Analysing correlated noise on the surface code using adaptive
decoding algorithms
Naomi H. Nickerson
1
and Benjamin J. Brown
2,3
1
Quantum Optics and Laser Science, Blackett Laboratory, Imperial College London, Prince Consort Road, London SW7 2AZ,
United Kingdom
2
Niels Bohr International Academy, Niels Bohr Institute, Blegdamsvej 17, 2100 Copenhagen, Denmark
3
Centre for Engineered Quantum Systems, School of Physics, University of Sydney, Sydney, New South Wales 2006, Australia
April 3, 2019
Laboratory hardware is rapidly pro-
gressing towards a state where quantum
error-correcting codes can be realised. As
such, we must learn how to deal with
the complex nature of the noise that may
occur in real physical systems. Single
qubit Pauli errors are commonly used to
study the behaviour of error-correcting
codes, but in general we might expect the
environment to introduce correlated er-
rors to a system. Given some knowledge
of structures that errors commonly take,
it may be possible to adapt the error-
correction procedure to compensate for
this noise, but performing full state to-
mography on a physical system to analyse
this structure quickly becomes impossible
as the size increases beyond a few qubits.
Here we develop and test new methods
to analyse blue a particular class of spa-
tially correlated errors by making use of
parametrised families of decoding algo-
rithms. We demonstrate our method nu-
merically using a diffusive noise model. We
show that information can be learnt about
the parameters of the noise model, and ad-
ditionally that the logical error rates can
be improved. We conclude by discussing
how our method could be utilised in a
practical setting blue and propose exten-
sions of our work to study more general
error models.
1 Introduction
The success of scalable quantum computation de-
pends on our ability to diagnose and repair errors
incident to the underlying physical qubits [1, 2,
3, 4, 5, 6, 7] of the system. To this end quantum
error-correcting codes [8, 9, 10, 11, 12, 13, 14]
have been developed to allow us to correct for
a finite density of physical errors that might
otherwise irreversibly disturb a quantum com-
putation. Of particular interest are topological
codes [10, 11], which are experimentally viable
due to their local structure and low-weight syn-
drome measurement operations [15, 16].
Quantum error-correcting codes are commonly
studied with a simple error model where it is
assumed that each physical qubit is subject to
some rate of local noise that is independent of the
state of other nearby qubits and its local environ-
ment. Identifying thresholds [11] and finding the
resource overheads [17, 18, 19] of different codes
using simple error models can provide a proof of
principle that a code can withstand local noise,
and offers a standardised benchmark to uniformly
compare different error-correcting codes with one
another [20, 21]. However, when we consider
some realisation of a quantum error-correcting
code, where we take into account physical fea-
tures of its underlying hardware [22, 23, 24], an
independent noise model is unlikely to be real-
istic. Instead, we might anticipate that corre-
lated errors [25, 26, 27, 28, 29, 30, 31] could oc-
cur, where the probability of a qubit suffering an
error depends on its neighbouring qubits having
also suffered errors.
Correlated noise has been identified in a num-
ber of physical systems that might serve to re-
alise quantum error-correcting codes. For exam-
ple, external fields and proximity effects in solid
state systems [30, 32] introduce correlated errors,
as do quantum codes that are coupled to a bo-
son environment [33, 34, 35, 36, 37, 13, 31, 38].
Correlated errors have also been identified when
Accepted in Quantum 2019-03-26, click title to verify 1
arXiv:1712.00502v4 [quant-ph] 1 Apr 2019
non-Markovian effects are modelled [26, 39]. It
has also been seen that correlated errors can be
introduced by certain schemes of quantum error-
correction. This has been observed, for instance,
in Ref. [19]. Correlated errors also occur when
considering syndrome measurements under the
circuit-based noise model [40, 41, 42, 43, 44], and
while applying transversal logical gates to quan-
tum error-correcting codes that are not in their
code space [45]. See also recent work. [46].
The aim of this Manuscript is to develop nu-
merical methods to characterise correlated error
models. Specifically, we develop families of de-
coding algorithms, each member of which is de-
signed to correct for a particular correlated error
model. We show that we can use a decoder fam-
ily, which we will refer to as an adaptive decoder,
to analyse a source of noise with some unknown
parameter by comparing the logical error rates
of different members of a decoder family. We
perform simulations to demonstrate our meth-
ods explicitly by adapting Edmonds’ minimum-
weight perfect matching algorithm [47, 48] to
identify string-like correlated errors introduced to
the toric code [10]. Gaining information on the
structure of correlations in noise can be used as a
diagnostic tool to minimize the effect of correla-
tions at the hardware level. It may also be used
to improve classical post-processing. We show
that we can use our adaptive scheme to find a
decoder with an improved threshold to deal with
the physical error model.
There are a number of methods and results
in the literature that look to determine correla-
tions that may emerge in fault-tolerant quantum
hardware. In Ref. [49] Combes et al. showed
that syndrome data can be used to discriminate
between different correlated noise processes act-
ing on small quantum codes including the classi-
cal repetition code and the five-qubit code [50].
