmark
October 25, 2018
The color code is both an interesting ex-
ample of an exactly solvable topologically or-
dered phase of matter and also among the most
promising candidate models to realize fault-
tolerant quantum computation with minimal
resource overhead. The contributions of this
work are threefold. First of all, we build upon
the abstract theory of boundaries and domain
walls of topological phases of matter to com-
prehensively catalog the objects as they are re-
alizable in color codes. Together with our clas-
sification we also provide lattice representa-
tions of these objects which include three new
types of boundaries as well as a generating set
for all 72 color code twist defects. Our work
thus provides an explicit toy model that will
help to better understand the abstract theory
of domain walls. Secondly, building upon the
established framework, we discover a number
of interesting new applications of the cataloged
objects for devising quantum information pro-
tocols. These include improved methods for
performing quantum computations by code de-
formation, a new four-qubit error-detecting
code, as well as families of new quantum error-
correcting codes we call stellated color codes,
which encode logical qubits at the same dis-
tance as the next best color code, but us-
ing approximately half the number of physi-
cal qubits. To the best of our knowledge, our
new topological codes have the highest encod-
ing rate of local stabilizer codes with bounded-
weight stabilizers in two dimensions. Finally,
we show how the boundaries and twist defects
of the color code are represented by multiple
copies of other phases. Indeed, in addition to
the well studied comparison between the color
code and two copies of the surface code, we
also compare the color code to two copies of
the three-fermion model. In particular, we
find that this analogy offers a very clear lens
through which we can view the symmetries of
the color code which gives rise to its multitude
of domain walls.
Markus S. Kesselring: markus.kesselring@fu-berlin.de
Contents
1 Introduction 2
2 Anyon models 3
3 The two-dimensional color code 4
3.1 The stabilizer formalism . . . . . . . . 4
3.2 The color code lattice and stabilizers . 4
3.3 Color code anyons . . . . . . . . . . . 5
3.4 Fusion . . . . . . . . . . . . . . . . . . 5
3.5 Self-exchange and braiding . . . . . . . 6
4 Boundaries 7
4.1 Lagrangian subgroups . . . . . . . . . 7
4.2 The boundaries of the color code . . . 8
4.3 Lattice representations of the color
code boundaries . . . . . . . . . . . . . 9
5 Small color codes and fault-tolerant
quantum computation 10
5.1 Code deformation . . . . . . . . . . . . 10
5.2 A four-qubit color code . . . . . . . . 10
6 Bulk domain walls 11
6.1 Symmetries in abelian anyon models . 11
6.2 Symmetries in the color code . . . . . 12
6.3 Lattice representation of domain walls 12
7 Twists 14
7.1 Twist defects . . . . . . . . . . . . . . 14
7.2 Fusion between twists . . . . . . . . . 15
7.3 Anyon localization and the quantum
dimensions of twists . . . . . . . . . . 16
7.4 Interplay between twists and domain
walls . . . . . . . . . . . . . . . . . . . 17
7.5 Confinement of twists at boundaries . 18
7.6 Detaching corners and twist condensa-
tion . . . . . . . . . . . . . . . . . . . 19
8 Stellated color codes 21
9 Unfolding boundaries and twists 22
10 Conclusions and future work 24
Acknowledgments 24
A Fermions in the color code 25
B The three-fermion model and the color
code 25
C Pachner moves and color code twists 26
D Stellated surface codes 29
References 29
1
arXiv:1806.02820v3 [quant-ph] 24 Oct 2018
1 Introduction
Topological phases of matter are of significant inter-
est both from a condensed matter perspective [1] as
well as for their potential application in fault-tolerant
quantum computation [29]. Indeed, among the lead-
ing architectures that have been proposed for scal-
able quantum-information processing are topological
quantum error-correcting codes. The physics of these
phases is enriched further when their low-energy ex-
citations are connected by symmetries, in which case
they give rise to non-trivial domain walls and point-
like topological defects [1016] that have been utilized
in a number of instances [12, 1727] to improve pro-
tocols for topological quantum computation and to
reduce overhead requirements. Advancing the study
of topological defects and their corresponding symme-
tries gives us new tools to improve architectures for
topological quantum computation and, moreover, ex-
tends our understanding of the underlying topological
phases that support them.
The color code [28] is a particularly interesting
model due to both its symmetries and its ability to
realize fault-tolerant quantum computational opera-
tions. The importance of the use of different bound-
aries to implement the Clifford gates fault-tolerantly
using transversal rotations has been recognized since
its inception [28]. Further work on the computational
power of color codes and their higher-dimensional
counterparts have also been studied directly with var-
ious boundary configurations in Refs. [2932] and
in the guise of the unfolded surface code model in
Ref. [33, 34]. Remarkably, these models are known
to saturate bounds on the computational power of lo-
cal models using local gates [3538]. It has also been
shown that the boundaries of the color code can be
exploited to perform Clifford gates using code defor-
mation [39, 40] either by the braiding of punctures [41]
or by use of lattice surgery [42] in Refs. [27, 43]. Fur-
ther to this, certain color code twist defects [17, 44],
and their application to quantum computation [17, 27]
have also been studied.
At a more fundamental level, a study in Ref. [45]
examines the symmetries of the color code model to
identify continuous domain walls; the fundamental
objects that are used to generate twist defects. Be-
yond this, the symmetries and twist defects with the
color code model have been expressed using a tensor
network formalism in Ref. [46]. This is complemented
by Ref. [47] where a general study of twist defects and
domain walls using the framework of tensor networks
is given.
A recent surge in activity [4854] has shown that we
may be on the verge of experimental demonstrations
of the fault-tolerant components that will make up a
scalable quantum computer. Moreover, with its afore-
mentioned versatility for performing logical gates, to-
gether with its high threshold error rates [55] and
its relatively economical resource demands [56], the
color code is regarded among the leading candidate
architectures for the realization of topological quan-
tum computation. As such, we have recently seen ex-
perimental proposals to realize color code models with
a number of physical implementations and platforms.
In particular Majorana devices [5762] may be suit-
able as they allow for the measurement of relatively
high-weight stabilizers.
The color code model evidently has a very rich
structure [45, 46] that we may be able to exploit fur-
ther to discover better and more resource efficient pro-
tocols to perform quantum-computational operations
in a fault-tolerant manner. It is the primary goal
of the present manuscript to systematically explore
the different objects that can be realized with mod-
ifications to the color code lattice through a study
of its underlying topological excitations. We discover
three new types of boundary, together with lattice re-
alizations of all 72 different twist defects that can be
realized with the color code model. We catalog all
of these objects and quantify their properties for use
with topological quantum computation.
We go on to find a number of new applications of
the boundaries and twist defects of the color code for
quantum information processing. For example, we
argue that the new boundaries we discover, which we
refer to as Pauli boundaries, may be of practical ben-
efit when performing topological quantum computa-
tion. Specifically, we find that, like with the surface
code model [40], we can move punctures with Pauli
boundaries in the color code model using single-qubit
measurements. In contrast, to the best of our knowl-
edge, we know of no way that we can move punctures
with the known color code boundaries without the use
of two-qubit Bell-type measurements [41]. We addi-
tionally find a new four-qubit error detecting code as
a small instance of a new variant of the color code
that has Pauli boundaries, one that could well serve
as a blueprint for an experimental realization of quan-
tum error correcting protocols, reminiscent of the ones
presented in Refs. [32, 50].
Beyond these examples, we find that our study of
twist defects reveals new local codes with high encod-
ing rates. It is known that two-dimensional commut-
ing projector codes obey the bound kd
2
cn [63],
where k is the number of encoded qubits, d is the
code distance, n is the number of physical qubits in
the code and c is a constant to maximize. By consid-
ering color codes where we place a twist defect in the
center of the lattice, we find a code where c converges
towards 4 as the rotational symmetry of the code
increases. In particular we present two codes with
bounded stabilizer weights w, for w 8 we find a code
with c = 4, and for w 6 a code with c = 8/3 2.7.
In contrast, the twisted surface code [22] has a value
c = 4/3 and the standard color code has a value c = 2
(c = 4/3) when w 8 (w 6). The standard sur-
Accepted in Quantum 2018-10-09, click title to verify 2
face code has a value c = 1 for a lattice that is ap-
propriately oriented [20, 64], or when punctures are
arranged in a suitable pattern [65]. With our simple
example we see that there may be many other benefits
to be gleaned by exploring all of the twists available
with the color code. We present our new family of
stellated color codes, as well as our new four-qubit
error-detecting code, in self-contained sections that
can be understood with only prerequisite knowledge
of the stabilizer formalism for the reader interested
only in the quantum information applications of our
work.
Finally, we study how the new objects we introduce
to the color code lattice can be represented on multi-
ple copies of other topological phases through unfold-
ing [26, 33, 46, 6668]. In particular, in addition to
representing the new objects we discover on two copies
of the surface code, we also explore a lesser known [69]
connection between the color code and two copies of
the three-fermion model [70]. We find this connection
to give a very natural presentation of the underlying
structure of the color code model. Developing the
connections between the color code model and copies
of other, simpler phases may help to discover new and
improved schemes for topological quantum computa-
tion. Examples of which have been demonstrated, for
instance in Refs. [26, 71]. Similar mappings have also
provided new ways of writing decoding algorithms for
the color code [66].
The remainder of the manuscript is structured as
follows. In Sec. 2 we review the general theory of
anyon models and in Sec. 3 we review the stabilizer
formalism, the color code and the low-energy excita-
tions of this model. In Sec. 4 we examine the theory
of boundaries in abelian topological phases and dis-
cover new boundaries with the color code. Using these
new boundaries we discuss improved methods of de-
forming color code punctures and a new four-qubit
error-detecting code in Sec. 5. In Sec. 6 we present
the symmetries of the color code model and introduce
domain walls to the lattice that map excitations non-
trivially while preserving the basic data of the parti-
cle model. Then, in Sec. 7 we extend the discussion
of domain walls by terminating lattice dislocations to
introduce the twists of the color code. Through the
study of twists we find new two-dimensional codes
with high encoding rates, discussed in Sec. 8. Lastly,
in Sec. 9 we study the connection between the color
code and two copies of the surface code model to show
how color code defects are represented in this simpler
model. We finally conclude by proposing new direc-
tions of research that make use of the toolbox we have
elucidated in Sec. 10.
We also give supplemental information on the color
code in the Appendices. In App. A we discuss the six
fermionic excitations of the color code, and in App. B
we discuss the connection between the color code and
two copies of the three-fermion model. In App. C
we discuss the dual color-code lattice, and consider
systematically introducing certain twists to the model
with Pachner moves. Finally, in App. D we present
a new family of surface codes, namely, the stellated
surface codes, which are related to the new color codes
we present.
2 Anyon models
We require notation to describe the anyonic low-
energy excitations of the model of interest without
explicitly referring to the underlying system. Here
we review topological quantum field theory which de-
scribes anyonic particles while remaining agnostic of
microscopic details. For a more detailed and complete
discussion of the theory of anyons see Refs. [46, 15]
or the Appendix E of Ref. [10].
The low-energy excitations of a topological phase of
matter are described by an anyon model. An anyon
model is a set of excitations together with additional
data that describes interactions between these exci-
tations. We write a set of labels a, b, c, ··· C to
denote the distinct charges of the system. Included
in all anyon models is the vacuum charge, 1 C. The
vacuum is the label that indicates ‘no charge’.
Fusion describes the combination of two particles
a, b C. The fusion product of two particles a and b
is denoted with the binary operation a×b. The fusion
rules of an anyon model are described by the fusion
multiplicities N
c
a,b
, such that
a × b =
X
c
N
c
a,b
c, (1)
where c C. Each anyon a has a unique antiparticle
labeled
a which can fuse with it to the vacuum charge,
N
1
a,a
= 1.
Anyons have a Hilbert space associated with them
known as a fusion space. It is possible for the di-
mension of the fusion space to increase as anyons are
added to the system. The factor with which the fusion
space increases for an anyon a is called its quantum
dimension, denoted d
a
1, which is determined by
the equation
d
2
a
=
X
c∈C
d
c
N
a
a,c
, (2)
where d
1
= 1. Anyon a is abelian if d
a
= 1 and is
non-abelian if d
a
> 1. An anyon model is abelian if it
contains only abelian anyons. In fact, for all abelian
anyon models, fusion outcomes are deterministic such
that for each a, b there exists a unique label c such
that N
c
a,b
= 1, and N
d
a,b
= 0 for all d 6= c.
Lastly, we address the exchange and braid statistics
of anyons. Exchanging two anyons of charge a results
in a phase θ
a
, called the topological spin. Diagram-
Accepted in Quantum 2018-10-09, click title to verify 3
matically we express this with the equation
θ
a
a a
=
a a
, (3)
where the lines indicate the trajectories of particles.
Note that θ
a
= θ
a
. We call a bosonic if it exchanges
trivially with itself, i.e. θ
a
= +1, and fermionic if
θ
a
= 1.
Braiding two abelian anyons a and b clockwise re-
sults in a phase M
a,b
, called the monodromy. We can
similarly express this operation with a diagrammatic
equation
M
a,b
a b
a b
=
a b
a b
. (4)
This equation can be transformed by adding a braid
to the two sides of Eq. (4) to obtain
M
a,b
a b
=
a b
, (5)
an expression which will provide useful to calculate
the monodromy of anyons in the considered model. If
M
a,b
= +1, we say a and b braid trivially.
Lastly, we can derive the data involving anyons if
we know enough about the particles into which we
can decompose it. If an anyon c is the only fusion
product of a and b, as is the case in abelian anyon
models, then its self-exchange statistic is
θ
c
= θ
a
θ
b
M
a,b
, (6)
and the monodromy with another anyon d C is given
by
M
c,d
= M
a,d
M
b,d
. (7)
3 The two-dimensional color code
We next review the two-dimensional color code [28],
and summarize how the microscopic description gives
rise to its corresponding anyon model. Refs. [2931,
56] provide alternative reviews of the model.
3.1 The stabilizer formalism
A useful tool to describe a large class of quan-
tum error-correcting codes and many-body quantum
states is the stabilizer formalism [72]. Stabilizer codes
are specified by their stabilizer group, S. This is
an abelian subgroup of the Pauli group acting on n
qubits. We denote the generators of the Pauli group
X
j
and Z
j
, where indices 1 j n denote individual
spin-half particles, or qubits, of the system.
The stabilizer group defines a code subspace
spanned by basis vectors |ψi. The code subspace, or
Figure 1: (a) the 6.6.6 and (b) the 4.8.8 color code lattice.
Physical qubits live on the vertices. Each plaquette p hosts
two stabilizers, one measuring the Z parity of the qubits on
the boundary of p, and the other measures the X parity.
Examples of stabilizers are shown in panel (a).
code-space for short, is the common +1 eigenspace of
all the stabilizers of the stabilizer group. Specifically,
we have
s |ψi = |ψi, (8)
for all s S. The additional requirement that
6∈ S guarantees that the stabilized subspace is
non-empty. The dimension of the code-space is 2
k
,
where k = nm, is the difference between the number
of physical qubits n and the number of independent
generators of the stabilizer group m. When interpret-
ing the stabilized subspace as a logical code-space, k
is the number of logical qubits it supports.
Pauli operators which commute with the stabilizer
group but are not included in it are logical opera-
tors. More formally, the set of logical operators is
such that L
:
= N(S)\S, where N(S) is the normal-
izer of S. Operators in L preserve the code-space
yet act non-trivially on it. The weight w(P ) of a
Pauli operator is the number of qubits upon which
P acts non-trivially. It allows defining the code dis-
tance, d
:
= min
P ∈L
w(P ), which is commonly used as
a first order approximation to quantify the degree of
protection provided by the code-space.
3.2 The color code lattice and stabilizers
The color code is defined on a trivalent three-colorable
lattice, regular examples of which are shown in Fig. 1.
Specifically, each vertex of the lattice has exactly
three incident edges and each plaquette, indexed p,
can be assigned one of three colors, red, r; green, g;
or blue, b, such that no two plaquettes of the same
color touch.
With the lattice defined we consider a system where
a qubit is placed on each vertex of the lattice. The
stabilizer group of the color code is then generated by
operators associated to the plaquettes of the lattice.
Specifically we have two stabilizer generators,
s
x
p
=
Y
jp
X
j
and s
z
p
=
Y
jp
Z
j
, (9)
for each plaquette p, where the product is taken over
all the qubits that lie on the boundary of p. We show
examples of stabilizers on the lattice in Fig. 1.
Accepted in Quantum 2018-10-09, click title to verify 4
3.3 Color code anyons
We have claimed that the color code model gives
rise to anyonic quasiparticle excitations. However,
we have thus far only introduced the color code as a
code-space of some larger physical Hilbert space using
the stabilizer formalism. In fact, we have also implic-
itly defined a Hamiltonian system whose low-energy
excitations are anyons [2, 28]. The code-space of the
model is also the ground state subspace of the exactly
solvable Hamiltonian H, whose interaction terms are
stabilizer generators, which for the color code is
H =
X
p
s
z
p
X
p
s
x
p
. (10)
Applying Pauli matrices, P , to states in the code-
space yields excited eigenstates of H. Plaquettes
where at least one stabilizer generator is violated by,
i.e., anti-commutes with, P are then interpreted as
quasi-particle excitations, namely, anyons. The type
of anyonic quasi-particle is determined by which sta-
bilizer(s) are violated on a given plaquette, and the
color of the plaquette as we will describe in more de-
tail below. For a detailed discussion on color code
excitations, see Ref. [28] and for an in depth discus-
sion on the connection between the stabilizer code and
commuting Pauli Hamiltonians see Ref. [8]. Given
the connection between the Hamiltonian presented in
Eq. (10), we will continue the discussion in terms of
stabilizer states as we defined them previously.
We can characterize the sixteen anyonic charges of
the color code by the operators that create them. The
operators that create excitations are Pauli operators
with string-like support. Importantly, string opera-
tors commute with all the plaquette operators except
for those at their endpoints. As such, anyonic excita-
tions are created at the endpoints of string operators
where plaquette operators are violated.
To study the sixteen color code anyons, we first fo-
cus on a subset of nine charges. This subset is an over-
complete generating set, in that it can generate any
charge from the color code anyon model through fu-
sion. Even though four generators would also suffice,
we find this generating set particularly convenient to
characterize the the model. The nine charges in this
set can be denoted by two labels which are a Pauli la-
bel, P = x, y, z, and a color label, C = r, g, b. Given
that such excitation is specified by one of three colors
and one of three Pauli labels we arrive at nine anyons
which are labeled as CP . Shortly, it will become con-
venient to arrange the generating charges in a 3×3
grid as in Tab. 1 where color labels are arranged in
columns, and Pauli labels in rows.
The excitations we create on the lattice are deter-
mined by the string-operator used to create the ex-
citation. The color label is determined by the color
of the violated stabilizer at the endpoint of a string
and the Pauli label corresponds to the Pauli opera-
tor which was applied to create the excitation. For
rx gx bx
ry gy by
rz gz bz
Table 1: The nine generating anyons in the color code can be
labeled using a color label and a Pauli label. The three colors,
r, g, and b are give in the columns, the Pauli labels x, y, and
z are associated with rows. This arrangement of anyons will
prove useful in what follows, as it succinctly conveys braid
statistics and fusion rules of the color code anyon model.
instance, in Fig. 2 (a) we show a string operator that
creates two gx excitations at its endpoints. This is
because the string is composed of Pauli-X operators,
and the operator shown in the figure violates the two
green s
Z
p
plaquettes at its endpoints. One can readily
check that the other stabilizers on the lattice commute
with this operator.
3.4 Fusion
We now offer a brief intuition on how the fusion rules
arise from the Hamiltonian of the color code, Eq. (10).
A more detailed discussion is given in Ref. [28]. Fol-
lowing the example shown in Fig. 2 (a) we have al-
ready seen that two anyons of the same type are cre-
ated at the endpoints of the string. In fact, we can
extend the string by applying Pauli-X rotations on
the pairs of qubits on the green edges adjacent to the
plaquettes that support the excitations. Green edges
are those that connect two green plaquettes.
In general, it holds that pairs of excitations of a
Figure 2: The fusion rules for color code anyons. (a) Each
color code anyon is its own antiparticle. Here, as an example,
an operator is shown which annihilates (or creates) a pair of
gx anyons. Corresponding operators can be found for pairs
of all color code anyons. (b) Three anyons sharing the Pauli
label but each with a different color label annihilate. This can
also be read as a fusion (or decay) process: From bottom to
top for example, it corresponds to the fusion of a red and a
green anyon to form a blue anyon. Notice, the Pauli label x
was chosen arbitrarily; the same holds for y or z. (c) Three
anyons sharing the same color label but each with one of
the three Pauli labels can also be annihilated. Again, and
alternative reading is the fusion of an x anyon and a y anyon
to form a z anyon. The red color label can be replaced by
either green or blue.
Accepted in Quantum 2018-10-09, click title to verify 5
given color are created by applying strings of the ap-
propriate Pauli rotations to edges of its corresponding
color [28]. Given that we can separate pairs of exci-
tations of the same type CP by extending a string of
Pauli operators P type via edges of color C, we arrive
at the fusion rule
CP × CP = 1. (11)
Physically, this fusion rule represents
2
charge con-
servation as it shows that anyons of a given type
must be created in pairs. This rule can also be in-
terpreted to say that color code anyons are their own
anti-particles.
We can similarly study string operators acting on
the color code lattice to determine fusion rules be-
tween anyons with different color and Pauli labels. In
Fig. 2 (b) a Pauli-X operator is applied to a single site.
This operator violates stabilizers on its three adjacent
plaquettes, each of which has a different color due to
the three-colorability of the lattice. The excitations
created by this operator can be separated arbitrarily
far from the other excitations with the application of
another appropriately chosen string. We thus arrive
at the fusion rule
C
1
P × C
2
P = C
3
P, (12)
for any Pauli label P and pairwise distinct color la-
bels C
1
, C
2
, C
3
. We thus observe a conservation law
between the color labels of the color code excitations:
Two charges of a common Pauli-type and distinct col-
ors fuse to give the charge of the same Pauli type and
the third color.
Finally, we can also observe a fusion rule between
different Pauli labels. In Fig. 2 (c) we show a string
operator composed of Pauli-X, Pauli-Y and Pauli-Z
operators with three red endpoints where stabilizers
are violated. At two of the endpoints we have a sin-
gle violated s
x
p
(s
z
p
) stabilizer which indicates an rz
(rx) excitation. At the third endpoint both stabilizer
generators are violated, indicating an ry excitation at
this location. Generalizing, we obtain the fusion rule
that connects Pauli-labels of the excitations
CP
1
× CP
2
= CP
3
, (13)
for arbitrary color C and pairwise distinct Pauli labels
P
1
, P
2
, P
3
. The reader may recognize a duality be-
tween the color label and the Pauli label in Eq. (12)
and Eq. (13). Indeed, this duality is a key feature
of the color code anyon model as will become clear
throughout the manuscript.
With the fusion rules we have developed so far we
begin to see the utility of Tab. 1. Indeed, the table
echos Eq. (12) as the fusion product of two distinct
excitations from the same row give the third excita-
tion of the given row. Similarly, the fusion product
of two excitations of different Pauli label from a com-
mon column fuse to give the third excitation from the
given column. This reflects the fusion rule shown in
Eq. (13). More trivially, we have that the fusion prod-
uct of two excitations from the same element of the
table fuse to give the vacuum excitation as in Eq. (11).
With the rows and columns of Tab. 1 now sepa-
rately understood, it remains to consider the fusion
product of pairs of charges that share neither a row
nor a column. As we will show in the following subsec-
tion, the nine excitations in Tab. 1 are in fact bosonic
excitations. Aside from the vacuum particle, the exci-
tations of the color code that are not shown in Tab. 1
are fermionic excitations. All of the remaining exci-
tations can be expressed as the fusion product of two
of elements. In fact, the color code gives rise to six
distinct fermions. For more details on the fermionic
excitations of the color code we refer the reader to
App. A. Moreover, in App. B we show that we can
describe the excitations of the color code using two
non-interacting copies of a model of three fermions.
3.5 Self-exchange and braiding
We referred to color code anyons as either bosonic or
fermionic. To justify these names we now derive their
self-exchange statistics. Consider Fig. 3, where we
compare two sets of string operators which move rz
anyons along closed loops. On the left, two separate
loops are shown. This is the left-hand side of Eq. (6),
only that the bottom-left(right) and top-left(right)
ends of the trajectories are here connected. On the
right-hand side of Fig. 3, a single closed loop of an rz
operator which crosses itself in the middle is depicted.
This is the right-hand side of Eq. (6), again with the
bottom-left(right) end of the trajectories connected
Figure 3: Bosons in the color code have trivial spin θ
a
= +1.
As an example compare string operators which move an rz
anyon along a closed loop. On the left, two circular paths
looping back to their initial positions are shown. Each of
these operators is the product of the z type stabilizers on
the blue and green plaquettes they enclose and hence itself
again a stabilizer. On the right however, the middle section
is changed, such that path crosses itself once in the middle
to form a single larger loop. However, this redrawing of the
path did not change how the operators act, and the left-hand
side is exactly equal to the right-hand side. Further more,
according to Eq. (3), the difference between these operators
gives us the spin of the moved anyon, here θ
rz
= +1. The
same holds for all nine anyons shown in Tab. 1.
Accepted in Quantum 2018-10-09, click title to verify 6
Figure 4: To obtain the braid statistics of the color code
anyons, we study the commutators of string operators. As
an illustrative example, an rx anyon is moved from the left to
the right out of frame. Then, the line along which said anyon
was moved is crossed top to bottom by three other anyons.
Comparing the two different orders in which these operators
are applied reveals the phase obtained by braiding the anyons,
see Eq. (5). In (a) we see that strings moving anyons sharing
the color label (red in our example) exchange trivially. Here,
not a single qubit is acted upon by both strings. In general
the number of qubits supporting two crossing strings of the
same color is even. (b) Crossing strings moving anyons with
the same Pauli label also exchange trivially, as strings moving
them act in the same Pauli basis. (c) shows the third case,
neither the Pauli label nor the color label of the two anyons
associated with the crossing string operators match. This
leads to an odd number of qubits on which the strings act in
anti-commuting bases. In our example one qubit is acted on
first by an X and then by Z. This leads to a phase of 1 if
the anyons are braided.
to the top left(right). Hence, the difference between
the left and right half of the figure corresponds to the
self-exchange statistic of rz. But clearly, the effect of
the shown operators is exactly the same and hence
θ
rz
= +1, rz is a boson. This is true for all nine
non-trivial bosons, as can be seen by changing the
Pauli basis in which the considered operators act, or
by moving the string operators such that they move
anyons of different colors. The self-exchange of the
color code fermions can be derived using Eq. (6) and
Eq. (14) given below together with any decomposi-
tion of fermions into two bosons shown in Eq. (23)
and Tab. 8 in App. A.
Finally, we look at the phase obtained by braiding
pairs of color code excitations. Braiding a bosonic
anyon a with color label C
a
and Pauli label P
a
with
a second boson b with labels C
b
and P
b
results in the
phase
M
a,b
= (1)
δ(C
a
6=C
b
)δ(P
a
6=P
b
)
(14)
where δ(statement) is an indicator function that re-
turns 1 if statement is true and 0 otherwise.
Like for the fusion rules, we find Tab. 1 proves a
useful tool to visualize the braid statistics between
bosonic excitations. This is to say, anyons a and b
which share a row or a column braid trivially, i.e.
M
a,b
= +1. If a and b share neither a row nor a
column however, we have that M
a,b
= 1. To prove
the above, we compare the order in which the cross-
ing operators moving anyons a and b are applied. As
in Eq. (5), the difference reveals the phase M
a,b
. As
can be seen in Fig. 4, pairs of string operators asso-
ciated to anyons with the same color label or Pauli
label exchange trivially, and thus we find M
a,b
= +1.
When both, the color label and the Pauli label is dif-
ferent however, the operators anticommute and we
get M
a,b
= 1. The braiding rules involving fermions
can be derived from the above, using Eq. (7) and any
decomposition of fermions into two bosons shown in
Eq. (23) and Tab. 8 in App. A.
4 Boundaries
Unless we are dealing with a closed 2D manifold, the
spatial extent of a topological phase must end at some
boundary, such as a gapped domain wall with the triv-
ial phase. It is important to gain a better grasp on
the physics at the boundary of a topological phase.
Not only is this interesting from a condensed mat-
ter perspective [13, 14, 7376], but can be exploited
in quantum computation [23, 3941, 77, 78]. In this
section, after reviewing concepts of the boundaries of
abelian topological codes, we introduce lattice real-
izations of three new color code boundaries and show
that this completes the boundaries available to the
model. We finally discuss the prospective applications
of these boundaries for quantum computational tasks
as well as small scale experimental implementations
of quantum error-correcting codes.
4.1 Lagrangian subgroups
Gapped boundaries are characterized by Lagrangian
subgroups. These mathematical objects are subsets
of anyons in the anyon model C which can condense
at the boundary of interest. Anyon condensation is
the process wherein an anyon meets and is absorbed
by the boundary. The reverse process is also possible
where a single excitation is created or ‘emitted’ by a
boundary. Anyons which do not condense are said to
be confined by the boundary. Boundaries that can
condense excitations are important for quantum in-
formation applications as these are topologically pro-
tected degrees of freedom whose state is determined
by the charges they have absorbed.
The subset of anyons that can condense at a bound-
ary is constrained to obey certain conditions that
we will now briefly review. Indeed, it is shown in
Refs. [73, 74] that the subsets of excitations in C that
are absorbed at gapped boundaries correspond to La-
grangian subgroups M. A Lagrangian subgroup ful-
fills three conditions: Firstly, as one might expect,
a Lagrangian subgroup is a group, and thus closed
under fusion. Further, if a particle can condense at
a given boundary, so can its antiparticle. Likewise,
the vacuum particle can be condensed at any gapped
Accepted in Quantum 2018-10-09, click title to verify 7
Figure 5: The conditions on Lagrangian subgroups picto-
rially summarized. The gray line represents the boundary
between the topological phase at the bottom and the trivial
phase above. (I) If two anyons a and b can condense, so
can the outcome of their fusion c = a × b. And the emis-
sion of an anyon a can be read as the condensation of its
antiparticle a. (II.a) Anyons which can condense must have
bosonic self-exchange statistics. The shown paths are stabi-
lizers, hence the left and the right situation are equal, which
implies trivial self-exchange for condensible anyons. (II.b)
Two anyons which can condense must braid trivially. Again,
the depicted paths are stabilizers and must hence commute,
which implies the trivial braiding of the anyons. (III) An
anyon d which can not condense must exchange non-trivially
with at least one anyon which can condense. This means the
set of condensible anyons is maximal. Sets of anyons which
fulfill these conditions are called Lagrangian subgroups of the
anyon model.
boundary. Secondly, all of the anyons within the sub-
group are bosons and braid trivially with all other
anyons in the subgroup. The third condition is that
M is maximal in the sense that no additional anyon
can be added to the subgroup such that the other
conditions are fulfilled. In the following, we revisit
the argument given in Sec. IV of Ref. [73].
First, we start by showing that M is closed under
fusion, a M and b M = a × b M. a × b
can condense by decaying into a and b, which can
then condense separately. These is pictorially shown
on the left of Fig. 5 (I). Before showing that M does
actually form a group, we first turn out attention to
the other claims.
The second requirement of the boundary can be
understood by considering string operators s
a
. These
operators act such that they pull an anyon a M
from the boundary, transport it along a trivial tra-
jectory through the bulk and condense it back at the
same boundary such that s
a
leaves the charge at the
boundary unchanged. Given that the string opera-
tors follow a trivial trajectory and do not change the
net charge of the boundary, these operators do not
affect the logical state of the system. Furthermore,
given that such operators do not create excitations,
we can regard these operators as stabilizers such that
s
a
|ψi = |ψi for all ground state vectors |ψi.
Using such operators, we can argue that all anyons
which can condense must be bosons as follows: Con-
sider the four operators shown in Fig. 5 (II.a). Com-
paring s
a
, s
0
a
to ˜s
a
, ˜s
0
a
, we see the endpoints have
been exchanged. Hence, the left and the right in
Fig. 5 (II.a) differ, according to Eq. (3), by a phase θ
a
given by the spin of charge a. But since these four op-
erators are all stabilizers, they act trivially and hence
θ
a
= +1. This holds for all a M, and thus all
condensible anyons are bosons.
To show the trivial braiding between condensible
bosons, consider operators s
a
and s
b
, which move dif-
ferent anyons a, b M. Their trajectories intersect
in a single point, as depicted in Fig. 5 (II.b). Note
how the order in which they are applied is changed on
the right compared to the left. Again, since s
a
and
s
b
are stabilizers, the left and the right situation are
exactly equivalent. And according to Eq. (5), com-
paring the two cases reveals the phase M
a,b
obtained
when braiding a and b. We conclude that anyons in a
Lagrangian subgroup braid trivially with each other
such that M
a,b
= 1 a, b M.
The third claim about the maximality of these
subgroups can be intuitively understood by requir-
ing braiding non-degeneracy. This means every any-
onic excitation can be measured by at least one other
anyon through braiding, the only exception being
charges in M. See the left half of Fig. 5 (III).
This braiding non-degeneracy also holds close to a
boundary, if for example anyon d is condensed in
or confined at the boundary. It can be stated as,
d 6∈ M a M s.t. M
a,d
6= +1. In this case, the
only anyons which can perform a full braid around d
are anyons which are able to condense, a M. This
can be seen on the right half of Fig. 5 (III). Thus, an
anyon close to the boundary can either be seen by at
least one anyon a M, or is itself in M.
Now, we can address that M indeed forms a group.
We already discussed that it is closed under fusion.
From the maximality we can conclude that the triv-
ial charge 1 is also in M, as it braids trivially with
all anyons. To see that for each charge a M its
antiparticle a M holds, we can for example use
Eq. (6). Choosing b = a and c = 1 = a × a and using
θ
a
= θ
a
, we find M
a,a
= 1.
4.2 The boundaries of the color code
To find all boundaries of the color code, we explore
the Lagrangian subgroups of the excitations of the
model. Given that Lagrangian subgroups consist only
of bosonic excitations, we concentrate only on the
bosonic excitations of the color code. We find Tab. 1
offers a particularly convenient representation of the
bosons of the color code model for identifying La-
grangian subgroups. Precisely, subsets consisting of
the trivial anyon and three bosons from a row or col-
umn constitute the Lagrangian subgroups of the color
code.
The fact that these sets of anyons are closed under
fusion follows readily from the discussion in Sec. 3.4.
Accepted in Quantum 2018-10-09, click title to verify 8
The braiding requirement is met as well, as discussed
in Sec. 3.5. There we argued that bosons in the ta-
ble braid trivially with each other if and only if they
share a row or a column. It is readily checked that the
subgroups of the rows and columns of Tab. 1 are max-
imal. We consider trying to add additional anyons to
the Lagrangian subgroups of a row or column. The
addition of a fermion would obviously violate the self-
exchange criterion of the subgroup. And trying to
add another boson only leaves bosons from a differ-
ent column or row. But they braid non-trivially with
two of the elements of the Lagrangian subgroup. We
conclude that, together with the trivial charge 1, the
three rows and three columns of Tab. 1 correspond
exactly to the six possible Lagrangian subgroups of
the color code anyon model.
Alternatively, if the anyon model in question can be
decomposed into two layers of a simpler anyon model,
boundaries can be found that way. Namely, bound-
aries correspond to folds between the two layers of
the simpler model with its domain walls, see Sec. 6
and Ref. [74]. The famous equivalence of the color
code and two copies of the toric code is explored in
Sec. 9. A second equivalence between the color code
and two copies of the three-fermion model is discussed
in App. B.
We next go on to find lattice representations of
these boundaries where the respective label subsets
condense.
4.3 Lattice representations of the color code
boundaries
We have identified six Lagrangian subgroups of the
color code anyon model. However, to the best of our
knowledge, only three of these subgroups have been
explored previously. In what follows we review the
boundaries that have already been studied and we
find and discuss the new boundaries that correspond
to the other three Lagrangian subgroups of the color
code model.