Our results build on this work by examining
different codes and we give numerical evidence
that these ideas are applicable to degenerate
codes [51]. Moreover, the use of efficient decoding
algorithms that we propose make our diagnostic
techniques scalable to systems of large size. We
point out that recently, during the preparation of
the present manuscript Ref. [52] showed that it is
possible to monitor time-dependent parameters
of uncorrelated noise models to improve thresh-
old error rates using methods similar to ours.
Other methods have also been found to de-
termine correlated errors that occur during the
error correction process. In Ref. [42] the au-
thors simulate a circuit-level error model to di-
rectly calculate the likelihood that correlated er-
rors occur while syndrome data is measured.
The information obtained with this procedure is
used to improve decoding schemes. Further, in
Ref. [53], drifts in experimental parameters of a
small repetition code are identified and corrected
by measuring changes in the rate at which error-
detection events occur. See also Ref. [54] that
recently became available. These works make
use of local syndrome data to determine differ-
ent parameters in the error model. In contrast,
we make use of a decoding algorithm that uses
global syndrome data to distinguish more generic
correlated error models where error events may
correlate over a large number of adjacent qubits.
We also remark on previous work to specialise
decoders to account for known correlated errors.
It has been shown [55, 56, 57, 58, 59, 60, 61,
62, 63] that we can obtain improved thresholds
by using a decoder that specifically accounts for
known correlations that occur under depolarising
noise. Hutter and Loss have also identified that
specialised decoders can be used to correct for
known correlated errors introduced by a thermal
environment [29]. The present work extends these
ideas by demonstrating that we can use learned
knowledge to calibrate a decoder to give improved
thresholds without complete knowledge of the er-
ror model a priori.
The remainder of the Manuscript is structured
as follows. In Section 2 we introduce the toric
code and develop decoding algorithms for gen-
eral correlated errors. In Section 3 we show that
the generalised decoders we present can be used
to find signatures that distinguish correlated er-
rors. Section 4 shows that we can use error model
data to obtain improved decoding schemes. In
Sec. 5 we give a discussion on how we might ap-
ply our scheme to the practical development of
fault-tolerant quantum hardware. We conclude
in Section 6 by discussing potential directions to
develop our methods in order to diagnose more
general forms of correlated error.
Accepted in Quantum 2019-03-26, click title to verify 2
Figure 1: The toric code with qubits arranged on the
edges of a sqaure lattice of size L = 7. (a) A plaquette
operator is shown in blue. (b) A star operator is shown
in orange. (c) and (d) show logical operators Z
1
and
X
1
, respectively.
2 The toric code and error models
We study correlated errors acting on the well-
studied toric code model. In this Section we
briefly review the toric code, and introduce the
notation we use to describe correlated error
events. We go on to describe the minimum-weight
perfect-matching (MPWM) algorithm that we
will use to analyse correlated errors.
2.1 The toric code
The toric code [10] is a quantum error-correcting
code of n = 2L
2
qubits, arranged on the edges on
a square lattice of L × L vertices with periodic
boundary conditions, as shown in Fig. 1. The
code space of the toric code is described by its
stabilizer group, S [64]. The stabilizer group is
an Abelian subgroup of the Pauli group, P, where
the code states of the toric code, |ψ
j
i, are the
common +1 eigenspace of all elements S
k
S
S
k
|ψ
j
i = (+1)|ψ
j
i, j, k. (1)
Elements of the stabilizer group are commonly
known as stabilizers.
The stabilizer group of the toric code consists
of two types of stabilizers, star operators, A
v
, and
plaquette operators, B
f
, defined as follows
A
v
=
Y
j3v
X
j
, B
f
=
Y
jf
Z
j
, (2)
where j is an index over the qubits and X
j
and
Z
j
operators are Pauli-X and Pauli-Z operators
acting on qubit j. The set f are the qubits on
the boundary of face f and j is the set of ver-
tices v at either end of the edge that supports
qubit j. The stabilizer group consists of one star
operator for each vertex, v, and one plaquette
operator for each face f. Colloquially, a star op-
erator is the product of Pauli-X operators acting
on all the qubits on edges incident to vertex v,
and a plaquette operator is the product of Pauli-
Z operators acting on all the edges that bound
faces of the lattice, f. We show pictures of a pla-
quette and a star operator in Figs. 1(a) and (b),
respectively.
The code space of the toric code encodes two
logical qubits that are acted upon by logical oper-
ators L, whose elements are Z
j
and X
j
with j =
1, 2. The logical operators are non-contractible
loops of Pauli-Z or Pauli-X operators acting on
the primal or dual edges of the lattice respec-
tively. One pair of these logical operators, X
1
and Z
1
, is shown in Figs. 1(c) and (d).
Quantum error-correcting codes are designed
to identify and correct errors suffered by the code.