The Lagrangian subgroups that have been consid-
ered explicitly form the columns of Tab. 1. Each of
the columns share a common color label, as such, we
find that we can obtain boundaries that absorb bosons
with a common color label. We call these the red,
green and blue color boundaries. These color bound-
aries were presented in the original work introducing
the color code [28]. We show an example of a blue
color boundary in Fig. 6 (a). We remark that the
stabilizer generators at the boundary of the model re-
spect the same rules presented in Sec. 3 where each
plaquette p supports two stabilizers s
x
p
=
Q
jp
X
j
and s
z
p
=
Q
jp
Z
j
. The figure also shows string op-
erators that create single blue excitations from the
boundary.
In addition to the three Lagrangian subgroups that
correspond to the columns of Tab. 1, we also find three
Figure 6: The color code boundaries and corners between
them. (a) shows a blue boundary as an example of a color
boundary, as well as the condensation of the three blue
bosons. (b) An example of a Pauli boundary. Here the z
boundary is shown and the condensation of the three bosons
with Pauli label z. (c) Corners between two color bound-
aries appear if a qubit is only part of a single plaquette of
the third color. (d) A corner between two Pauli boundaries
is characterized by a qubit which is only part of two striped
boundary plaquettes. (e) Corners between Pauli and color
boundaries are also possible, here shown a z Pauli bound-
ary meeting a blue color boundary. (f) shows the stabilizers.
As usual, hexagons carry associated X- and Z-type weight-
six stabilizer. The color boundaries shown have quadrilateral
plaquettes which host weight-four X-type and Z-type sta-
bilizers. The striped triangular plaquettes featured in Pauli
boundaries are only stabilized by a single stabilizer. We indi-
cate the basis in which the triangular plaquettes are measured
by the direction of the stripes.
Lagrangian subgroups that correspond to the rows of
the table. We can therefore expect the existence of
boundaries that absorb all of the bosonic excitations
with a common Pauli label, independent of their color
label. These boundaries will be referred to as the
x, y and z Pauli boundaries. A similar conjecture
has been made in Ref. [33] by consideration of the
unfolded color code, i.e., two copies of the toric code.
We discuss this in more detail in Sec. 9.
In Fig. 6 (b) we show a lattice realization of the
z Pauli boundary, which condenses bosons with the
Pauli label P = z. At the boundary we show weight-
three plaquettes. In principle we could choose any
number of stabilizers acting at these plaquettes p,
s
x
p
=
Q
jp
X
j
, s
y
p
=
Q
jp
Y
j
or s
z
p
=
Q
jp
Z
j
.
But since we require all stabilizers to commute, we see
that we can only choose one stabilizer per plaquette.
This is because s
x
p
, s
y
p
and s
z
p
mutually anti commute
as they each share three qubits. Similarly, two of these
boundary stabilizers on adjacent plaquettes must be
of a common Pauli type in order to commute with one
another. Adjacent boundary stabilizers have common
support only on one single qubit. In the figure we ar-
bitrarily choose s
z
p
stabilizers, and we show examples
of differently colored excitations with a z Pauli label
that are emitted from the boundary. To the best of
our knowledge, the Pauli boundary has not been ex-
plicitly represented on the color code lattice before.
The following table summarizes the discussion
Accepted in Quantum 2018-10-09, click title to verify 9
rx gx bx
ry gy by
rz gz bz
Pauli
boundaries
color
boundaries
Table 2: To visualize the color code boundaries, we can
return to the arrangement of the bosonic anyons presented
in Tab. 1. The nine non-trivial bosonic anyons in the color
code are arranged such that anyons in the same column or
in the same row exchange trivially. Each row and column
corresponds to a boundary. Anyons which can be condensed
at a common color (Pauli) boundary share their color (Pauli)
label.
given above. It explicitly shows that the columns
of the table represent the Lagrangian subgroups that
give rise to color boundaries and the rows repre-
sent the Lagrangian subgroups that give rise to Pauli
boundaries.
Finally, we present the interfaces between the dif-
ferent boundary types. General interfaces between
boundaries have attracted some interest in the con-
densed matter community [79]. For quantum in-
formation applications, points where boundary type
changes are referred to as corners [23].
We show different corners in Fig. 6 (c), (d) and (e).
The first interface separates two color boundaries, see
Fig. 6 (c). This interface has already been utilized in
Ref. [28] to find a two-dimensional color code capa-
ble of transversally performing all Clifford operations.
See also Refs. [29, 30] for a more detailed discussion
on similar boundary interfaces. In Fig. 6 (d) we show
an interface between two different Pauli boundaries.
We show the corner between an x Pauli boundary
and a z Pauli boundary which generalizes immediately
to all Pauli-Pauli corners. At the top-right corner of
the figure we show two boundary plaquettes that are
commonly supported on two qubits where two Pauli
boundaries can meet such that the boundary stabiliz-
ers commute. Such corners can be used to construct
codes with Pauli boundaries, as discussed in Sec. 5.2.
Finally, in Fig. 6 (e), we show a corner where a Pauli
boundary meets a color boundary. Once again, it is
readily checked that the stabilizers represented in the
figure commute. This example can also be generalized
easily to obtain all nine color/Pauli corners.
5 Small color codes and fault-tolerant
quantum computation
In addition to giving us a better understanding of
the fundamental physics that underlies the color code
phase, a second benefit of systematically laying out all
of the different features of the color code is that it al-
lows us to discover new modes of fault-tolerant quan-
tum computation with the model. In what follows we
give two such examples of applications of Pauli bound-
aries. Specifically, we show that the new boundaries
we have introduced give rise to better methods of per-
forming code deformations, and we also find a new ex-
ample of a four-qubit quantum error correcting code
which can be interpreted as the smallest instance of a
color code with Pauli boundaries.
5.1 Code deformation
We first consider code deformations, and argue that
braiding punctures with Pauli-boundaries to perform
fault-tolerant entangling gates will be advantageous
over equivalent operations using punctures with col-
ored boundaries as have been considered previously
in the literature [41].
One of the leading paradigms for performing logi-
cal gates is through the braiding of punctures in some
topological quantum error-correcting code. This, tra-
ditionally, put the color code at a disadvantage com-
pared to the toric code. In the toric code, single
qubit measurements are sufficient to create and move
punctures. In contrast, the creation and movement of
punctures featuring the color boundaries of the color
code require two qubit Bell measurements [41]. We
find that we can move punctures with Pauli bound-
aries using only single-qubit measurements. This is
readily checked because we can measure a patch of
individual qubits of the color code lattice using single-
qubit Pauli measurements in a common Pauli basis to
introduce or increase the size of a puncture. In doing
so, we remain in a known eigenstate of the stabiliz-
ers at the boundary of the new puncture that has
been produced by the measurements. We leave a de-
tailed derivation of this argument as an exercise to the
reader. This property of Pauli boundaries is advan-
tageous for practical implementations for color code
quantum computation as single-qubit measurements
are invariably easier to perform than two-qubit par-
ity measurements and are therefore less likely to in-
troduce noise during deformations. This feature may
also play an important role for measurement based
fault-tolerant quantum computation [39] using color
codes.
5.2 A four-qubit color code
Our extended toolbox has also enabled us to find a
new four-qubit color code which we expect will be of
interest for near-term experimental realization with
small-scale quantum devices. They complement what
has been seen in small-scale trapped ion quantum
computers [50, 54] and relates to insights into the
smallest interesting color code [32].
To describe the features of our new code, we first
present the family of triangular color codes with Pauli
boundaries. These codes feature an x Pauli bound-
ary, a y Pauli boundary and a z Pauli boundary. The
Accepted in Quantum 2018-10-09, click title to verify 10
Figure 7: Examples of small triangular color codes with Pauli
boundaries. The stabilizers acting on bulk plaquettes are the
usual. On the boundary plaquettes, indicated by the stripes,
only one stabilizer is measured. These boundary stabilizers
are shown in the top left. (a) shows the smallest member
of this family, a J4, 1, 2K code. It consists of four physical
qubits, three weight-three stabilizers and hosts one logical
qubit. Its logical operators are, for instance, X
L
= X
3
X
4
,
Y
L
= Y
2
Y
4
and Z
L
= Z
1
Z
4
. (b) and (c) show the next
smallest codes with Pauli boundaries, consisting of nine and
sixteen qubits respectively. In (c) examples of two logical
operators, X
L
and Z
L
, are shown.
three smallest members from this family of codes are
shown in Fig. 7. Bulk plaquettes feature the two reg-
ular stabilizers given in Eq. (9), the striped boundary
stabilizers host a single weight-three stabilizer each,
the basis in which it acts is indicated by the direction
of its stripe pattern. These codes encode a single logi-
cal qubit each. Examples of their logical operators are
shown in Fig. 7 (c). Remarkably, the smallest instance
of this code, shown in Fig. 7 (a), is a new example of
a J4, 1, 2K code with only weight-three stabilizers.
6 Bulk domain walls
In this section we introduce transparent domain walls,
or invertible domain walls [14], to the model. These
domain walls change the charge labels of anyons which
cross them in a non-trivial way, such that fusion,
braiding and self-exchange of the anyons is conserved.
Specifically, we discuss closed domain walls here; open
domain walls and in particular their terminal points
are considered in Sec. 7. As we are interested in ap-
plying the theory to the color code, we restrict the dis-
cussion to abelian anyon models and refer the readers
to Refs. [15, 80] for the general case.
We first provide a comprehensive study of trans-
parent domain walls for abelian anyon models in Sub-
sec. 6.1. Then we apply this theory in our concrete ex-
ample of the color code, whose symmetries we study in
Subsec. 6.2. Lattice representations of domain walls
are given in Subsec. 6.3.
6.1 Symmetries in abelian anyon models
Transparent domain walls change the labels of anyons
crossing them according to a map. This map is an
automorphism, ϕ : C C which maps the anyon
model onto itself. Acting on an anyon a C it returns
Figure 8: The consistency conditions on a domain wall can
be derived by studying contractible loops crossing the domain
wall. From left to right the figures correspond to Eqs. (15-
17).
again an anyon ϕ(a) = a
0
C. We require these
automorphisms to leave the basic data of the anyon
model, i.e its fusion, self-exchange and braid statistics,
invariant. See Ref. [15]. Explicitly, this means
ϕ
N
c
a,b
= N
c
0
a
0
,b
0
!
= N
c
a,b
, (15)
ϕ (θ
a
) = θ
a
0
!
= θ
a
, (16)
ϕ (M
a,b
) = M
a
0
,b
0
!
= M
a,b
. (17)
These conditions can be obtained by requiring that
anyons moved along contractible loops crossing the
domain wall constitute stabilizers and hence leave the
state invariant. Considering a branching loop, like
on the left of Fig. 8, we can follow Eq. (15). If an
anyon c on the left of a domain wall decays into a
and b which are then moved over the domain wall,
then on the right a
0
and b
0
must fuse to c
0
. And us-
ing that stabilizers commute with each other, we can
deduce Eq. (16) from the paths shown in the middle
of Fig. 8. We observe that any phase θ
a
picked up by
the self-exchange on the left must be canceled by the
phase θ
a
0
obtained by the reversed exchange on the
right. Lastly, to get Eq. (17), consider the two un-
linked loops on the right of Fig. 8. Again, any phase
M
a,b
picked up by the braid performed on the left
must be canceled out by the phase M
a
0
,b
0
which is
obtain by the reverse braid on the right.
The automorphisms fulfilling these conditions for
an anyon model C are called the symmetries of C.
And each transparent domain wall can be described
Figure 9: Two closed domain wall, drawn in purple. The
arrows are to indicate the directionality of the domain wall.
The domain wall drawn as a solid line is described by the
automorphism ϕ, the dashed one by ϕ
0
. Where the domain
walls are fused, a new domain wall is obtained which is de-
scribed by ϕ
0
ϕ. Note that domain walls have an orientation
and changing it corresponds to taking the inverse ϕ in the
group of automorphisms.
Accepted in Quantum 2018-10-09, click title to verify 11
and labeled by one such automorphism ϕ. The set of
symmetries of an anyon model forms a group [15]. The
group identity is the trivial map, corresponding to the
absence of a domain wall, and the group multiplica-
tion is the composition of maps. Thus, composition
maps also fulfill the conditions and are therefore also
symmetries of the anyon model. Physically, compos-
ing maps corresponds to multiple domain walls being
combined or fused. See Fig. 9 for a schematic of how
domain walls change anyons and how the fusion of
domain walls corresponds to the composition map.
6.2 Symmetries in the color code
To classify the transparent domain walls in the color
code, it proves once again useful to go back to the 3×3
grid of bosonic color code anyons, shown in Tab. 1. As
mentioned earlier, fusion of two different anyons in a
common row or column results in the third anyon in
said row or column. As such we find that the fusion
rules are preserved by all transformations on the grid
that respect that all elements of a given row or column
remain together in any other row or column. Likewise,
we know that excitations braid trivially if and only if
they share a column or row of the grid. Like fusion
then, the mutual braid statistics of this generating
set of excitations shown in the table are respected
provided all excitations that share a common row or
column remain in rows and columns with the same
excitations.
The above discussion implies that the symmetries of
the generating set of bosonic excitations, and thus the
excitations of the color code in general, are obtained
by finding all the transformations that respect the
rows and columns of the grid. Indeed, these are row
permutations, column permutations, and the trans-
position of the grid. Given there are three rows and
three columns, there are six different permutations of
the rows, and of the columns, and finally one trans-
pose. We thus arrive at 6 ·6 ·2 = 72 symmetries of the
color code. This agrees with the exhaustive numerical
search given in Ref. [45].
Further insight on the symmetry group can be
gained using the following decomposition: (S
3
× S
3
)n
2
. One S
3
comes from permutations of the columns,
i.e. colors, the second from the rows, i.e. the Pauli
labels. Since permuting rows and columns commutes,
the corresponding group is given by the direct prod-
uct, S
3
× S
3
. Finally, the
2
matrix transposition
exchanging color and Pauli labels acts via the semi
direct product (n). As a compact notation we can use
the wreath product (o) to write the symmetry group
as S
3
o
2
. This decomposition follows directly from
interpreting the color code anyon model as two copies
of the three-fermion model, see App. B for more de-
tails.
To label symmetries which exchange two colors, we
use a capital letter of the third, unchanged color. For
example, the symmetry which exchanges the green
and the blue color label while leaving the red label,
we denote as R. Equivalently, G and B respectively
leave green and blue color labels invariant while ex-
changing the other two. Similarly, the exchanges of
the Pauli labels we denote as X, Y or Z, for symme-
tries which leave the x, y or z Pauli label invariant.
Finally, we use D for the duality symmetry between
color and Pauli labels, corresponding to the transpo-
sition of the 3×3 anyon grid, shown in Tab. 3. All of
the symmetries mentioned so far are of order 2, i.e.
applying them twice leads to the trivial symmetry.
Combining two of the above order 2 color symme-
tries creates one of the two cyclic color permutations.
Thus we can label them RB and BR, for example.
Conjugation of one color exchange with a second re-
sults in the third, e.g. RGR = B. The analogous
statement holds for the permutations of Pauli labels;
XZ and ZX are the cyclic Pauli label permutations
and we can check that XY X = Z. Conjugating
a color exchange with D creates a symmetry which
solely exchanges Pauli labels, for example DBD = Z,
or conversely DZD = B.
x
y
z
r g b
B
Y
D
Table 3: The effect of the three generating symmetries on
the anyons. B is the blue invariant color permutation which
exchanges the red and green color label. Y the y-invariant
Pauli permutation which exchanges x and z Pauli labels. And
D the duality symmetry which exchanges color and Pauli
labels as shown.
To generate all 72 symmetries, three are already
sufficient. A minimal choice might be, for instance,
X, Y and D, two different row permutations and the
duality. However, here, we choose B, Y and D, a row
exchange, a column exchange and the transposition,
as shown in Tab. 3. This is to demonstrate all three
different types of domain walls in what follows.
6.3 Lattice representation of domain walls
Now, we give lattice representations of color code do-
main walls which correspond to the generating sym-
metries above. In particular, these are the B, Y and
D symmetries. They can generate all 72 color code
domain walls, as discussed in Sec. 6.2. And they
also constitute examples for three different classes of
twists, one only acting on the color label, one on
the Pauli label and duality transformation which ex-
changes the two. To obtain these closed domain walls,
we transform the stabilizers along a closed path. This
provides an apparent physicality to the location of the
Accepted in Quantum 2018-10-09, click title to verify 12
Figure 10: A 6.6.6 color code lattice with a blue-invariant
domain wall B, marked by the purple line. Along the domain
wall a fourth color, yellow, is introduced. Two commuting
stabilizers are measured on each plaquette, one x-type and
one z-type. Anyons crossing the domain wall keep their Pauli
label, but red and green color labels are exchanged. Note
how we omitted the arrow indicating the orientation of the
domain walls, as it corresponds to an order two symmetry
B = B.
domain walls, which although artificial provides a way
to later emphasize their terminal points whose local-
ity is physical. These terminal points are called twist
defects and are the focus of Sec. 7. We emphasize
that neither of the transformations can be achieved
by means of a local unitary with support only where
the domain wall will come to lie.
An example of a domain wall changing the color
label is shown in Fig. 10. It exchanges solely the red
and the green color label of anyons crossing it. It cor-
responds to the symmetry B, shown in Tab. 3. Along
the domain wall, each plaquette hosts two stabiliz-
ers, an x- and a z-type. Notice how in this example
a fourth color is used to color the faces. The brak-
ing of three-colorability is in fact a general feature of
color permuting domain walls. In App. C we show
how color permuting domain walls can be introduced
by applying a series of Pachner moves on the dual
lattice. Using the Pachner move technique also al-
lows us to generate domain walls which permute the
three color labels in a cyclic fashion, see Fig. 36 (a)
and Fig. 37 (a). The two other color exchanging
domain walls, R and G can be drawn equivalently.
In Refs. [17, 44, 81], similar domain walls have been
demonstrated for equivalent models.
Next, we turn our attention to domain walls per-
muting Pauli labels. They can be constructed with-
out changing the underlying lattice. Instead, just the
basis in which stabilizers act along the domain wall
are measured is altered. Such a domain wall is rep-
resented by a violet line in Fig. 11. To generate a y-
invariant domain wall Y , the two stabilizers for each
plaquette lying on top of the domain are changed.
The first one acts in the x-basis on qubits on one side
of the domain wall, and in the z-basis on qubits on the
other side. The second one acts in the z-basis on one
side and the x-basis on the other side. This is equiv-
Figure 11: A 6.6.6 color code with a y-invariant Pauli domain
wall Y represented by a purple line. Along the line, every
plaquette hosts two stabilizers, they are shown in the bottom
left. They can be obtained by conjugating the basis in which
the stabilizers act by a Hadamard on one side of the domain
wall.
alent to conjugating the basis in which the domain
wall stabilizers act with a Hadamard on one side of
the domain wall, as discussed in Ref. [45]. These also
correspond to locality preserving gates in the topo-
logical quantum field theory, which were studied in
Ref. [82]. Domain walls which permute the Pauli label
of anyons in a cyclic manner can also be constructed
in a similar fashion by conjugating them partially by
a unitary which cyclically permutes x, y and z.
Lastly, we give a lattice representation of the do-
main wall which transposes the 3×3 arrangement of
anyons. It corresponds to the symmetry D and we re-
fer to it as the domino domain wall, since it features
stabilizers that resemble domino bricks, see Fig. 12.
Such a domain wall can be constructed by pulling a
zigzag line of edges apart which duplicates all of the
qubits lying along the line. In the created space we
color squares of the lattice which define rectangular
stabilizers that each act on six qubits which are sup-
ported on two adjacent squares that make a rectangle.
Each of these new stabilizers has two colors, one on
Figure 12: An example of an domino domain wall D. The
dislocation line hosts rectangular stabilizers supporting six
qubits. Like domino bricks, they are colored in a different
color on either half. Their colors dictate the basis in which
they act, shown in the bottom left.
Accepted in Quantum 2018-10-09, click title to verify 13
either side, resembling domino bricks. These colors
are determined by the colors of the neighboring pla-
quettes. To each of these domino-brick shaped cells
we assign one weight-six stabilizer, where the colors
of the brick determine which operator acts on rectan-
gle. Stabilizers colored green and blue are measured
in the x-basis, red and blue in the y-basis, and red and
green in the z-basis. This results in the symmetry D
which transposes the 3×3 grid. This domain wall it
leaves the bosons rx, gy and bz that cross it invariant.
See Fig. 12 for a lattice representation of the domain
wall and the effect it has on anyons. Changing what
pair of colors is associated with which Pauli basis, six
different domain walls can be obtained. To the best
of our knowledge, this is the first time a lattice rep-
resentation of this domain wall has appeared in the
literature.
7 Twists
In the previous section, we considered continuous
transparent domain walls that change the charge la-
bels of an abelian anyon model while preserving the
fusion rules and braid statistics between the charges.
In what follows we build on our discussion by ter-
minating the domain walls. Again, we restrict our
discussion to abelian anyon models. At the terminal
points of these domain walls, we realize twist defects
or simply twists [11, 12, 14, 15]. These remarkable
point-like objects have many properties in common
with non-abelian anyons. This is especially intriguing
considering that we started with an abelian theory. As
such they are of particular interest for applications in
quantum information processing.
We begin by introducing twist defects in Sub-
sec. 7.1. In Subsec. 7.2 the fusion of twists is dis-
cussed. We then study the twist defects of the color
code and discuss how they interact with its anyons
in Subsec. 7.3, with domain walls in Subsec. 7.4 and,
lastly, with boundaries and corners in Subsec. 7.5 and
Subsec. 7.6.
7.1 Twist defects
We start by labeling the twist defects. To this end,
we consider an anyon a that is moved once around
a twist before returning to its initial position. The
anyon will cross the domain wall ϕ which the twist
terminates once under this monodromy, and as such
is mapped onto the excitation ϕ(a). We label a twist
ϕ if it acts on C with the mapping ϕ under a single
clockwise monodromy.
With this definition we observe particle anti-
particle like creation of pairs of twists when adding an
open domain wall to the bulk of a phase, see Fig. 13.
Indeed, we observe a twist ϕ at one terminal point of
the domain wall and ϕ at the other.
Figure 13: Twists terminating a domain wall correspond to
the automorphism as the domain wall. Consider a domain
wall which maps an anyon a according to ϕ if the anyon gets
transported in one direction across. If the anyon goes the
other way, the inverse ϕ is applied. The two twists lying on
the terminal point of the domain wall are drawn as purple
squares. Moving the anyon a counter-clockwise around the
twists applies the automorphism ϕ for one twist and ϕ for
the other twist terminating the domain wall.
We now go on to discuss the explicit example of
twist defects in the color code. Specifically, we con-
sider a generating set of twist defects, B, Y and D,
which are the terminal points of their corresponding
domain walls. This generating set will be sufficient to
generate all of the twists of the color code using twist
composition rules we discuss later in Sec. 7.2.
We first look at a lattice realization of the blue-
invariant twist B lying at the terminal point of the
blue-invariant domain wall, see Fig. 14. On the lat-
tice, the terminal point of the domain wall lies at a
plaquette with an odd number of vertices. This means
it can no longer support both an x-type and a z-type
stabilizer, as they would not commute. Instead, a
single basis has to be chosen in which the parity is
measured.
A Y twist lies at the terminal point of the domain
wall that permutes the Pauli-X and Pauli-Z label of
the color code bosons. Unlike a B twist we do not
need to change the structure of the lattice to introduce
a Y twist. However, at the terminal plaquette of the
domain wall we can impose only a single stabilizer of
Pauli-Y terms, as is depicted in Fig. 15.
Figure 14: Shown is a lattice representation of the B twist.
It features a weight-seven stabilizer which is just measured
in one basis, the y-basis for example. If anyonic charges are
moved once around it, their color label changes according to
the automorphism B.
Accepted in Quantum 2018-10-09, click title to verify 14
Figure 15: A y-invariant domain wall can be terminated by
a plaquette which hosts one single stabilizer acting in the
y-basis. This constitutes a Y twist which exchanges x and
z Pauli labels of color code anyons under monodromy.
We finally look at the twist D that terminates the
domino-brick domain wall. As shown in Fig. 16, at
the terminal point of the domain wall we have one
weight-five and two weight-seven stabilizers.
While in some cases it is possible to propose twist
defect stabilizers via an educated guess, the stabilizers
can be rather complex; this is particularly true in the
case of D twists. In fact, it is possible to essentially
eliminate any guess work and derive the stabilizers for
the twist defect by considering a sequence of Pauli op-
erations which by construction preserves the ground
space. First, we create a pair of anyons, a and ¯a.
Excitation a is then transported around the defect `
times, following its label change every time it crosses
the domain wall ϕ such that ϕ
`
(a) = a. Upon com-
pleting this repeated monodromy we can annihilate
the two excitations to obtain a new stabilizer opera-
tor from the product of the Pauli operators used to
realize this process. In order to find low-weight stabi-
lizers, the resulting string operator can be contracted
by multiplying it with conventional stabilizers that lie
on the lattice.
Figure 16: The domino twist D has three changed stabiliz-
ers. One of the Domino brick stabilizers is cut so that it only
supports five qubits. Additionally we have two weight-seven
stabilizers, overlapping on six oft their qubits.
Figure 17: Similar to how the fusion of domain walls cor-
responds to composing maps, the same holds for twists. In
this example, two twists described by the maps ϕ and ϕ
0
are
combined to create a twists described by ϕ
0
ϕ.
7.2 Fusion between twists
The fusion of domain walls can be described by com-
position maps, see Sec. 6.1. Twists correspond to
terminal points of domain walls, and hence the fu-
sion of twists is also described by the composition
of maps, see Fig. 17. Given that the symmetry of
an anyon model forms a group, the fusion of pairs of
twists result in other twists. This means places where
a domain wall changes type or where multiple domain
walls meet are also twists.
For an example, in the color code we can consider
the fusion of two color exchanging twists, R and B, to
form a cyclic color permuting twist, RB, see Fig. 18.
Note that the lattice representation of the RB twist is
not presented in the main text, we refer the reader to
App. C, in particular to Fig. 36 (a). This fusion can
be visualized nicely on the dual lattice using Pachner
moves, see Fig. 38 in App. C.
In Fig. 17 we have deformed the domain walls up-
wards at the point where the twists meet. We thus
obtain a twist ϕ
0
ϕ. Instead we could have chosen
to bend the domain walls downwards, such that the
twist maps a according to ϕ ϕ
0
(a). This ambiguity
in the label is a general feature of twist fusion. To
rectify this, we need to consider conjugacy classes of
the symmetry group. The label used to describe a
twists does in general depend on an arbitrarily cho-
Figure 18: An R twist on the left and a B twist on the right
are fused to form an RB twist. The RB twist features two
plaquettes with odd weight, where only one stabilizer each is
measured.
Accepted in Quantum 2018-10-09, click title to verify 15
Figure 19: This figure depicts the fusion of two twists de-
scribed by the symmetries ϕ and ϕ
0
. On the left, we illustrate
how the reference point used to label the fused twist matters.
If the reference point is chosen to lie above the twist, as in
the top left, we obtain a map ϕ
0
ϕ. Whereas a reference
point below the twist results in the description ϕ ϕ
0
, as
shown in the bottom left. However, one description can be
changed to the other by wrapping a domain wall around the
twists. This is seen on the right, where the lower descrip-
tion is obtained by enclosing the three twist with a closed ϕ
domain wall.
sen starting point. Consider Fig. 19, where a twist is
labeled by two different maps on the left half of the
figure. Shown on the right, we see that the two maps
are equivalent up to conjugation. Here, conjugating
the twist is equivalent to changing gauge or engulf-
ing the entire twist compound inside a closed domain
wall. If we do this in the example of Fig. 19, as shown
on the right, we can convince our selves that the two
maps are equivalent up to conjugation with ϕ.
In a general example, where multiple domain walls
meet, even more than two different but valid descrip-
tions might be found. And in general, without fixing a
global gauge and precise conventions, twists are only
identified up to conjugacy.
In the color code, twists belong to one of nine equiv-
alence classes. Tab. 4 lists the conjugacy classes A to
I of all 72 twists. Within an equivalence class twists
X Y Z XZ ZX
A B B B D D
R B C C C E E
G B C C C E E
B B C C C E E
RB D E E E F F
BR D E E E F F
D () X Y Z XZ ZX
G H H H I I
R H G I I H H
G H I G I H H
B H I I G H H
RB I H H H I G
BR I H H H G I
Table 4: The equivalence classes of the 72 color code twists.
Each entry corresponds to a twist obtained by fusion of the
twist labeling the row and column. In the second table, every
twist is additionally fused with the domino twist D. We find
nine conjugacy classes, labeled A to I here.
Figure 20: Shown is a twist described by the map ϕ. We
consider an anyon a which can decay into ϕ(b) and b. If
the ϕ(b) anyon is now moved counter-clockwise around the
twist, it crosses the domain wall described by ϕ. Hence it
turns into a b anyon, which can fuse with b to form the trivial
particle 1. This process is called localization of anyon a at
twist ϕ.
share many properties, such as their order or their
quantum dimension (see Tab. 5).
7.3 Anyon localization and the quantum di-
mensions of twists
Similar to how certain anyons can condense at bound-
aries, twists can also absorb anyons. This process is
called localization of an anyon at a twist. We can ex-
press the localization of an anyon a at twist ϕ with
the fusion rule
ϕ × a = ϕ. (18)
To determine the anyons that can localize at a twist,
we study its automorphism ϕ together with the fusion
multiplicities of the anyon model N
c
a,b
. In particular,
if there exists some excitation b C such that
N
a
ϕ(b),b
= 1, (19)
then we have that ϕ × a = ϕ. To show why this is
true, we study Fig. 20. The figure shows a twist ϕ and
an anyon a which decays into a pair of particles ϕ(b)
and b. ϕ(b) is then moved counter-clockwise around
the twist such that the resulting excitation ϕ ϕ(b) =
b annihilates with b and restores the system to the
ground state.
To illustrate anyon localization in the color code,
consider the three examples shown in Fig. 21. The B
twist can localize all blue bosonic anyons by decay-
ing them into red and green anyons. If one is moved
around the twist it changes color and can annihilate
with the other anyon. Cyclic color permuting twists
can annihilate all sixteen color code anyons. This is
because all anyons are composed of the nine bosonic
anyons, which in turn can decay into two anyons of
different color. If one is moved around the twist one or
two times, it can be brought to match color with the
other one, so that they annihilate. Lastly, the D twist
can condense one of the fermionic particles. Fermions
in the color code are composed of two bosonic parti-
cles with differing color and Pauli labels, see App. A.
Accepted in Quantum 2018-10-09, click title to verify 16
Figure 21: Left: A B twist, which exchanges the red and
the green color labels of anyons is shown. A blue anyon with
any Pauli label can localize at the twist. We decay it into a
red and a green anyon, which when moved around the twist
match color and annihilate. Middle: A cyclic color twist
can localize anyons of any color. This is because an anyon
can obtain all three color labels if moved around the twist
sufficiently many times. Right: The D twist can condense a
single fermionic anyon. In this case, the f
1
= rx × bz anyon
gets localized. The dashed line indicates the different Pauli
label of the red anyon.
The domino twist D as we defined it can localize the
fermion f
1
= rx × bz. To see this, we take one of its
constituent bosons, say rx, move it around the twist
once to transform it into bz which annihilates with
the other constituent part of f
1
.
The process of anyon localization at twists, as de-
scribed by Eq. (18), is reminiscent of the fusion in non-
abelian anyon models. Further, as with non-abelian
anyons, we can encode logical qubits in the fusion
space of twists. To increase the number of logical
qubits, more twists can be added to the system. The
quantum dimension of a twist defect is the factor by
which the fusion space dimension grows, as a twist
of that type is added. We calculate the quantum di-
mension of twist ϕ, denoted d
ϕ
, with a formula that
is analogous to that of non-abelian anyons given in
Eq. (2)
d
2
ϕ
=
X
a∈C
d
a
N
ϕ
ϕ,a
, (20)
where N
ϕ
ϕ,a
tells us if the anyon a can localize at the
twist ϕ. In addition to calculating the quantum di-
mension of a twist defect directly from its fusion rules,
we also point out work where, like with non-abelian
anyons [83, 84], the quantum dimension of twist de-
fects can be evaluated by studying the entanglement
entropy of the ground state of topological phases that
include twist defects [8587].
Since all anyons in the color code are abelian, their
dimension is 1 and hence the quantum dimension of a
twist is just the square root of the number of anyons
it can localize. As mentioned above, the B twist for
example can, apart from the trivial charge, localize
all three blue bosonic anyons. Thus, its quantum di-
mension is d
B
= 2. Cyclic color twists can localize all
sixteen anyons and have quantum dimension d
BR
= 4.
The domino twist D can localize only a single fermion
and the vacuum and has dimension d
D
=
2. The
X Y Z XZ ZX
1 2 2 2 4 4
R 2 2 2 2 4 4
G 2 2 2 2 4 4
B 2 2 2 2 4 4
RB 4 4 4 4 2 2
BR 4 4 4 4 2 2
D () X Y Z XZ ZX
2
8
8
8
8
8
R
8
2
8
8
8
8
G
8
8
2
8
8
8
B
8
8
8
2
8
8
RB
8
8
8
8
8
2
BR
8
8
8
8
2
8
Table 5: The quantum dimension of the 72 color code
twists. Each entry corresponds the twist obtained by fus-
ing the twists labeling the row and column. In the second
table, every twist is additionally fused with the domino twist
D.
quantum dimension of all 72 color code twists is given
in Tab. 5.
Anyons which do not localize at a twist might get
identified with other non-trivial anyons. Say anyon
a localizes at twist ϕ and hence a × ϕ = ϕ, and an
anyon c can decay to a and b, i.e. c = b ×a. Then, we
cannot distinguish the two twist-anyon compounds,
c × ϕ = b × a × ϕ = b × ϕ. Microscopically, twist-
anyon compounds are anyons located close to the
twist defect. Mathematically, the twist-anyon com-
pounds form a G-crossed braided tensor category, for
more details see Ref. [15]. In this manuscript we are
solely concerned with the ground-state space, or code-
space when viewed as a quantum code. This means
we will focus on twists which are locally equivalent to
the twist with the trivial particle 1 localized.
7.4 Interplay between twists and domain walls
We now study crossing domain walls and the twists
which terminate them. This is relevant when twists
are moved over domain walls using, for instance, code
deformations [12, 17, 21, 44, 88]. To figure out how
the twist changes after crossing a domain wall we con-
sider closed contractible loops engulfing the crossing
point of the domain walls. If an anyon was moved
along such a contractible path until it reaches its orig-
inal position, it should return to its initial charge la-
bel. Consider a twist which acts as the map ϕ on
the anyon labels, as depicted in Fig. 22. After it
crosses a domain wall ϕ
0
it transforms into a twist
acting as ϕ
00
= ϕ
0
ϕ ϕ
0
. An anyon a which moves
on the aforementioned path then gets mapped by
ϕ
00
ϕ
0
ϕ ϕ
0
= . Remember, the equality stems
from the fact that the loop followed by the anyon is
contractible. This means the symmetry describing the
Accepted in Quantum 2018-10-09, click title to verify 17
Figure 22: Consider a twist described by ϕ, here drawn at the
bottom, terminating a solid domain wall. It crosses a dotted
domain wall ϕ
0
. This transforms the twist to ϕ
00
. An anyon
a starting in the top-left quadrant and traveling counter-
clockwise around the crossing point gets finally mapped to
ϕ
00
ϕ
0
ϕ ϕ
0
(a). Since its path is contractible, the final
label has to be equal to its original label a. Following this
equivalence for a generating set of anyons identifies the twist
ϕ
00
= ϕ
0
ϕ ϕ
0
as the conjugation of ϕ with ϕ
0
.
transformed twist is obtained by conjugating its origi-
nal symmetry with the one of the crossed domain wall.
Hence, twists crossing domain walls will stay within
the same equivalence class, see Sec. 7.2.