We consider error operators E P. The proce-
dure for error correction involves measuring the
generators of the stabilizer group. Stabilizers
commute with the logical operators of the stabi-
lizer code, and as such we do not disturb encoded
logical information when stabilizer measurements
are made. The outcomes of the stabilizer mea-
surements can be used to identify the error as
the elements of the stabilizer group S
j
that do
not commute with E will return a 1 outcome.
We see this with the eigenvalue equation
S
k
E|ψ
j
i = ES
k
|ψ
j
i = (1)E|ψ
j
i, j, k. (3)
We refer to elements of the generating set of S
that return 1 eigenvalues as defects. The col-
lection of measurement outcomes of the selected
stabilizer generators is known as the error syn-
drome. In the case of the toric code that we study
here we use the star and plaquette operators in
Eqn. 2 to represent the canonical set of stabilizer
generators.
We require a decoder, a classical post-
processing algorithm that interprets syndrome in-
formation, to estimate the most likely error that
might have occurred on the lattice. A decoder
will subsequently return a correction operator, C,
to attempt to reverse error E. Specifically, we
seek a decoder such that the value
P = prob(CE S), (4)
Accepted in Quantum 2019-03-26, click title to verify 3
is optimised. To do so, the decoder must make
use of some knowledge of the type of errors that
commonly occur in the system. In the remain-
der of this Section we describe general correlated
noise models acting on the toric code, the syn-
dromes they give rise to, and the decoder we use
to interpret the syndrome information.
2.2 Error events
In the previous Subsection we introduced error
operators E P. However, we wish to design
a decoder which takes into account more infor-
mation about the structure of an error operator.
Here we make rigorous the concept of a structured
error model by considering an error operator as a
series of error events.
To study the performance of quantum error-
correction procedures against correlated errors we
generate random Pauli errors of the form
E =
Y
j
e
(1σ
j
)/2
j
, (5)
where operators e
j
P describe a set of error
events that may occur through the noise chan-
nel. Error event e
j
occurs with likelihood p
j
such
that σ
j
= 1 with probability p
j
and σ
j
= +1
otherwise.
Typically quantum error-correcting codes are
tested using an independent and identically dis-
tributed(i.i.d.) noise model. For instance, the
error events of the bit-flip noise model are sin-
gle qubit Pauli errors acting on each qubit of the
code i.e. e
j
= X
j
for all j that index the qubits of
the system. Each error event in the model occurs
with uniform probability p
j
= p for all j. Al-
ternatively, the depolarising noise model is some-
times considered, where similarly, error events e
j
act uniformly on all qubits j of the system, i.e.
p
j
= p, and then error events can take values X
j
,
Y
j
or Z
j
with uniform probability 1/3. In our
analysis that follows we will consider more gen-
eral bit-flip error models with error events that
are supported on multiple qubits. The restriction
to bit-flip errors simplifies the problem of decod-
ing as all of the stabilizer defects are detected by
the plaquette operators.
2.3 Decoding with the minimum-weight per-
fect matching algorithm
In the next Section we generalise the well-studied
MWPM decoding algorithm [11]. For numerical
Figure 2: Defects on the lattice (left) are mapped onto
a graph with weighted edges (right) where the weights
correspond to the probabilities that its adjacent pair of
defects were created by an event that corresponds to
errors along that edge.
studies we make use of the Blossom-V implemen-
tation of Edmonds MWPM algorithm [47] due to
Kolmogorov [48]. This decoding algorithm is par-
ticularly well suited to the diffusive noise model
we will introduce in the next Section.
The algorithm due to Edmonds takes as input
a graph with weighted edges and returns a match-
ing graph, i.e. a graph where each vertex has one
and only one incident edge, such that the sum
of the weights of all the edges are minimal. We
denote edges, e, by pairs of vertices of the graph
such that e = (u, v) where u and v denote ver-
tices of the graph. We also denote by W (e) the
weight assigned to edge e.
To apply the MWPM algorithm to the prob-
lem of decoding we assign to each stabilizer de-
fect measured by a plaquette operator a vertex
of the matching graph. Ideally we then construct
a complete weighted graph where we weight each
edge with the function
W (e) = log(prob(e)), (6)
where prob(e) is the probability that the error
model introduced an error event, or a combina-
tion of error events, such that defects u and v of
edge e occurred. In general evaluating W (e) may
be a complex summation of exponentially many
terms. We will typically approximate this value
by calculating only low-order terms of the sum-
mation. We summarise the construction of the
complete graph in Fig. 2.
To find a correction operator we apply the Ed-
monds MWPM algorithm to the generated graph
to find a matching graph where the edges corre-
spond to a collection of error events that is likely
to have caused the error syndrome. We use this
information to find correction operator C.