Again, let us illustrate this with a color code exam-
ple. Consider a color exchanging twist, say G which
crosses the domino domain wall described by the dual-
ity symmetry D. Using the procedure described above
for a generating set anyons, rx, rz, bx and bz shows
that the twist transforms to a Pauli exchanging twist
Y , see Fig. 23. This is to show the duality transfor-
mation D does affect twists in very much the same
way as it does anyons. The Pauli label of twists and
the color labels are exchanged.
As a second example we consider a color exchang-
ing twist G crossing a color exchanging domain wall
R which turns it into a B twist. This follows from the
Figure 23: A G twist (bottom) transforms to a Y twist (top)
if moved over a domino domain wall D (middle). This can be
seen by starting with a generating set of four anyons, rx, rz,
bx and bz, in the top-left quadrant. Moving them counter-
clockwise around the crossing point to the top right, they
transform to rz, rx, bz and bx, respectively. Moving them
back to their initial position over the top half of the vertical
domain wall should return them to their initial charge. The
only domain wall with this effect is the one terminated by a
Y twist.
Figure 24: Lattice representation of a twist crossing a do-
main wall. A G twist (bottom) is moved over an R domain
wall (middle) and transforms to a B twist (top). Apart from
the plaquettes hosting the twists marked by purple squares,
all plaquettes host two independent stabilizers.
fact that GRG = B. Again, we see that the domain
wall has an equivalent action on the twist label as it
would on anyon labels, see Fig. 24 for a lattice repre-
sentation. This example can again be understood nat-
urally using the Pachner move construction for twists
and domain walls, see Fig. 41 in App. C.
7.5 Confinement of twists at boundaries
In this subsection, we explore how twists and domain
walls interact with boundaries. In particular, a do-
main wall covering a boundary can change the bound-
ary type. This implies that a twist defect confined at
the boundary corresponds to a corner. A corner is
the point where two different boundaries meet, as de-
scribed in Sec. 4. If the boundary type is not changed
by the automorphism applied by the domain wall then
the corresponding twist can condense at said bound-
ary.
Let us begin by investigating how boundaries
change when domain walls are present. Consider a do-
main wall described by the automorphism ϕ close to
a boundary which is characterized by its Lagrangian
subgroup M, see in Fig. 25 (a). If an anyon a is con-
densible at M, i.e. a M, then the anyon ϕ(a) can
condense at the boundary by first moving over the
domain wall. This transforms it to ϕ ϕ(a) = a
M. Clearly, an anyon ϕ(a) can still condense when
the domain wall covers the boundary, as depicted in
Fig. 25 (b).
In fact, the anyons ϕ(a) form a Lagrangian sub-
group M
0
:
= ϕ(M) which may condense at the re-
sulting boundary. This follows from the properties of
Lagrangian subgroups, discussed in Sec. 4.1, and the
conditions put on the symmetries of the anyon model,
see Eqs. (15-17). To illustrate this, the covering do-
main wall was removed in Fig. 25 (c) and the new
boundary M
0
drawn as a dashed line. In this pic-
ture the twists correspond to the corners between the
different boundaries. In an alternative reading of the
above we have that the corners correspond to confined
Accepted in Quantum 2018-10-09, click title to verify 18
Figure 25: Corners, points where the boundary type changes,
can be identified with confined twist defects. In (a), a pair of
twists ϕ and ϕ sit close to a boundary M where anyons a
M can condense. The boundary behind the domain wall is
able to condense anyons ϕ(a), if moved over the domain wall
first. (b) shows same situation as (a), but the twists and their
connecting domain wall are now covering the boundary. This
corresponds to (c), where boundary covered by twists and
the domain wall now correspond to corners and a changed
boundary denoted by the dashed gray line. If the original
boundary is described by the Lagrangian subgroup M, then
the dashed domain wall is described by M
0
:
= ϕ(M). Note,
this encapsulates trivial corners, i.e. the case where M =
M
0
. We call twists ϕ
0
for which ϕ
0
(M) = M condensible
at the boundary M.
twists. In contrast, if an automorphism ϕ
0
leaves the
boundary type unchanged, i.e. ϕ
0
(M) = M, we say
the twist ϕ
0
can condense at the boundary M. Twist
condensation is discussed in more detail in Sec. 7.6.
This leads to an ambiguity, corners can not, in gen-
eral, be uniquely identified with a single twist, as they
might fuse with condensible twists. If ϕ(M) = M
0
and ϕ
0
(M) = M, then ϕ ϕ
0
(M) = M
0
. This means
a twist ϕ ϕ
0
, the fusion outcome of a condensible
twist ϕ
0
with a corner twist ϕ is again a valid inter-
pretation of the corner.
Let us now turn to the corners in the color code for
concrete examples of the aforementioned ambiguity in
the association of corners with twists. Consider the
common triangular color code with the three color
boundaries, red, green and blue. This code can be
constructed in the several ways from a disk with a
single red boundary and adding open domain walls to
segments of the uniform red boundary. We show two
such ways in Fig. 26. We first consider applying a do-
main wall that implements the G symmetry to cover
one red colored boundary segment which transforms
it to the blue color boundary. Similarly, the B domain
wall creates a green boundary when covering the sec-
ond part of the red boundary. Hence, the corner that
lie between the red-blue(red-green) boundary can be
interpreted as a G (B) twist. The blue-green cor-
ner then is the fusion of the two twists which we call
GB = BR. This is depicted on the right in Fig. 26.
Alternatively we can use the cyclic domain walls RB
and BR to transform a segment of the red boundary
onto a blue or green boundary, respectively. Hence,
as shown on the left in Fig. 26, we can label each
Figure 26: The identification of color code corners with
twists is ambiguous. We start with a color code with just a
single red boundary, as shown in the first row. Consecutively,
we transform one boundary after the other by covering them
with domain walls, until the common triangular color code in
the bottom center is obtained. This construction shows the
non-uniqueness of the interpretation of corners as twists. On
the left, we end up with three BR twist, one in each corner,
where on the right only one corner is identified with a cyclic
twist.
of the three corners as a cyclic color permuting twist
BR. The two approaches to realize the color code
with three differently colored boundaries differ by an
R twist fused with the two corner twists that lie at
the red boundary. However, given the red boundary
is invariant if we cover it with a red domain wall, we
see that the physics of the two models we have pro-
duced is invariant under the fusion of the G and B
twist of Fig. 26 with red twists. We thus see that the
two pictures we have produced are consistent.
Corners between Pauli boundaries can be identified
equivalently with Pauli permuting twists. Corners be-
tween Pauli and color boundaries host twists shown
in the second half of Tab. 4, who involve the D twist
when generated from the standard set of symmetries.
This is shown in more detail in Fig. 28, only that
there, the corners are detached and pulled into the
bulk as twists. This process is discussed in Sec. 7.6.
7.6 Detaching corners and twist condensation
Not only can we identify corners with certain twists,
we can also detach corner twists and move them into
the bulk. Consider Fig. 27 (a), where we interpret
the M
0
boundary as a domain wall ϕ covering the
M boundary, and the corners as twists. These twists
can be moved, for example through code deformation,
such that they come to lie in the bulk, see Fig. 27 (b).
This leads to Fig. 27 (c), the twist is detached from
the corner and pulled into the bulk.
Notice how this framework also encapsulates the
Accepted in Quantum 2018-10-09, click title to verify 19
Figure 27: (a) Shows the same situation as depicted in
Fig. 25 (b), where the M
0
boundary is seen as a domain
wall covering the M boundary, and the corners as twists. In
(b), the domain wall is extended such that the ϕ twist now
lies in the bulk. This corresponds to (c), where the domain
wall covering the boundary is omitted to illustrate that it
constitutes a different boundary. Now, the situation looks
like a single twist whose domain wall terminates between the
M and M
0
boundaries. This can be read as a detached
corner which is pulled into the bulk as a ϕ twist. (d) shows
the case where M = ϕ
0
(M), the trivial corner. Twists which
map the Lagrangian subgroup of a boundary back to itself are
able to condense at and being emitted from said boundary.
trivial corner, i.e. the case M = M
0
. This means
a twist ϕ
0
which maps the Lagrangian subgroup de-
scribing a boundary back onto itself, i.e. ϕ
0
(M) = M,
can condense at or be emitted from said boundary, see
Fig. 27 (d).
To illustrate the above with practical examples, let
us turn our focus again to the color code model. Start-
ing with twists condensation, we can observe that all
Pauli twists can condense at color boundaries, since
they map the Lagrangian subgroups of color bound-
aries back onto themselves. Additionally, color ex-
changing twists which leave one color invariant can
condense at and be emitted from the boundary of said
color. As an example, consider Fig. 28 (a), where an
R twist is emitted from a red boundary. Pauli bound-
aries can emit Pauli exchanging twists which leave the
relevant label invariant, as well as all purely color per-
muting twists and twists which are the fusion product
of these examples. See Fig. 28 (b) for an example of
an X twist emitted from an x Pauli boundary. Let
us now consider the different color code corners, dis-
cussed in Sec. 4.3 and depicted in Fig. 6. Corners
between two color boundaries can be interpreted as
the twist which exchanges them. For instance, a cor-
ner between a blue and a red color boundary can be
detached and pulled into the bulk as a G twist, see
Fig. 28 (c). Alternatively, the cyclic color twist RB
can be pulled into the bulk as a detached blue-red cor-
ner, see Fig. 28 (d). The same holds for Pauli/Pauli
corners, consider Fig. 28 (e) showing a Y twist as a
detached x-z corner. The last type of corner where
color and Pauli boundaries meet, corresponds to D
type twists. Fig. 28 (f) shows a detached color/Pauli
corner as a D twist in the bulk.
Twist condensation and detaching corner twists
Figure 28: Lattice representations of twists condensing at
boundaries and detached corner twists. (a) an R twist emit-
ted from a red color boundary. (b) an X twist emitted from
x Pauli boundary. (c) a G twist as a detached blue-red
color/color corner. (d) a BR twist as a detached blue-
red color/color corner. (e) a Y twist as a detached x-z
Pauli/Pauli corner. Notice, on the boundary plaquettes the
changed orientation of the stripe pattern indicates a change
in basis. (f) a D twist as a detached red-x color/Pauli corner.
can be used in fault-tolerant quantum computation.
Topological degrees of freedom associated to twist de-
fects, in which twist encoded logical qubits can be
stored, become local degrees of freedom when the
twist is condensed at a boundary. Measuring these
local degrees of freedom then corresponds to measure-
ments on logical qubits. In the reverse process, where
twists are emitted from corners or boundaries, the
logical degrees of freedom increase. This is similar to
considering non-abelian anyons and a boundary where
such anyon excitations condense. This is similar to
how merging and splitting codes using lattice surgery
changes the number of logical qubits [42]. This is
no coincidence, as discussed in Ref. [23], identifying
corners with twists unifies encoding schemes relying
on boundaries with twist encoding schemes. This of-
fers a new understanding of lattice surgery used in
recent proposals for efficient fault-tolerant quantum
computation schemes [89, 90]. Lastly, identifying cor-
ners as twists can help in the search for hardware ef-
ficient quantum error-correcting codes. For instance,
Accepted in Quantum 2018-10-09, click title to verify 20
the code presented in Ref. [22] may be understood as
a surface code where one corner twist is detached and
pulled into the center of the lattice. This lowers the
overhead of physical qubits needed to encode logical
qubits at a given code distance. Even more efficient
codes based the same insight are presented in App. D
and in the following Sec. 8, where G twists in the cen-
ter of the code can be interpreted as detached corners.
8 Stellated color codes
We next present a new family of two-dimensional sta-
bilizer codes that make use of the twist defects of
the color code phase. The proposed codes reduce
the number of physical qubits required per code qubit
while preserving desirable properties such as code dis-
tance and geometric locality of stabilizer generators in
a 2D euclidean lattice.
For topological stabilizer codes on flat manifolds,
the number of logical qubits k one can encode using n
physical qubits with code distance d is upper bounded
by the holographic bound [63] for codes defined by
geometrically local commuting constraints embedded
in 2D Euclidean space
k c
n
d
2
. (21)
Here, c is a constant which depends only on the lo-
cality of the constraints defining the error-correcting
code. One can compare these c-values of different
codes, where a higher c roughly corresponds to a
more efficient encoding scheme. For the surface code
with three-body boundary terms, as proposed by Ki-
taev [2], one obtains c =
1
2
. This value can be dou-
bled to c = 1 by rotating the code and using two-
body boundary terms, resulting in the rotated surface
code [20, 64, 91]. The same c-value can be achieved
by appropriately arranging punctures [65]. In surface
codes, a further increase of the c value is possible us-
ing twist defects. The triangular surface code [22]
and certain configurations of twists and holes in hy-
brid encoding schemes [23] achieve c =
4
3
. The same
value, c =
4
3
, is found for the triangular 6.6.6 color
code. Using a 4.8.8 lattice, the triangular color code
with color boundaries reaches c = 2 [56]. As far as we
are aware of, this is the highest c-value of any known
flat two-dimensional topological error correction code
with bounded stabilizer weight w 8. Hyperbolic
codes [9294] exceed the bound given in Eq. (21), but
are expected to come along with significant experi-
mental challenges, as they can not be embedded in
two spatial dimensions while maintaining the locality
of stabilizers.
Inspired by the triangular surface code [22], we
start by presenting an analogous color code model
that achieves c =
8
3
. It can be obtained by fatten-
ing [95] the vertices of the triangular surface code to
quadratic faces. Alternatively, this process can be
Figure 29: Three members of the family of stellated 4.8.8
color codes. The three shown examples have a code distance
d = 9. Left: the s = 3 case is shown. This code encodes
two logical qubits. The middle and the right show the cases
s = 5 and s = 7, encoding four and six logical qubits, re-
spectively. The logical operators apply X, Y and Z on qubits
supported by strings connecting different boundaries. Some
logical operators are shown in the middle code as examples.
understood in terms of code concatenation with the
J4, 2, 2K-code, as discussed in Ref. [68]. It leads to the
left most code shown in Fig. 29. In the figure, a single
qubit lies on each vertex of the lattice and each pla-
quette p supports two stabilizers, s
x
p
=
Q
jp
X
j
and
s
z
p
=
Q
jp
Z
j
, where p denotes the set of qubits
lying on the vertices adjacent to p. This is the same
as the standard color code stabilizers, see Eq. (9). We
leave a puncture with a green boundary at the center
of the model where no stabilizer is enforced.
The properties of the model we have described are
understood naturally from the perspective of non-
abelian twist defects. We first point out that the
color symmetry is broken when we undergo a full mon-
odromy of the central puncture. As such, we find a
single G-twist is condensed in the central puncture,
connected to the boundary with a domain wall. Fur-
ther, there are exactly three corner twists of type G
lying at the interfaces where red and blue boundaries
meet. The ground-state degeneracy then is equiva-
lent to the fusion space of four G-twists that are con-
strained to fuse to the vacuum. The central puncture
encodes a qubit, which we treat as a gauge qubit.
The described code is a representative of a whole
code family we call stellated color codes. Codes in this
family are parameterized by s, the order of rotational
symmetry of the code. The representatives with s =
3, 5 and 7 of the stellated 4.8.8 color code family are
shown in Fig. 29. Again, each code supports s+1 color
permuting twists for odd s. Note that in principle s
can be chosen to be even, in which case the trivial
twist, i.e. no twist, is featured in the center. The case
where s = 4, for example, is a rectangular 4.8.8 color
code featuring alternating red and blue boundaries.
For odd s, a stellated code encodes k = s 1 logical
qubits and reaches a c-value of c = 4
4
s
. Stellated
codes with even s encode k = s 2 logical qubits and
we find c = 4
8
s
. In both cases, c approaches 4 as
s grows. Allowing s to diverge doubles the highest
known values of c for flat topological error-correcting
codes such that c 4. The advantage is achieved
Accepted in Quantum 2018-10-09, click title to verify 21
Figure 30: Examples of stellated color codes on a 6.6.6 lat-
tice. These are the three smallest codes with s = 5, depicted
with code distances d = 3, 5, 7 from left to right. The shown
codes each encode two logical qubits. The encoding rate for
the 6.6.6 stellated color codes approaches c =
8
3
2.7 for
large s, which, to the best of our knowledge, is the high-
est encoding rate of any flat topological quantum code with
w 6.
by locally changing the curvature around the central
plaquette which is known to enhance the encoding
rate in hyperbolic codes [93, 94].
Given that the measurement of high-weight sta-
bilizers can be challenging experimentally we also
present the stellated color codes based on the 6.6.6
honeycomb lattice. Their encoding rate is c =
8
3
8
3s
(c =
8
3
16
3s
) for odd (even) s. See Fig. 30 for three
s = 5 representatives. These codes are, to the best
of our knowledge, the most efficient topological error
correction codes with stabilizer weights w 6 which
are embeddable in two spacial dimensions.
Interestingly, the discussed increase of s can also be
applied to the twisted surface code model [22]. The
encoding rates for these models halve compared to
the stellated 4.8.8 color codes. However, the stel-
lated surface codes we present may be of interest
as typically surface code based architectures have
higher fault-tolerance thresholds than those based on
the color code model [9]. To substantiate such a
claim, one would likely need to conduct numerical
analyses which goes beyond the scope of the present
work. We give a brief description of stellated surface
codes in App. D. One should also investigate differ-
ent ways of performing logical operations on qubits
encoded in stellated codes. As a first approach one
could take the description of these models in terms
of their twists and generalize the schemes presented
in, for instance, Refs. [8, 22, 27, 43, 90], where fault-
tolerant approaches to braid and fuse twist defects
are considered. We leave the problem of optimiz-
ing fault-tolerant protocols for quantum computation
with these models to future work.
9 Unfolding boundaries and twists
It is well established [33, 6668] that we can locally
map the color code onto two copies of the toric code
model [2]. This is clear because the excitations of the
color code have the same braid statistics and fusion
rules of their analogous excitations in the mapped pic-
ture, and hence these two models are members of a
common quantum phase [96, 97]. So too then must
two copies of the toric code share the Lagrangian sub-
groups and excitation symmetries of the color code,
and thus give rise to analogous boundaries and twists.
In this section we concisely describe these analogous
objects. We study the different boundaries of the
model, as well as a generating set of twist defects.