Accepted in Quantum 2019-03-26, click title to verify 4
We remark that for the well-studied case of
i.i.d. noise we will typically have W (e) =
log(prob(e)) d(u, v) where d(u, v) is the sep-
aration of vertices u and v of edge e on the lat-
tice [11]. Throughout this work we consider the
Manhattan distance where
d(u, v) = d
x
+ d
y
, (7)
where d
x
(d
y
) is the separation between u and v
along the x(y) direction on the lattice.
We refer to the MWPM decoder where W (e)
d(u, v) which is designed to deal with i.i.d. noise
model as the standard decoder. Other work has
also considered introducing an extra logarithmic
term to account for degeneracy in errors that
cause defects at vertices u and v [65, 59], but
for simplicity we do not use this extension here.
In this work we will generalise the construction of
the MWPM decoder by assigning different func-
tions to W (e) depending on the anticipated cor-
related noise model.
The MWPM decoder will find a single error
that was likely to have caused a given syndrome
with respect to an error model that is deter-
mined by W . We expect that our methods can
be improved using maximum-likelihood decod-
ing schemes that identify the most-likely equiv-
alence class of errors that caused a given syn-
drome [65, 55, 56, 66]. However, in general,
few efficient maximum-likelihood decoding algo-
rithms are known [66]. Here we assume that iden-
tifying a single error that was likely to cause the
syndrome is a good approximation of maximum-
likelihood decoding. This has been demonstrated
to be the case for the independent and identically
distributed noise model in Ref. [11, 67].
3 Noise characterisation
In this Section we consider how to adapt the
MWPM decoding algorithm, such that it can
be parametrised to target different noise models.
This parameterisation allows us to define a family
of decoders, which we call an adaptive decoder.
Specifically, we consider a MWPM decoder where
edges are assigned weights by the function W
λ
(d)
where d is the distance between the two defects,
and the parameter λ denotes different members
of the decoder family. In general, λ can represent
many variables, but here, for simplicity, we will
consider only integer values. We show that we
Figure 3: Error events of the ballistic noise model for
ξ = 3. The model introduces error strings along vertical
and horizontal lines that introduce pairs of stabilizer de-
fects that are separated by distance ξ. (a) An error event
acting on a horizontal edge marked by a square box. A
bit flip is introduced to the edge in the square box, and
the ξ 1 horizontal edges immediately below the edge
where the error event occurs. (b) An error event acting
on a vertical edge marked by a square box. The error
event introduces a bit flip on the vertical edge and the
ξ 1 vertical edges to the immediate right of the edge
where the event occurs.
can distinguish different noise models by study-
ing logical error rates as a function of λ with an
adaptive decoder. We find that, for certain noise
models with unknown parameters, the logical er-
ror rates, or thresholds of different decoders can
allow us to deduce the noise parameters.
We first consider a simple example noise model
to analytically justify the use of logical error rates
to learn noise parameters. We will then test two
different adaptive decoders, which we will define
shortly, to analyse a diffusive noise model using
numerical experiments.
3.1 An illustrative example of the diagnostic
tools
We first develop some intuition for our diagnos-
tic method using a simple correlated noise model,
that we call the ballistic noise model, by quoting
some results that are shown in Appendix A. We
will argue that improved logical error rates are
achieved with a targeted decoder that uses in-
formation about the noise model in the regime
where the error rate is small. We use this exam-
ple to motivate the simulations we conduct in the
remainder of the paper.
In the ballistic error model, error events form
‘straight lines’ of Pauli errors with a fixed weight,
Accepted in Quantum 2019-03-26, click title to verify 5
Figure 4: An error caused by the ballistic noise model in
the path-counting regime. The noise model introduces
strings of length ξ = L/4. A standard decoder will have
difficulty dealing with this error configuration as there
are two inequivalent correction operators that have the
same weight. Conversely the targeted decoder which is
designed to anticipate errors that introduce defects that
are separated by a distance ξ will correct errors from this
error model with high probability.
ξ. More specifically, the error event associated to
vertical(horizontal) qubit edge j introduces a bit-
flip to edge j and the subsequent ξ 1 edges that
are located immediately to the right of edge j
(below edge j). Each error event occurs on edge
j with uniform probability p
j
= p. We show ex-
amples of error events of the ballistic noise model
where ξ = 3 in Fig. 3.
In Appendix A we show that for arbitrary ξ 3
in the regime where p is very low, otherwise
known as the path-counting regime [11, 18], the
logical error rates can be improved by using a tar-
geted decoder. It follows that we can construct
a family of two different decoders to distinguish
i.i.d. noise from the ballistic noise model for fixed
ξ 3 by comparing the logical error rates of
the two different decoders. The two MWPM de-
coders of this family are the following, the first is
the standard decoder, where the matching graph
is constructed with edges weighted according to
the separation of stabilizer defects. This decoder
seeks the lowest-weight correction operator. The
second decoder in the family is the targeted de-
coder, which anticipates the ballistic noise model
and assigns weight a to edges where stabilizer de-
fects are separated by distance , and otherwise
assigns very high weights to edges of the match-
ing graph.
Consider the ballistic noise model with ξ = L/4
in the path-counting regime where 1/p 4ξL.
In this example the common error configurations
that will cause either the standard decoder or the
targeted decoder to fail are those where two non-
overlapping error events occur on the same row
or column. However, there are far fewer non-
overlapping error configurations that will cause
the targeted decoder to fail compared with the
standard decoder. We give an example error con-
figuration in Fig. 4 caused by the ballistic noise
model that will be corrected using the targeted
decoder, but will not be corrected by the stan-
dard decoder.
For the case of the error configuration shown
in Fig. 4, the standard decoder will construct
a graph that will assign equal probabilities to
two inequivalent correction operators, C
1
and C
2
,
independent of L, where C
1
repairs the error,
C
1
E S, and the other correction is such that
C
2
E 6∈ S. The standard decoder will struggle to
deal with this case as both C
1
and C
2
will have
weight L/2 and as such, the standard decoder
will not be able to identify the error events that
have caused the given syndrome. In contrast, the
targeted decoder that is designed to identify the
error events that introduce defects separated by
distance ξ, and will assign very low weights to
the edges of the set that correspond to the cor-
rect correction operator, and very high weights to
the other edges.
One can count the number of non-overlapping
error configurations using Eqn. (11). There are
exactly L(L 2ξ + 1)/2 error configurations of
two error events where ξ = L/4 such that there
are are exactly L/2 bit flips along a given row.
Each of these error configurations will cause the
standard decoder to fail with probability 1/2 in-
dependent of the size of the system. However,
of these non-overlapping configurations, there
are only 6ξ = 3L/2 error configurations where
the targeted decoder will fail with probability
1/2, other non-overlapping error configurations
of weight L/2 will be decoded successfully by the
targeted decoder.
Having shown that there are O(L
2
) low-weight
error configurations that will cause the standard
decoder to fail for the ballistic noise model, com-
pared with only O(L) low weight configurations
that will cause the targeted decoder to fail, we
can expect to be able to identify the ballistic
noise model by comparing logical error rates of
the noise source with respect to the two different
decoding algorithms of the family by taking sys-
tem of suitably large size. In general, we show
in Appendix A that R, the ratio of the failure
Accepted in Quantum 2019-03-26, click title to verify 6
probability of the targeted decoder and the failure
probability of the standard decoder in the path
counting regime calculated with respect to the
ballistic noise model, vanishes exponentially with
L for constant ξ 3. Specifically, we find
R
ξ + 1
2
2
ξ
L/2ξ
. (8)
To this end we can expect to easily identify the
ballistic noise model in the path counting regime
by comparing the logical error rates of the two
decoders of the family.
3.2 Diffusive noise
We now consider a diffusive error model where
each error event introduces a continuous random
walk of bit-flip errors that extends across a num-
ber of nearby qubits such that two defects appear
at its end points. We use this error model to nu-
merically demonstrate our diagnostic method.
We index error events e
f
with the faces of the
lattice f. Event e
f
will occur with probability
p
f
= p where p is uniform for all faces. Given
event e
f
occurs, a random walk is generated for a
fixed number of steps, ξ L, between adjacent
faces of the lattice. At a given step the walk di-
rection is chosen uniformly from the four adjacent
faces. Each edge, indexed j, that is crossed dur-
ing the walk an integer w times will suffer error
X
w
j
. An error event of this kind generates a sin-
gle error chain, that is, where only two syndrome
defects are created at the initial and final face of
the walk.
Finally, we point out that while we observe the
behaviour in the case of odd and even values of
ξ to be qualitatively similar, we do not compare
odd and even values of ξ directly, as we observe
there to be a slight discrepancy in their perfor-
mance. We clearly see this effect in Fig. 5. We
speculate that this is due the fact that, unlike
the odd case, for even ξ there is a non-zero prob-
ability that error events are members of the sta-
bilizer group, and therefore act trivially on the
code space.
3.3 Evaluating ξ for the diffusive noise model
using the single-weight adaptive decoder
Here we demonstrate numerically that using
adaptive decoding the unknown parameter ξ of
the diffusive noise model can be identified. We
3
5
7
9
ξ
1
2 4 6 8 10 12
0.01
0.02
0.05
0.10
λ
p
th
a) Threshold dependence on decoder parameter
b) Example thresholds for 𝜉=3
0.04 0.05 0.06 0.07 0.08 0.09 0.10
0.5
0.6
0.7
0.8
0.9
1.0
error
successrate
𝝀=2
𝝀=3
0 2 4 6 8 10
0
20 000
40 000
60 000
80 000
100 000
!="
W(d)
#="
0 5 10 15 20 25
0.5
0.6
0.7
0.8
0.9
1.0
Decoding success rate
L = 100, xi = 2, p = 0.11, efferr = 0.0770453
Decoder parameter, "
1
3
5
7
9
0 5 10 15 20
0.00
0.02
0.04
0.06
0.08
0.10
p
th
ξ
2
12
30
40
0 10 20 30 40
0
2
4
6
8
Correlation length,
Optimal decoder parameter,
*
L
30
40
50
60
i)
ii)
Decoder weighting function
Error event rate, p
3
5
7
9
ξ
1
2 4 6 8 10
0.01
0.02
0.05
0.10
λ
p
th
Defect separation, 𝑑
𝑊(𝑑)
𝝀
Δ
Figure 5: Threshold error rates as a function of the de-
coder parameter λ under a diffusive noise model that is
characterised by parameter ξ. (a) Solid lines show the
results for ξ = 3, 5, 7 and 9. Dashed line indicates the
threshold error rate for the noise model where ξ = 1.