To briefly review, all anyons in the toric code are
abelian and take labels 1, e, m and ε = e ×m, where
e and m are bosons and ε is a fermion. Braiding
two different non-trivial toric code anyons results in a
phase, M
e,m
= M
e,ε
= M
m,ε
= 1. The Lagrangian
subgroups of the toric code are thus M
e
= {1, e} and
M
m
= {1, m} [3, 14, 98]. The boundaries that con-
dense the excitations M
e
(M
m
) are more colloquially
known as rough (smooth) boundaries. Moreover, the
model has a single non-trivial domain wall which ex-
changes e and m excitation [10, 12, 14], and thus gives
rise to a single twist defect that resembles a Majorana
mode [10, 12, 21].
We specify the excitations of the system of two
copies of the toric code as a
±
for a = e, m, ε where
the ± symbol indicates which of the two copies the
excitation lies on. We can consistently map the color
code excitations onto the dual-layer toric code model
under the following correspondence
rx e
, rz m
+
,
bx e
+
, bz m
,
(22)
which succinctly describe a generating set for C under
fusion [66]. We can also populate Tab. 1 as shown in
Tab. 6.
x
y
z
r g b
e
e
m
+
m
+
e
e
+
ε
ε
+
m
m
+
e
+
m
e
+
m
Table 6: Color code anyons after unfolding the color code
into two toric code layers. e and m refer to electric and mag-
netic charges, the superscript indicates on which toric code
layer the anyon lives. Certain color code anyons correspond
to combinations of two toric code anyons. The shown corre-
spondence is found if the lattice is contracted on the blue and
red lattices to form the and + layer toric codes respectively,
following the unfolding procedure laid out in Ref. [33]. Note
that this relabeling may be defined up to an automorphism
ϕ of the anyon model.
We next investigate the boundaries of the color
code. The boundaries described by Lagrangian sub-
groups that correspond to the columns of Tab. 6 have
been studied in Ref. [33]. In particular, the system
boundary where the top copy of the toric code is rough
and the bottom copy is smooth corresponds to a red
boundary in the color code picture. Likewise, the
Accepted in Quantum 2018-10-09, click title to verify 22
Figure 31: Illustrated is the folding of a toric code to obtain
a triangular color code. The top, (b) - (e) row which starts
with a toric code with a regular boundary configuration re-
sults in a triangular color code with color boundaries. The
bottom row (f) - (i) starts with a toric code with a more
unconventional domain wall arrangement and ends with a
triangular color code with Pauli boundaries. The square (a)
represents a patch of toric code with smooth boundaries. In
(b) and (f) we introduce twists at the corners, connected by
domain walls which change the boundary type from smooth
to rough, see (c) and (g). Now, the code is folded in half on
top of itself to get (d) and (h). Depending on the domain
wall configuration, we get a code equivalent to a triangular
color code with color boundaries (e) or Pauli boundaries (i).
boundary where the top (bottom) copy of the toric
code is smooth (rough) corresponds to a blue bound-
ary. The green boundary corresponds to a fold where
e
+
× e
and m
+
× m
are absorbed. In Fig. 31 (c)
and (d), we show the planar code with two rough
boundaries and two smooth boundaries being folded.
In Ref. [33] it was argued that this model is analogous
to the standard color code with a red, blue and green
boundary [28].
Before describing the Pauli boundaries in this
model, it will be helpful to understand the toric code
model with rough and smooth boundaries as a model
with only smooth boundaries but where twist defects
lie at its corners. Indeed, in Ref. [23] it was sug-
gested that we should regard a rough boundary as
a smooth boundary with a non-trivial domain wall
stretched across its surface. Using this picture we see
that twist defects lie at the corners of the planar code.
In Fig. 31 (b) we show a configuration of domain walls
that gives rise to the conventional planar code with
two rough boundaries and two smooth boundaries.
The color code with three different Pauli bound-
aries can also be understood from a folded system. In
contrast, we can regard the color code with three dis-
tinct Pauli boundaries as the folded code model with a
domain wall configuration shown in Fig. 31 (f). With
this domain wall configuration we obtain through
folding boundaries where both the top and bottom
layers are both rough or both smooth, see Fig. 31 (h).
These boundaries correspond to x and z Pauli bound-
aries, respectively. Finally, the y Pauli boundary
correspond to a fold with a non-trivial domain wall
stretched across it. With this we complete the bound-
aries of the unfolded color code, as we have now explic-
Figure 32: A domain wall and twist in two layers of toric
code which corresponding to a D symmetry in the color code.
It consists of a toric code domain wall in the second layer.
All stabilizers are shown in the legend below.
itly realized boundaries that condense the Lagrangian
subgroups of the rows of Tab. 6.
We finally look at the domain walls and twist de-
fects of the unfolded color code. Again, we consider a
generating set of three domain walls. As discussed in
Sec. 6 it is sufficient to find a domain wall that trans-
poses Tab. 6 together with two more domain walls
that permute the rows of the table.
The domain wall that exchanges the excitations ac-
cording to a transposition of Tab. 6 is the mapping
e
m
. Physically, this domain wall is the well
studied domain wall of the toric code [12, 14] on the
bottom layer of the two-copy system. We show this
domain wall together with its termination twist defect
in Fig. 32.
Next, we discuss the domain wall corresponding to
the G symmetry. In the unfolded color code it corre-
sponds to a layer swap. This means toric code anyons
moved over this domain wall get mapped according to
a
±
a
, where a = e, m, ε. See Fig. 33 for a lattice
representation of the layer swapping domain wall and
its terminating twist.
Figure 33: A layer swapping domain wall and twist in two
layers of toric code. This corresponds to the G symmetry in
the color code. The stabilizers along the domain wall have
support on both layers, see the legend below.
Accepted in Quantum 2018-10-09, click title to verify 23
Figure 34: This domain wall and twist in two toric code layers
correspond to the Bsymmetry in the color code. Certain
anyons get cloned to the other layer when passing it.
The B domain wall in the picture of two toric codes
maps as follows, e
e
e
+
and m
+
m
m
+
. This
means certain anyons get copied over to the other
layer when passing the domain wall. On the lattice,
it can be realized as shown in Fig. 34. It is finally
worth remarking that the two domain walls analogous
to the G and B maps that we have just presented can
be realized in the two-copy system by, respectively,
applying the swap gate and CNOT gate transversally
between the two copies of the model on a region of
the lattice. We thus realize a closed domain wall at
the boundary of this region in the fashion of Ref. [45].
Finally, we discuss an equivalence between the
color code model and two copies of the three-fermion
model [69]. This connection is explored in App. B.
The mapped system gives particularly natural way to
view the symmetries, and thus the twist defects of the
color code model.
10 Conclusions and future work
In this work, we have identified all of the bound-
aries, domain walls, and twist defects of the color
code model, giving rise to a complete and compre-
hensive classification. On one hand, this endeavor
has been motivated from the mindset of condensed
matter physics. Specifically, using an exactly solv-
able toy model, we have been able to introduce a con-
crete example to give more substance to the abstract
mathematical theory of phases. Not only this, but
we have also described the model of focus in terms
of other simpler phases to achieve a better grasp on
the physics of the system. It is the hope that our
study will encourage readers to explore other explicit
phases to help build our understanding of the theory
of phases.
In addition to the exploration of fundamental
condensed-matter physics, a second and equally im-
portant motivation of our work has been to aid the
discovery of new applications of topological phases in
quantum information science. Indeed, we have ap-
plied our analysis to the development of topological
quantum codes. Topological quantum computing re-
mains the most promising road towards realizing a
scalable quantum computer. In order to contribute
to the theory of quantum codes, we have introduced
a new small code, a family of stellated codes with
competitive encoding rates and improved methods of
performing code deformations. These new applica-
tions clearly exhibit the potential to use our catalog to
help find new codes and protocols for processing quan-
tum information. However, the examples we have pre-
sented may only be the tip of the iceberg of quantum
computational schemes that are yet to be discovered
using the framework we have built for the color code
model. In particular, the present study invites more
research on how to devise fully fault-tolerant schemes
for quantum computing within such a framework as
well as on implementations.
To extend this research path further, we encour-
age the reader to experiment with the different topo-
logical features we have laid out to find other new
schemes for encoding and manipulating topologically
protected quantum information. In particular we be-
lieve taking a more systematic approach [65, 99] to
combine the different objects we have described may
be a fruitful avenue towards new discoveries. We
hope that the present study stimulates such further
research.
Acknowledgments
The authors thank A. Bauer, H. Bomb´ın, C. Brell,
J. Bridgeman, T. Jochym-O’Connor, A. Kubica,
D. Litinski, H. Poulsen Nautrup, A. Nietner, R.
Raussendorf, N. Tarantino, and T. J. Yoder for help-
ful and encouraging conversations. BJB is also grate-
ful to T. Adair, S. Bartlett, K. Korzekwa, and S.
Roberts for discussions on code deformations. Fur-
ther, we thank Zhenghan Wang for pointing out the
connection between the color code and two copies of
the three-fermion model. MSK is supported by the
DFG (CRC183, project B02). FP was supported by
the Alexander von Humboldt Foundation. JE is sup-
ported by DFG (CRC 183, project B02, EI 519/14-
1, and EI 519/7-1), the ERC (TAQ), the Templeton
Foundation, and the BMBF (Q.com). BJB is sup-
ported by the University of Sydney Fellowship Pro-
gramme, the Australian Research Council via the
Centre of Excellence in Engineered Quantum Sys-
tems(EQuS) project number CE170100009, and the
Villum Foundation.
Accepted in Quantum 2018-10-09, click title to verify 24
A Fermions in the color code
In the main text, we have discussed the existence of
six fermionic anyons in the color code. As mentioned,
fermions are the result of fusion of bosonic anyons
which do not share either the color nor the Pauli label.
For example, we obtain a fermion which we call f
1
by fusing rz and bx, i.e. f
1
= rz × bx. Since we
can let rz decay using the rule rz = gz × bz, we can
also identify a triple of bosonic anyons with the same
fermion, f
1
= gz × bz × bx. And using bz × bx = by,
we obtain an alternative way of writing f
1
in terms of
two bosons, f
1
= gz × by. In fact, one can also write
f
1
in a third way as the combination of two bosonic
fermions, namely f
1
= ry × gx. Once again, we can
use the 3×3 grid of color code bosons to represent
fermions, see Tab. 1. Specifically, in Tab. 7 we place
a spot in each of the cells of the grid that represent
the two bosons that constitute the fermion.
f
1
=
=
=
Table 7: The fermion f
1
written in three different ways as
the fusion of two bosonic anyons.
We can write all six color code fermions in three
ways as the fusion product of two bosons,
f
1
= rz × bx = gz × by = ry × gx,
f
2
= gy × bz = rz × gx = ry × bx,
f
3
= gy × bx = rx × gz = gy × bz,
f
4
= rx × gy = gz × bx = rz × by,
f
5
= rz × gy = gx × bz = rx × by,
f
6
= rx × bz = ry × gz = gx × by.
(23)
And again draw them in the 3×3 grid, see Tab. 8.
Here, different symbols where used for the different
ways of composing the fermions.
f
1
=
F
F N
N
f
6
=
N F
F
N
f
3
=
N
F
N F
f
4
=
N
F
F N
f
5
=
F N
F
N
f
2
=
F
N F
N
Table 8: The six fermionic anyons in the color code. Each
one can be created by fusion of two bosonic anyons. Each
table is the 3×3 grid introduced in Tab. 1. The different
shapes distinguish the different ways in which a fermion can
be created by fusing two bosons, as per Eq. (23).
Note, for fermions with even index, the bullets lie
on the diagonal, for fermions with odd index, they lie
on the anti-diagonal. Indeed, the distinction between
fermions with even and odd index goes deeper. Braid-
ing odd (even) fermions with even (odd) fermions
is trivial, whereas braiding them with distinct odd
(even) fermions results in a phase 1,
M
f
i
,f
j
= (1)
δ(i6=j)
(1)
ij
. (24)
Furthermore, fusion of two different odd (even)
fermions results in the third odd (even) fermion.
Whereas fusing an odd fermion with an even fermion
results in a bosonic anyon.
Fermions can not condense at boundaries, but there
are twists which localize fermionic anyons. For exam-
ple, as mentioned in Sec. 7.3 and shown in Fig. 21,
the D twist can localize the f
1
anyon, and apart from
the trivial charge 1 only the f
1
anyon. The other five
fermions also have a corresponding twist which only
localizes them. These are the twists in the conjugacy
class G in Tab. 4.
As discussed in Sec. 9, the color code is equivalent
to two copies of the toric code model. In the toric
code, there exists a single fermionic anyon which is
the product of fusing an electric flux e with a mag-
netic flux m, i.e. ε = e × m. Two copies of the
toric code feature six fermions, as the fusion of a
fermion on one layer and a boson on the other layer
produces a fermion, see Eq. (6). Written as two
toric codes, the fermions of the color code are hence
ε
, ε
e
+
, ε
m
+
, ε
+
, ε
+
e
, and ε
+
m
.
In App. B, we discuss the equivalence of two copies
of the three-fermion model and the color code. This
equivalence also makes the difference between what
we here called even and odd fermions very apparent,
as they correspond to the fermions on different layers
of the three-fermion model.
B The three-fermion model and the
color code
In the main text in Sec. 9, we have explored the equiv-
alence between the color code (CC) features and fea-
tures in two copies of the toric code (TC). In this ap-
pendix, we address a similar equivalence. Namely the
equivalence between two copies of the three-fermion
model (3F) and the color code [69].
The three-fermion model is discussed in Ref. [70]
where it is called the (D
4
, 1) modular tensor category.
It can physically be obtained as low energy excita-
tions of a topological subsystem code, as discussed
in Ref [100]. In the three-fermion model, there are
four anyonic charges. Apart from the trivial charge
1, there are three fermionic particles, f
1
f
2
and f
3
.
Fermionic here means θ
f
i
= 1. The fusion rules
are the same as in the toric code, all particles are
Accepted in Quantum 2018-10-09, click title to verify 25
their own antiparticles and fusion of two distinct non-
trivial anyons results in the third non-trivial anyon,
f
i
× f
j
= f
k
for i 6= j 6= k 6= i. Braiding two differ-
ent(identical) fermions results in a phase of 1(+1)
such that M
f
i
,f
j
= (1)
δ(i6=j)
. Note that braiding and
fusion in the three-fermion model is exactly the same
as for the toric code, only the spin of two particles is
changed.
Let us now turn towards the features of 3F, its
boundaries and symmetries. The three-fermion model
does not have any boundaries. Due to the absence
of any non-trivial bosons, the only Lagrangian sub-
group is the trivial group. On the other hand, it has
a richer symmetry compared to the toric code. All
permutations between the fermions are valid symme-
tries, thus its symmetry group is S
3
, the permutation
of a set of three elements. Hence it has |S
3
| = 6 do-
main walls and twists, including the trivial one. It
has three order-two symmetries which exchange two
fermions while leaving the third invariant, and two
order-three symmetries which cyclically permute the
three-fermions, one or the other way.
Two copies of the three-fermion model is equiva-
lent to the color code model, 3F 3F = CC. To con-
vince ourselves this is true, we show that the anyons in
3F3F are the same as the color code anyons. Braid-
ing and fusion of anyons between layers is trivial; on
a single layer, the braiding and fusion statistic is the
same in 3F as in TC. Since we know TC TC = CC,
we only need to show how the self-exchange statis-
tics in 3F 3F is the same as in CC. The spin of
a joint particle c = a
× b
+
is given by the product
of the self-exchanges on each layer, θ
c
= θa
θ
b
+
, as
per Eq. (6). In particular, the only way to obtain a
fermionic anyon is to fuse a fermion on one layer with
the trivial particle on the other layer. Hence, we ob-
tain six fermions, the 10 other non-trivial anyons are
bosonic, just as in the color code. One possible way of
writing the color code anyons in terms of 3F anyons
is shown in Tab. 9. Confirming the known CC fusion
rules and braid statistics is left as an exercise to the
reader.
x
y
z
r g b
f
1
f
+
1
f
2
f
+
3
f
3
f
+
2
f
2
f
+
2
f
3
f
+
1
f
1
f
+
3
f
3
f
+
3
f
1
f
+
2
f
2
f
+
1
Table 9: An example of writing the 3×3 grid of color code
anyons (Tab. 1) in terms of particles from two copies of the
three-fermion model 3F. The superscript indicates the layer,
the subscript the label of the fermion within a 3F layer.
With the convention used in Tab. 9 is such that the
fermion labels are related to those in App. A such that
f
+
j
= f
2j1
and f
j
= f
2j
. It is also interesting that
the color label of a boson can be determined uniquely
as the sum modulo 3 of the indices of the two parti-
cles from the three fermion layers. Equivalently, the
Pauli label corresponds to the difference modulo three
between said indices.
Now, let us study the boundaries of the color code
when read as two copies of the three-fermion model.
As mentioned, the 3F model does not have a non-
trivial boundary, as it has not a single non-trivial
bosonic particle. Instead, all six color code bound-
aries correspond to foldings between the two layers
of the 3F model, potentially with additional domain
walls on one layer. The x Pauli boundary is just the
direct fold, where f
i
turns into f
+
i
and vice versa.
This means particles of the form f
i
f
+
i
can condense
at said boundary. The y Pauli boundary consists of a
fold with a domain wall on the layer which permutes
the fermions cyclically, f
i
7→ f
i1
, where is addi-
tion modulo 3. The z Pauli boundary features the
inverse domain wall, f
i
7→ f
i1
. The color bound-
aries consists of folds together with domain walls of
order two. For the red boundary it is the domain
wall which exchanges fermions 2 and 3 in the layer,
f
2
f
3
. For the green and blue boundary the do-
main walls have the effect f
1
f
3
and f
2
f
3
,
respectively.
The symmetries of the color code can be con-
structed using the 3F symmetries. Each 3F layer has
the symmetry group S
3
, this combines to S
3
× S
3
for two layers. Additionally, we can introduce a
layer swap symmetry, which acts as f
i
f
+
i
on
the anyons. This
2
action on the group leads to
(S
3
×S
3
) n
2
= S
3
o
2
, which as we know from the
main text is the color code symmetry group. Again,
this group can be generated by three elements, the
layer swap with the generators on one of the 3F lay-
ers for example.