Each value of ξ shows a unique behavior. Each thresh-
old value is obtained by repeated simulation of the noisy
lattice and decoding process to identify a crossing point.
Two threshold examples for a correlation parameter of
ξ = 3 are shown in (b), corresponding to the points
labelled (i) and (ii) in (a). Each point in these plots
represents the aggregated result of at least 20,000 indi-
vidual simulations.
Accepted in Quantum 2019-03-26, click title to verify 7
use an adaptive decoder based on the MWPM
decoding algorithm that we refer to as the single-
weight adaptive decoder. The weighting func-
tion for this decoder assigns low weights to edges
of the matching graph with length λ, and to
all other edges assigns high weights that scale
with the separation of pairs of stabilizer defects.
Specifically, it assigns edge weights with the func-
tion
W
λ
(d) =
(
d d = λ,
d otherwise,
(9)
where d is the Manhattan distance between two
defects, with L. The weighting function for
the single-weight adaptive decoder is shown in
the inset of Fig. 5(a). We will see that the logical
failure rate of the single-weight decoder depends
on the choice of λ relative to ξ.
We numerically simulate this scenario for in-
teger values of λ between 1 and 10 and values of
ξ = 1, 3, 5, 7, 9. To obtain the threshold error rate
we plot the logical error rate P
fail
as a function
of p for at least three large system sizes. To min-
imise the effects of simulating finite size systems
we study systems where L
ξ given that we
expect random walks to typically achieve walks
of length
ξ. The value of p for which P
fail
coincides for all L determines p
th
. We find the
crossing point by fitting the data to the function
P
fail
= A
0
+ A
1
x + A
2
x
2
where x = (p p
th
)L
1
and A
j
, p
th
and µ are constants to be evalu-
ated. We show an example threshold plot for
ξ = 3 using the single-weight adaptive decoder
with λ = 2, 3 in Fig. 5(b).
We remark on the clear trend in Fig. 5(a) where
the threshold presented as a function of p tend to
decay quickly with ξ. This decay is partly due to
the fact that the weight of the error increases with
ξ even while p remains constant. A comparison
of the effective single qubit error rates are shown
In fact, given good knowledge of the value of ξ
the threshold remains relatively high even in the
presence of the spatial correlations. We discuss
this further in Sec. 4.
Having calculated the threshold for each error
model using a range of decoders from the family,
we go on to estimate the value of ξ by plotting
p
th
as a function of λ for each noise model, as
shown in Fig. 5(a). We observe that each value
of ξ generates a distinct plot as λ is varied. We
can therefore regard this as a signature of a given
noise model, as such, we refer to this plot as a
a) Parameter estimation
b) Continuous 𝝀* estimation
0 1 2 3 4 5 6 7
0.010
0.015
0.020
0.025
0.030
0.035
0.040
0.045
λ
p
th
●●
Discrete λ
*
Continuous λ
*
3 5 7 9 11
1
2
3
4
5
ξ
λ
*
0.08
0.02
0.04
0.06
Figure 6: Estimating ξ values using signature plots in
Fig. 5(a). (a) Estimates of λ
are shown. The orange
points show the discrete values of λ for which the adap-
tive decoder has the greatest threshold error rate. A
continuous estimation is made by fitting a parabola to
the data points of the signature plot with odd λ close to
the peak λ value (red data points). This fitting is shown
in (b). We observe that λ
grows with ξ with the trend
λ
1.3 · ξ
0.39
. We show the fitting with the dotted
blue line in (a).
Accepted in Quantum 2019-03-26, click title to verify 8
signature plot. We now investigate how we can
use the signature plot to estimate the unknown
parameter of the noise model, ξ and determine
other features of the noise model.
Studying Fig. 5(a) we first remark on the dis-
crepancy between the data obtained for even and
odd values of λ. We have considered only odd
length error chains, which means single error
events will never introduce defects separated by
an even distance. As such, a decoder with even
λ will predominantly be correcting for second or-
der error events, which we would expect to have a
poor performance. This effect is indeed observed
in the data. When the value of λ is odd, even
when it does not match ξ the decoder is prefer-
entially correcting for first-order error events. It
is interesting to see that this feature of the noise
model, namely that the noise model only intro-
duces error strings of odd length, can be readily
identified from the signature plot.
In Fig. 6 we plot estimates for the value of λ
for which the threshold error rate is maximal. We
denote this value λ
. We determine λ
using two
rudimentary methods. First of all we estimate λ
as the discrete value of λ for which the threshold
is maximal. We show these discrete data points
in orange in the Figure. Notably, for the rela-
tively small values of ξ we study, we observe that
λ
increases with ξ in one discrete step. This ob-
servation indicates that λ
is sensitive to changes
in the value of ξ.
Ideally, we would like to determine ξ precisely
using the signature plot. However, using the
method for estimating λ
we have suggested so
far, it is difficult to resolve ξ from λ
as there are
error models with different values of ξ that return
the same value λ
. To deal with this we refine the
estimate of λ
by using more data points from the
signature plot. In the Fig. 6(b) we fit a quadratic
curve to the odd λ data points close to λ
, which
are shown in Fig. 5. The maximal values of the
quadratic fittings are shown in Fig. 6(a) with red
data points. We observe that the maximal values
increase like λ
1.3 · ξ
0.39
which is plotted in
blue against the acquired data points. We might
expect a relationship λ
ξ that represents the
average distance covered by an error event that
performs a uniform random walk. We speculate
that this discrepancy in the fitting is due to mi-
croscopic details and small size effects. Nonethe-
less, our numerical results show that the value of
λ
, which is found using the signature plot, esti-
mates the unknown parameter ξ precisely.
3.4 Estimating ξ using the Gaussian decoder
One might argue that the adaptive decoder we
used in the previous section uses too much knowl-
edge of the error model to be applicable in any
real system, since we use the fact that the random
walk is of a fixed length. In this section we show
that a similar result is achieved with a more ‘gen-
eral purpose’ adaptive decoder, which could ap-
ply to many more possible noise models. We call
this the Gaussian decoder, as it assigns weights
to the edges of the matching graph according to
a Gaussian function based on the separation of
two defects
W
λ
(d) = d
10
4
9999 e
(dµ)
2
2σ
2
(10)
where d is the separation of the two defects mea-
sured via the Manhattan distance on the lattice,
µ = λ and σ = λ/2. A schematic of the weight-
ing function of the Gaussian adaptive decoder is
shown in the inset of Fig. 7. The decoder is de-
signed to assign bias to identify common error
events that cause defects that are separated by a
distance that is on average d λ which we ex-
pect from the diffusive noise model. Errors that
are not consistent with a diffusive random walk
are assigned probabilities that are proportional to
their separation as we would using the standard
decoder.
In the previous section we produced signature
plots by computing the threshold error rate for
each decoder parameter λ, and then using the
result to estimate the correlation length ξ. In
practice it will be simpler to use logical failure
rates directly, rather than thresholds, as they can
be computed for a single code of fixed size, and
a single value of physical error rate. As such,
for this case, we use logical error rates instead of
threshold error rates to find a signature plot.
At the top of Fig. 7(a) we generate a signature
plot using the logical failure rate of the adaptive
decoder as a function of λ for different odd values
of ξ ranging from ξ = 3 to ξ = 51, for a lattice
size of L = 100 such that
ξ L. Only the
odd values of λ are shown. We see that, as in the
previous Subsection, the peak logical success rate
varies with ξ. In Fig. 7(b) we aim to estimate ξ
using the signature plot. Two different estimates
Accepted in Quantum 2019-03-26, click title to verify 9
0 5 10 15 20 25 30
0.4
0.5
0.6
0.7
0.8
0.9
1.0
λ
Decoding success rate
L = 100, xi = 3, p = 0.07, efferr = 0.0697621
a) Logical error rate dependence on decoder parameter 𝝀
b) Parameter estimation
0 2 4 6 8 10
0
20 000
40 000
60 000
80 000
100 000
anyon%separa*on,%d
!="
W(d)
#="
0 5 10 15 20 25
0.5
0.6
0.7
0.8
0.9
1.0
Decoding success rate
L = 100, xi = 2, p = 0.11, efferr = 0.0770453
Decoder parameter, "
1
3
5
7
9
0 5 10 15 20
0.00
0.02
0.04
0.06
0.08
0.10
p
th
ξ
2
12
30
40
0 10 20 30 40
0
2
4
6
8
Correlation length,
Optimal decoder parameter,
*
𝝈=𝝀/2
𝜇=𝝀
Defect separation, 𝑑
𝑊(𝑑)
0 2 4 6 8 10
0
20 000
40 000
60 000
80 000
100 000
anyon%separa*on,%d
!="
W(d)
#="
0 5 10 15 20 25
0.5
0.6
0.7
0.8
0.9
1.0
Decoding success rate
L = 100, xi = 2, p = 0.11, efferr = 0.0770453
Decoder parameter, "
1
3
5
7
9
0 5 10 15 20
0.00
0.02
0.04
0.06
0.08
0.10
p
th
ξ
2
12
30
40
0 10 20 30 40
0
2
4
6
8
Correlation length,
Optimal decoder parameter,
*
𝜉
3
11
31
41
●●●
●●
●●
Discrete estimation of λ
*
Continuous estimation of λ
*
0 10 20 30 40 50
0
2
4
6
8
10
12
Correlation length, ξ
Optimal decoding parameter, λ
Figure 7: (a) Logical error rates as a function of the
decoder parameter λ are shown for values of ξ =3,11,31
and 41 under a diffusive noise model. The inset shows a
schematic of the decoder weighting function. For each
value of ξ we see a peak in the logical error rate at
approximately λ
ξ. (b) Estimation of the optimal
decoder parameter λ
. We identify the discrete value of
λ for which the logical error rate is maximal (orange data
points), and make a continuous estimation by fitting a
quadratic function around the peak (red data points).