Through the folding of the three-fermion model, we
can interpret the corners between color code bound-
aries as color code twist defects. This is a practical
example of the theory given in Ref. [74]. Introducing
a domain wall in the folded 3F model along the crease
line changes the boundary type. Let us start with the
color code x boundary which corresponds to the triv-
ial fold between the two 3F layers. Adding a domain
wall which exchanges the second and third fermion,
i.e. f
2
f
3
, changes the x Pauli boundary to a red
color boundary. This corresponds to a D domain wall
in the color code, which as discussed in the main text,
changes an x Pauli boundary to a red color boundary.
C Pachner moves and color code
twists
The goal of this appendix is to show a systematic
way of deforming the color code lattice to create and
move color exchanging twist defects and the domain
walls they terminate. We will find that this can be
achieved using a series of simple Pachner moves on
Accepted in Quantum 2018-10-09, click title to verify 26
Figure 35: The Pachner moves needed to create and move
color twists are shown on the dual lattice (left) and the primal
color code lattice (right). (a) is a 2-2 Pachner move. Such
a move changes the three-colorability of the lattice as shown
on the right. After the move is applied, two plaquettes with
the same color touch. (b) is a combination of a 1-3 and a 2-2
Pachner move. On the primal lattice, this move introduces
(bi) or removes (bii) a square, depending on its direction. In
either case, the three-colorability of the lattice is violated,
either because a fourth color is introduced, or because two
plaquettes of the same color share an edge.
the dual of the color code lattice. This construction
also leads to a graphical understanding of some of
the features of color twists we have seen in the main
text. In particular, the fusion between color twists,
the effect color domain walls have on color twists and
condensation of color twists at color boundaries can
be visualized nicely in this framework.
Pachner moves [101, 102] are a set of transforma-
tions that generate the mapping of a triangulation of
a manifold onto any other triangulation of the same
manifold. As the color code lattice is trivalent, its
dual lattice is a triangulation [29, 30]. We find that
in general, applying Pachner moves to said triangula-
tion often introduces colored twists to the model as
they typically break the three-colorability of the lat-
tice.
To create twist defects, we distinguish between
three Pachner moves, shown in Fig. 35. Shown in
Fig. 35(a) is a 2-2 Pachner move, transforming two
triangles joint at one edge into two other triangles by
redrawing the diagonal. Fig. 35(bi) is a 1-3 followed
by a 2-2 Pachner move, transforming two triangles
into four by ‘adding a second diagonal’. While this
is not typically among the generators of the Pachner
moves, we find this move to be useful to discuss the
introduction of twists to the dual lattice. Fig. 35(bii)
also shows the inverse of (bi). The effect of these
moves on the dual lattice are shown on the left in
Fig. 35, the effect on the color code lattice is shown
on the right.
In the dual color code lattice, the triangular faces
correspond to physical qubits and the vertices to sta-
bilizers. Thus, the (a) Pachner move changes the
number of qubits on which the four involved stabi-
lizers have support by ±1. This will turn the usually
even plaquettes into odd plaquettes, which has the ef-
fect that x-type and z-type stabilizers no longer com-
Figure 36: Only two of the Pachner moves shown in Fig. 35
are possible in the 6.6.6 color code. Changed or added edges
are colored in purple on the dual lattice. The resulting dis-
location lines are shown here in the dual and primal lattice.
Note that the top lattice on the right side is rotated 30
compared to the rest.
mute. To go back to to a commuting Hamiltonian,
one of the stabilizers has to be removed. If, however,
a series of (a) moves are applied along a line, as shown
in Fig. 36 (a) and Fig. 37 (a), the parity of vertices
along the line gets changed twice and is even in the
end. This means the only odd plaquettes lie at the
terminal points and constitute the twists of the cre-
ated domain wall.
The (b) Pacher move only changes two surrounding
plaquettes, where it either adds (bi) or removes (bii)
an edge. Additionally, it adds or removes a square
stabilizer. Again, these moves can be applied along
a line such that vertices along the line change parity
twice. This restores the parity of all plaquettes to be
even, except for the ones at the terminal points. The
single plaquette at each terminal point changes its
Figure 37: The three Pachner moves from Fig. 35 applied
on the 4.8.8 color code. The resulting dislocation lines are
shown in purple. Note, the dashed line in (bii) indicates the
removed edges. The lattice in the middle on the right side is
rotated 45
compared to the rest.
Accepted in Quantum 2018-10-09, click title to verify 27
Figure 38: A configuration of a branching domain wall ter-
minated by three twists is shown on the dual lattice. On the
left an R twist, on the right a B twist and in the top an
RB twist. Alternatively, this picture can be understood as
showing the fusion between an R and a B twist to form an
RB twist. Fig. 18 depicts the exact same configuration on
the primal lattice.
support to an odd number of qubits and its stabilizers
need to be changed. See Fig. 36 (bi) and Fig. 37 (bi)
and (bii).
Extending and shortening the line along which the
Pachner moves are applied, allows for twists to be
moved and fused together. Consider as an example
Fig. 38, where three twists are shown, an R twist on
the left, a B twist on the right and an RB twist at
the top. This configuration can be obtained by fus-
ing an R and a B twist together. Notice how at the
point where the three domain walls meet, not a single
stabilizer has odd weight. Hence, there is no non-
trivial twist located there. The same configuration
was shown in the main text in Fig. 18.
The way color twists change when moved over color
domain walls can also be visualized using Pachner
moves. In Fig. 39, two crossing domain walls are de-
picted. We can understand the configuration as a pair
of G twists located at the bottom, one of which was
moved over an R domain wall in the middle which
Figure 39: A G twist (bottom), when moved over an R
domain wall (middle) changes into a B twist (top). This
process can be understood by starting with an R domain
wall and then applying Pachner moves which move the G
twist over said R domain wall. Since the lattice along the
R domain wall hosts yellow square stabilizers, the series of
Pachner moves changes such, that after crossing the R do-
main wall, it is terminated by a B twist on a blue vertex.
Fig. 24 depicts the same configuration of twists in the primal
lattice.
Figure 40: Adding a color domain wall changes the color
boundaries. (a) shows a common red color boundary in the
dual lattice. The dashed lines going upwards indicate the
deleted red vertex (plaquette in the primal lattice) that was
removed to create the red color boundary. In (b), a G domain
wall was added to the boundary by applying (bi) Pachner
moves. In (c), the in (b) introduces yellow plaquettes has
been changed to red. And finally, in (d) the lattice was a bit
deformed to make it apparent that the obtained boundary is
a blue boundary. We summarize, a G domain wall changes
a red to a blue color boundary.
transformed in to a B twist at the top. In doing so,
the series of Pachner moves changed from connect-
ing green vertices to connecting blue vertices as soon
as the twist was moved over the domain wall. This
shows, how a G twist changes to a B twist if moved
over an R domain wall. See Fig. 24 for the same con-
figuration in the primal color code lattice.
Next, let study how color domain walls change the
color boundaries of the color code. Consider Fig. 40,
where we show how a red color boundary turns into a
blue color boundary when a G domain wall is added.
The red color boundary corresponds to a single red
vertex (plaquette in the primal lattice), whose stabi-
lizer generators were removed from the code Hamil-
tonian. This is indicated in the figure by the dashed
lines extending to a red vertex far outside the top of
the figure. Adding a G domain wall at the boundary
by applying (bi) Pachner moves, adds new yellow ver-
tices. This transforms the boundary to a blue color
boundary, as is made clear in the figure by first re-
coloring the vertices and then slightly deforming the
lattice.
Figure 41: An R defect line condensed at a red color bound-
ary as drawn on the dual lattice. The color boundary on the
top corresponds to a removed red plaquette, indicated by the
dashed edges. An R domain wall introduced by (bi) Pachner
moves applied along a line can terminate at the red boundary.
Accepted in Quantum 2018-10-09, click title to verify 28
Lastly, let us turn to the condensation of a color
twist at a color boundary. Fig. 41 depicts an R do-
main wall which extends over a red color boundary.
As discussed, the red boundary corresponds to a re-
moved red vertex, as indicated by the dashed edges.
An R domain wall created by (bi) Pachner moves
can be extended such that it terminates at the re-
moved red vertex. This corresponds to the condensa-
tion of the terminating R twist at the red boundary.
Fig. 28 (a) depicts the same configuration of twists in
the primal lattice.
We find the connection between color permuting
twists and Pachner moves is illuminating as it visu-
alizes the algebraic structure of the color permuting
subgroup of the color code symmetry. It also im-
plicitly offers a protocol to braid and condense color
twists through code deformation techniques. Addi-
tionally, the Pachner move construction likely gen-
eralizes to show extensive twists [103] in higher di-
mensional color codes [2931, 95, 104], see also the
supplemental material of Ref. [105].
D Stellated surface codes
As mentioned in Sec. 8 where stellated color codes
were introduced, there also exist similar codes for
the toric code model. We call them stellated sur-
face codes. The triangular surface code introduced
in Ref. [22] can be interpreted as the s = 3 member
of this code family, where s denotes the rotational
symmetry of the code. Stellated surface codes with
odd symmetry s host a toric code twist in their cen-
ter, which is connected to the boundary via a domain
wall, shown in purple in Fig. 42. The c value increases
with increasing s. We find c = 2
2
s
for odd s and
c = 2
4
s
for even s, both approaching 2 for large
values of s.
Figure 42: The three first stellated surface codes with odd
s, shown with a code distance d = 7. Each plaquette hosts
one stabilizer which acts in the z-(x-)basis on the qubit if
the plaquette is colored light (dark) gray where it meets the
qubit. Along the domain wall, indicated by the dashed purple
line, stabilizers act in both bases on different qubits. At the
center of the code a stabilizer acts in the y-basis on the qubit
which it meets with two colors. Left: The s = 3 stellated
surface code is the Yoder-Kim triangular surface code [22].
The s = 5 and the s = 7 stellated toric codes are shown in
the middle and on the right, respectively.
The stellated surface codes with the three smallest
odd s are shown in Fig. 42. We use the convention
of placing physical qubits at vertices and associating
each plaquette with one stabilizer [64]. Light gray
plaquettes correspond to a z stabilizer, which checks
the parity of qubits adjacent to the plaquette in the
z-basis. Similarly, dark gray plaquettes correspond
to x stabilizers. Plaquettes along the purple dashed
domain wall are colored in both shades, dark and light
gray. They also host a single stabilizer each which acts
in different bases; qubits which lie on their light gray
side are measured in the z-basis, the ones on the dark
gray side in the x-basis, and the single qubit where
both shades of gray meet is measured in the y-basis.
References
[1] Xiao-Gang Wen. Quantum field theory of many-
body systems: from the origin of sound to an
origin of light and electrons. Oxford University
Press on Demand, 2004. URL https://books.
google.com/books?id=llnlrfdR4YgC.
[2] Alexei Yu. Kitaev. Fault-tolerant quantum com-
putation by anyons. Ann. Phys., 303:2, 2003.
DOI: 10.1016/S0003-4916(02)00018-0.
[3] Eric Dennis, Alexei Kitaev, Andrew Landahl,
and John Preskill. Topological quantum mem-
ory. J. Math. Phys., 43:4452, 2002. DOI:
10.1063/1.1499754.
[4] J. Preskill. Topological quantum computation,
2017. URL http://www.theory.caltech.edu/
˜
preskill/ph219/topological.pdf. (Chap-
ter 9 of Lecture Notes on Quantum Computa-
tion).
[5] Chetan Nayak, Steven H. Simon, Ady Stern,
Michael Freedman, and Sankar Das Sarma.
Non-Abelian anyons and topological quantum
computation. Rev. Mod. Phys., 80:1083, 2008.
DOI: 10.1103/RevModPhys.80.1083.
[6] Jiannis K Pachos. Introduction to topologi-
cal quantum computation. Cambridge Univer-
sity Press, 2012. URL https://books.google.
com/books?id=XDciVh6bAE0C.
[7] B. M. Terhal. Quantum error correcton for
quantum memories. Rev. Mod. Phys., 87:307,
2015. DOI: 10.1103/RevModPhys.87.307.
[8] B. J. Brown, D. Loss, J. K. Pachos, C. N. Self,
and J. R. Wootton. Quantum memories at finite
temperature. Rev. Mod. Phys., 88:045005, 2016.
DOI: 10.1103/RevModPhys.88.045005.
[9] Earl T Campbell, Barbara M Terhal, and
Christophe Vuillot. Roads towards fault-
tolerant universal quantum computation. Na-
ture, 549(7671):172, 2017. DOI: 10.1038/na-
ture23460.
[10] Alexei Kitaev. Anyons in an exactly solved
model and beyond. Ann. Phys., 321(1):2–111,
2006. DOI: 10.1016/j.aop.2005.10.005.
Accepted in Quantum 2018-10-09, click title to verify 29
[11] Maissam Barkeshli and Xiao-Gang Wen. U(1)×
U(1) o Z
2
Chern-Simons theory and Z
4
parafermion fractional quantum Hall states.
Phys. Rev. B, 81:045323, 2010. DOI:
10.1103/PhysRevB.81.045323.
[12] Hector Bomb´ın. Topological order with a twist:
Ising anyons from an Abelian model. Phys. Rev.
Lett., 105:030403, 2010. DOI: 10.1103/Phys-
RevLett.105.030403.
[13] Salman Beigi, Peter W. Shor, and Daniel
Whalen. The quantum double model with
boundary: Condensations and symmetries.
Commun. Math. Phys., 313:351, 2011. DOI:
10.1007/s00220-011-1294-x.
[14] Alexei Kitaev and Liang Kong. Model for
gapped boundaries and domain walls. Com-
mun. Math. Phys., 313:351, 2012. DOI:
10.1007/s00220-012-1500-5.
[15] Maissam Barkeshli, Parsa Bonderson, Meng
Cheng, and Zhenghan Wang. Symmetry,
defects, and gauging of topological phases.
arXiv:1410.4540, 2014. URL https://arxiv.
org/abs/1410.4540.
[16] Daniel Barter, Jacob C. Bridgeman, and Corey
Jones. Domain walls in topological phases
and the Brauer-Picard ring for Vec(Z/pZ).
arXiv:1806.01279, 2018. URL https://arxiv.
org/abs/1806.01279.
[17] Hector Bomb´ın. Clifford gates by code defor-
mation. New J. Phys., 13:043005, 2011. DOI:
10.1088/1367-2630/13/4/043005.
[18] Maissam Barkeshli, Chao-Ming Jian, and Xiao-
Liang Qi. Twist defects and projective
non-Abelian braiding statistics. Phys. Rev.
B, 87:045130, 2013. DOI: 10.1103/Phys-
RevB.87.045130.
[19] Benjamin J. Brown, Abbas Al-Shimary, and
Jiannis K. Pachos. Entropic barriers for two-
dimensional quantum memories. Phys. Rev.
Lett., 112:120503, 2014. DOI: 10.1103/Phys-
RevLett.112.120503.
[20] Matthew B. Hastings and A. Geller. Reduced
space-time costs Ising dislocation codes and ar-
bitrary ancillas. Quant. Inf. Comp., 15:0962,
2015. URL https://arxiv.org/abs/1408.
3379.
[21] James R. Wootton. A family of stabilizer codes
for D(Z
2
) anyons and majorana modes. J.
Phys. A: Math. Theor., 48:215302, 2015. DOI:
10.1088/1751-8113/48/21/215302.
[22] Theodore J. Yoder and Isaac H. Kim. The sur-
face code with a twist. Quantum, 1:2, 2017.
DOI: 10.22331/q-2017-04-25-2.
[23] B. J. Brown, K. Laubscher, M. S. Kesselring,
and J. R. Wootton. Poking holes and cutting
corners to achieve Clifford gates with the sur-
face code. Phys. Rev. X, 7:021029, 2017. DOI:
10.1103/PhysRevX.7.021029.
[24] Beni Yoshida. Gapped boundaries, group
cohomology and fault-tolerant logical gates.
Ann. Phys., 377:387–413, 2017. DOI:
10.1016/j.aop.2016.12.014.
[25] Sam Roberts, Beni Yoshida, Aleksander Ku-
bica, and Stephen D. Bartlett. Symmetry-
protected topological order at nonzero temper-
ature. Phys. Rev. A, 96:022306, 2017. DOI:
10.1103/PhysRevA.96.022306.
[26] Guanyu Zhu, Mohammad Hafezi, and Maissam
Barkeshli. Quantum origami: Applying fault-
tolerant transversal gates and measuring topo-
logical order. arXiv:1711.05752, 2017. URL
https://arxiv.org/abs/1711.05752.
[27] Ali Lavasani and Maissam Barkeshli. Low
overhead Clifford gates from joint measure-
ments in surface, color, and hyperbolic codes.
arXiv:1804.04144, 2018. URL https://arxiv.
org/abs/1804.04144.
[28] Hector Bomb´ın and Miguel A. Martin-
Delagado. Topological quantum distillation.
Phys. Rev. Lett., 97:180501, 2006. DOI:
10.1103/PhysRevLett.97.180501.
[29] ector Bomb´ın. Gauge color codes: optimal
transveral gates and gauge fixing in topological
stabilizer codes. New J. Phys., 17:083002, 2015.
DOI: 10.1088/1367-2630/17/8/083002.
[30] Aleksander Kubica and Michael E. Beverland.
Universal transversal gates with color codes: A
simplified approach. Phys. Rev. A, 91:032330,
2015. DOI: 10.1103/PhysRevA.91.032330.
[31] Fern H. E. Watson, Earl T. Campbell, Hus-
sain Anwar, and Dan E. Browne. Qudit color
codes and gauge color codes in all spatial di-
mensions. Phys. Rev. A, 92:022312, 2015. DOI:
10.1103/PhysRevA.92.022312.
[32] Earl T. Campbell. The smallest in-
teresting color code, 2016. URL
https://earltcampbell.com/2016/09/26/
the-smallest-interesting-colour-code/.
[33] Aleksander Kubica, Beni Yoshida, and Fer-
nando Pastawski. Unfolding the color code. New
Journal of Physics, 17(8):083026, 2015. DOI:
10.1088/1367-2630/17/8/083026.
[34] Michael Vasmer and Dan E. Browne. Univer-
sal quantum computing with 3D surface codes.
arXiv:1801.04255, 2018. URL https://arxiv.
org/abs/1801.04255.
[35] Sergey Bravyi and Robert onig. Classification
of topologically protected gates for local stabi-
lizer codes. Phys. Rev. Lett., 110:170503, 2013.
DOI: 10.1103/PhysRevLett.110.170503.
[36] Fernando Pastawski and Beni Yoshida. Fault-
tolerant logical gates in quantum error-
correcting codes. Phys. Rev. A, 91:012305, 2015.
DOI: 10.1103/PhysRevA.91.012305.
[37] Tomas Jochym-OConnor, Aleksander Kubica,
and Theodore J Yoder. Disjointness of stabilizer
Accepted in Quantum 2018-10-09, click title to verify 30
codes and limitations on fault-tolerant logical
gates. Phys. Rev. X, 8(2):021047, 2018. DOI:
10.1103/PhysRevX.8.021047.
[38] Paul Webster and Stephen D. Bartlett.
Locality-preserving logical operators in topolog-
ical stabilizer codes. Phys. Rev. A, 97:012330,
2018. DOI: 10.1103/PhysRevA.97.012330.
[39] R. Raussendorf, J. Harrington, and K. Goyal.
A fault-tolerant one-way quantum com-
puter. Ann. Phys., 321:2242, 2006. DOI:
10.1016/j.aop.2006.01.012.
[40] H. Bomb´ın and M. A. Martin-Delgado. Quan-
tum measurements and gates by code defor-
mation. J. Phys. A, 42:095302, 2009. DOI:
10.1088/1751-8113/42/9/095302.
[41] Austin G Fowler. Two-dimensional color-
code quantum computation. Phys. Rev.
A, 83(4):042310, 2011. DOI: 10.1103/Phys-
RevA.83.042310.
[42] Clare Horsman, Austin G. Fowler, Simon De-
vitt, and Rodney Van Meter. Surface code
quantum computing by lattice surgery. New
J. Phys., 14:123011, 2012. DOI: 10.1088/1367-
2630/14/12/123011.
[43] Andrew J. Landahl and Ciaran Ryan-Anderson.
Quantum computing by color-code lattice
surgery. arXiv:1407.5103, 2014. URL https:
//arxiv.org/abs/1407.5103.
[44] Jeffrey CY Teo, Abhishek Roy, and Xiao Chen.
Unconventional fusion and braiding of topo-
logical defects in a lattice model. Phys. Rev.
B, 90(11):115118, 2014. DOI: 10.1103/Phys-
RevB.90.115118.
[45] Beni Yoshida. Topological color code and
symmetry-protected topological phases. Phys.
Rev. B, 91:245131, 2015. DOI: 10.1103/Phys-
RevB.91.245131.
[46] Jacob C. Bridgeman, Stephen D. Bartlett, and
Andrew C. Doherty. Tensor networks with a
twist: Anyon-permuting domain walls and de-
fects in peps. Phys. Rev. B, 96:245122, 2017.
DOI: 10.1103/PhysRevB.96.245122.
[47] Dominic J Williamson, Nick Bultinck, and
Frank Verstraete. Symmetry-enriched topolog-
ical order in tensor networks: Defects, gaug-
ing and anyon condensation. arXiv:1711.07982,
2017. URL https://arxiv.org/abs/1711.
07982.
[48] M. D. Reed, L. DiCarlo, S. E. Nigg, L. Sun,
L. Frunzio, S. M. Girvin, and R. J. Schoelkopf.
Realization of three-qubit quantum error cor-
rection with superconducting circuits. Nature,
482:382, 2012. DOI: 10.1038/nature10786.
[49] R. Barends, J. Kelly, A. Megrant, A. Veitia,
D. Sank, E. Jeffry, T. C. White, J. Mutus,
A. G. Fowler, B. Campbell, Y. Chen, Z. Chen,
B. Chiaro, A. Dunsworth, C. Neill, P. O’Malley,
P. Roushan, A. Vainsencher, J. Wenner, A. N.
Korotkov, A. N. Cleland, and J. M. Martinis.
Superconducting quantum circuits at the sur-
face code threshold for fault tolerance. Nature,
508:500, 2014. DOI: 10.1038/nature13171.
[50] D. Nigg, M. uller, E. A. Martinez,
P. Schindler, M. Hennrich, T. Monz, M. A.
Martin-Delgado, and R. Blatt. Quantum
computations on a topologically encoded
qubit. Science, 345(6194):302–305, 2014. DOI:
10.1126/science.1253742.
[51] A. D. orcoles, Easwar Magesan, Srikanth J.
Srinivasan, Andrew W. Cross, M. Steffen,
Jay M. Gambetta, and Jerry M. Chow. Demon-
stration of a quantum error detection code
using a square lattice of four superconducing
qubits. Nat. Comms., 6:6979, 2015. DOI:
10.1038/ncomms7979.
[52] S. M. Albrecht, A. P. Higginbotham, M. Mad-
sen, F. Kuemmeth, T. S. Jespersen, J. Nyg˚ard,
P. Krogstrup, and C. M. Marcus. Exponen-
tial protection of zero modes in Majorana is-
lands. Nature, 531:206, 2016. DOI: 10.1038/na-
ture17162.
[53] Maika Takita, A. D. C´orcoles, Easwar Magesan,
Baleeg Abdo, Markus Brink, Andrew W. Cross,
Jerry M. Chow, and Jay M. Gambetta. Demon-
stration of weight-four parity measurements
in the surface code architecture. Phys. Rev.
Lett., 117:210505, 2016. DOI: 10.1103/Phys-
RevLett.117.210505.
[54] Norbert M. Linke, Mauricio Gutierrez, Kevin A.
Landsman, Caroline Figgatt, Shantanu Deb-
nath, Kenneth R. Brown, and Christopher Mon-
roe. Fault-tolerant quantum error detection.
Sci. Adv., 3:e1701074, 2017. DOI: 10.1126/sci-
adv.1701074.
[55] H. Bomb´ın, Ruben S. Andrist, Masayuki
Ohzeki, Helmut G. Katzgraber, and M. A.
Martin-Delgado. Strong resilience of topo-
logical codes to depoloarization. Phys. Rev.
X, 2:021004, 2012. DOI: 10.1103/Phys-
RevX.2.021004.
[56] Andrew J. Landahl, Jonas T. Anderson, and
Patrick R. Rice. Fault-tolerant quantum com-
puting with color codes. arXiv:1108.5738, 2011.
URL https://arxiv.org/abs/1108.5738.
[57] Barbara M. Terhal, Fabian Hassler, and
David P. DiVincenzo. From majorana fermions
to topological order. Phys. Rev. Lett., 108:
260504, Jun 2012. DOI: 10.1103/Phys-
RevLett.108.260504.
[58] D. Aasen, M. Hell, R. V. Mishmash, A. Higgin-
botham, J. Danon, M. Leijnse, T. S. Jespersen,
J. A. Folk, C. M. Marcus, K. Flensberg, and
J. Alicea. Milestones toward majorana-based
quantum computing. Phys. Rev. X, 6:031016,
2016. DOI: 10.1103/PhysRevX.6.031016.
[59] S. Plugge, L. A. Landau, E. Sela, A. Altland,
Accepted in Quantum 2018-10-09, click title to verify 31
K. Flensberg, and R. Egger. Roadmap to ma-
jorana surface codes. Phys. Rev. B, 94:174514,
Nov 2016. DOI: 10.1103/PhysRevB.94.174514.
[60] L. A. Landau, S. Plugge, E. Sela, A. Altland,
S. M. Albrecht, and R. Egger. Towards realis-
tic implementations of a majorana surface code.
Phys. Rev. Lett., 116:050501, Feb 2016. DOI:
10.1103/PhysRevLett.116.050501.
[61] D. Litinski, M. Kesselring, J. Eisert, and F. von
Oppen. Combining topological hardware and
topological software: Color code quantum com-
puting with topological superconductor net-
works. Phys. Rev. X, 7:031048, 2017. DOI:
10.1103/PhysRevX.7.031048.
[62] Daniel Litinski and Felix von Oppen. Braiding
by Majorana tracking and long-range CNOT
gates with color codes. Phys. Rev. B, 96:205413,
2017. DOI: 10.1103/PhysRevB.96.205413.
[63] Sergey Bravyi, David Poulin, and Barbara Ter-
hal. Tradeoffs for reliable quantum infor-
mation storage in 2D systems. Phys. Rev.
Lett., 104(5):050503, 2010. DOI: 10.1103/Phys-
RevLett.104.050503.
[64] Xiao-Gang Wen. Quantum orders in an exact
soluble model. Phys. Rev. Lett., 90(1):016803,
2003. DOI: 10.1103/PhysRevLett.90.016803.
[65] Nicolas Delfosse, Pavithran Iyer, and David
Poulin. Generalized surface codes and packing
of logical qubits. arXiv:1606.07116, 2016. URL
https://arxiv.org/abs/1606.07116.
[66] Hector Bomb´ın, Guillaume Duclos-Cianci, and
David Poulin. Universal topological phase of
two-dimensional stabilizer codes. New Jour-
nal of Physics, 14(7):073048, 2012. DOI:
10.1088/1367-2630/14/7/073048.
[67] Arjun Bhagoji and Pradeep Sarvepalli. Equiv-
alence of 2D color codes (without translational
symmetry) to surface codes. arXiv:1503.03009,
2015. URL https://arxiv.org/abs/1503.
03009.
[68] Ben Criger and Barbara Terhal. Noise thresh-
olds for the [[4, 2, 2]]-concatenated toric code.
arXiv:1604.04062, 2016. URL https://arxiv.
org/abs/1604.04062.
[69] Zhenghan Wang. private communication, 2017.
[70] Eric Rowell, Richard Stong, and Zhenghan
Wang. On classification of modular tensor cate-
gories. Commun. Math. Phys., 292(2):343–389,
2009. DOI: 10.1007/s00220-009-0908-z.
[71] Jonathan E. Moussa. Transversal Clifford
gates on folded surface codes. Phys. Rev.
A, 94:042316, 2016. DOI: 10.1103/Phys-
RevA.94.042316.
[72] Daniel Gottesman. Stabilizer Codes and Quan-
tum Error Correction. PhD thesis, California
Institute of Technology, 1997. URL https:
//arxiv.org/abs/quant-ph/9705052.
[73] M. Levin. Protected edge modes without sym-
metry. Phys. Rev. X, 3(2):021009, 2013. DOI:
10.1103/PhysRevX.3.021009.
[74] Maissam Barkeshli, Chao-Ming Jian, and Xiao-
Liang Qi. Classification of topological de-
fects in Abelian topological states. Phys. Rev.
B, 88:241103(R), 2013. DOI: 10.1103/Phys-
RevB.88.241103.
[75] Iris Cong, Meng Cheng, and Zhenghan Wang.
Defects between gapped boundaries in two-
dimensional topological phases of matter. Phys.
Rev. B, 96:195129, 2017. DOI: 10.1103/Phys-
RevB.96.195129.
[76] FJ Burnell. Anyon condensation and its appli-
cations. Annu. Rev. Condens. Matter Phys.,
(0), 2017. DOI: 10.1146/annurev-conmatphys-
033117-054154.
[77] Iris Cong, Meng Cheng, and Zhenghan Wang.
Topological quantum computation with gapped
boundaries. arXiv:1609.02037, 2016. URL
https://arxiv.org/abs/1609.02037.
[78] Iris Cong, Meng Cheng, and Zhenghan Wang.
Universal quantum computation with gapped
boundaries. Phys. Rev. Lett., 119(17):170504,
2017. DOI: 10.1103/PhysRevLett.119.170504.
[79] Netanel H Lindner, Erez Berg, Gil Refael, and
Ady Stern. Fractionalizing majorana fermions:
Non-abelian statistics on the edges of abelian
quantum hall states. Phys. Rev. X, 2(4):041002,
2012. DOI: 10.1103/PhysRevX.2.041002.
[80] Nicolas Tarantino, Netanel H Lindner, and
Lukasz Fidkowski. Symmetry fractionalization
and twist defects. New J. Phys., 18(3):035006,
2016. DOI: 10.1088/1367-2630/18/3/035006.
[81] J. C. Y. Teo. Globally symmetric topolog-
ical phase: from anyonic symmetry to twist
defect. J. Phys., 28(14):143001, 2016. DOI:
10.1088/0953-8984/28/14/143001.
[82] Michael E Beverland, Oliver Buerschaper,
Robert Koenig, Fernando Pastawski, John
Preskill, and Sumit Sijher. Protected gates
for topological quantum field theories. J.
Math. Phys., 57(2):022201, 2016. DOI:
10.1063/1.4939783.
[83] Alexei Kitaev and John Preskill. Topo-
logical entanglement entropy. Phys. Rev.
Lett., 96:110404, 2006. DOI: 10.1103/Phys-
RevLett.96.110404.
[84] Shiying Dong, Eduardo Fradkin, Robert G.
Leigh, and Sean Nowling. Topological entan-
glement entropy in Chern-Simons theories and
quantum Hall fluids. JHEP, 05:016, 2008. DOI:
10.1088/1126-6708/2008/05/016.
[85] Benjamin J. Brown, Stephen D. Bartlett, An-
drew C. Doherty, and Sean D. Barrett. Topo-
logical entanglement entropy with a twist.
Phys. Rev. Lett., 111:220402, 2013. DOI:
10.1103/PhysRevLett.111.220402.
Accepted in Quantum 2018-10-09, click title to verify 32
[86] Zhao Liu, Gunnar oller, and Emil J.
Bergholtz. Exotic non-Abelian topological de-
fects in lattice fractional quantum Hall states.
Phys. Rev. Lett., 119:106801, 2017. DOI:
10.1103/PhysRevLett.119.106801.
[87] Parsa Bonderson, Christina Knapp, and
Kaushal Patel. Anyonic entanglement and topo-
logical entanglement entropy. Ann. Phys., 385:
399, 2017. DOI: 10.1016/j.aop.2017.07.018.
[88] Y.-Z. You and X.-G. Wen. Projective non-
abelian statistics of dislocation defects in a Z
N
rotor model. Phys. Rev. B, 86(16):161107, 2012.
DOI: 10.1103/PhysRevB.86.161107.
[89] Hendrik Poulsen Nautrup, Nicolai Friis, and
Hans J Briegel. Fault-tolerant interface be-
tween quantum memories and quantum proces-
sors. Nature Commun., 8(1):1321, 2017. DOI:
10.1038/s41467-017-01418-2.
[90] Daniel Litinski and Felix von Oppen. Lattice
surgery with a twist: Simplifying clifford gates
of surface codes. Quantum, 2:62, 2018. DOI:
10.22331/q-2018-05-04-62.
[91] H Bomb´ın and MA Martin-Delgado. Topolog-
ical quantum error correction with optimal en-
coding rate. Phys. Rev. A, 73(6):062303, 2006.
DOI: 10.1103/PhysRevA.73.062303.
[92] Michael H Freedman, David A Meyer, and Feng
Luo. Z
2
-systolic freedom and quantum codes.
In Mathematics of quantum computation, pages
287–320. 2002. URL https://books.google.
de/books?id=evPKBQAAQBAJ.
[93] Nicolas Delfosse. Tradeoffs for reliable quantum
information storage in surface codes and color
codes. In 2013 IEEE International Symposium
on Information Theory, pages 917–921. IEEE,
2013. DOI: 10.1109/ISIT.2013.6620360.
[94] N. P. Breuckmann, C. Vuillot, E. T. Campbell,
A. Krishna, and B. M. Terhal. Hyperbolic and
semi-hyperbolic surface codes for quantum stor-
age. Quantum Sc. Tech., 2:035007, 2017. DOI:
10.1088/2058-9565/aa7d3b.
[95] H Bomb´ın and MA Martin-Delgado. Exact
topological quantum order in d=3 and beyond:
Branyons and brane-net condensates. Phys.
Rev. B, 75(7):075103, 2007. DOI: 10.1103/Phys-
RevB.75.075103.
[96] S. Bravyi, M. B. Hastings, and F. Verstraete.
Lieb-Robinson bounds and the generation of
correlations and topological quantum order.
Phys. Rev. Lett., 97:050401, 2006. DOI:
10.1103/PhysRevLett.97.050401.
[97] Xie Chen, Zheng-Cheng Gu, and Xiao-Gang
Wen. Local unitary transformation, long-
range quantum entanglement, wave function
renormalization and topological order. Phys.
Rev. B, 82:155138, 2010. DOI: 10.1103/Phys-
RevB.82.155138.
[98] Sergey B Bravyi and A Yu Kitaev. Quan-
tum codes on a lattice with boundary.
arXiv:9811052, 1998. URL https://arxiv.
org/abs/quant-ph/9811052.
[99] Nicolas Delfosse, Pavithran Iyer, and David
Poulin. A linear-time benchmarking tool for
generalized surface codes. arXiv:1611.04256,
2016. URL https://arxiv.org/abs/1611.
04256.
[100] H. Bomb´ın. Topological subsystem codes. Phys.
Rev. A, 81(3):32301, 2010. DOI: 10.1103/Phys-
RevA.81.032301.
[101] Udo Pachner. P. L. homeomorphic manifolds
are equivalent by elementary shellings. Eu-
rop. J. Combinatorics, 12:129, 1991. DOI:
10.1016/S0195-6698(13)80080-7.
[102] Mikio Nakahara. Geometry, Topology and
Physics. Institute of Physics, 2003. URL
https://books.google.com/books?id=
cH-XQB0Ex5wC.
[103] Andrej Mesaros, Yong Baek Kim, and Ying
Ran. Changing topology by topological de-
fects in three-dimensional topologically ordered
phases. Phys. Rev. B, 88:035141, 2013. DOI:
10.1103/PhysRevB.88.035141.
[104] H. Bomb´ın and M. A. Martin-Delagado.
Topological computation without braiding.
Phys. Rev. Lett., 98:160502, 2007. DOI:
10.1103/PhysRevLett.98.160502.
[105] Benjamin J. Brown, Naomi H. Nickerson, and
Dan E. Browne. Fault-tolerant error correction
with the gauge color code. Nat. Commun., 7:
12302, 2016. DOI: 10.1038/ncomms12302.
Accepted in Quantum 2018-10-09, click title to verify 33