We observe that λ
ξ as we expect from a diffusive
noise model where each error event will typically intro-
duce a pair of stabilizer defects that are separated by
distance
ξ.
of λ
are shown, a discrete estimation by choos-
ing the value of λ for which the logical success
rate is the greatest, and an estimate based on a
quadratic fit around the peak values. We observe
a trend λ
ξ as we might expect from the
diffusive noise model.
We finally remark that we chose a fixed rela-
tionship between µ and σ in our weighting func-
tion, but more generally the functional form can
be parametrized by more variables to explore a
wider space of functions. This may be an inter-
esting route for further work. Using our choice of
adaptive decoder the results we have presented
show that the logical failure rate of the decoder
is sensitive to our choice of λ, and that we can
use this information to improve the failure rates
of our decoding scheme.
4 Improved decoding schemes
In the previous subsection we demonstrated that
we can establish features of noise models by use of
an adaptive decoder, and we showed this explic-
itly for the case of the diffusive noise model. Be-
ing able to gain a knowledge of the noise acting on
underlying hardware is a valuable tool for improv-
ing the performance of quantum error-correcting
codes. Particularly in the regime where the sys-
tem is too large to allow for tomography. Infor-
mation about the error processes involved may
make it easier to identify the source of the noise,
and it may be possible to improve the decoding
algorithm to account for these known errors in
classical post-processing. In this Section we show
that we can use signature plot data to improve
decoding algorithms to increase both threshold
error rates and logical success rates.
We will focus on the data obtained in Sub-
sec. 3.3 to design better decoding algorithms.
Our aim is to use signature plots to design a de-
coder that performs better than the best decoder
we had already identified by sweeping the param-
eter λ. We propose a specialised decoder, that
identifies the two highest peaks on the signature
plot, and uses a weighting function that assigns
very low weight two both of these values of λ,
and high values elsewhere, otherwise the func-
tion assigns a large weight that scales with the
separation of the defects.
We compare the specialised decoder with both
the standard decoder, and the fixed-weight de-
Accepted in Quantum 2019-03-26, click title to verify 10
p
th
(eff err)
, standard decoder
p
th
(eff err)
, fixed-weight decoder
p
th
(eff err)
, specialised decoder
p
th
(event)
, standard decoder
1 3 5 7 9
0.02
0.04
0.06
0.08
0.10
0.12
0.14
correlation length, ξ
p
th
Figure 8: The threshold under diffusive noise for varying
values of ξ for the standard decoder, the fixed weight de-
coder with λ = λ
and a specialised decoder where we
assign a very low weight to more than one value of the
λ. The threshold is shown as a function of the correla-
tion parameter ξ, for the three approaches to decoding.
We see that for ξ > 3, with an appropriate choice of a
non-monotonic weighting function we obtain improved
threshold error rates, thus indicating that we can design
better decoders using the data available in the signature
plot. The thresholds in the effective single qubit error
rate are shown in the solid lines. The dashed blue line
gives a comparison to the event error rate.
coder which has been calibrated with λ = λ
where we use the discrete value of λ
found in the
previous section, i.e. the orange points shown in
Fig. 6(a). Fig. 8 compares the thresholds of the
different decoders. We find that the specialised
decoder outperforms the other decoders for suf-
ficiently high ξ. The calibrated fixed-weight de-
coder also outperforms the standard decoder for
larger values of ξ, but does not perform as well as
the specialised decoder. Importantly, our results
show that updating the weighting function using
information from the signature plot improves the
logical failure rates of the surface code.
We remark that the specialised decoder we
have suggested here is by no means optimal, we
have only demonstrated that even using a very
simple method for using information from the sig-
nature plot the decoder can be improved. We ex-
pect that more sophisticated techniques will al-
low further improvements in the threshold, but
we leave a thorough exploration of the decoder
performance that can be achieved using signature
plots to future work.
It is common to think about noise only as a rate
of Pauli errors on single qubits, and it is useful to
consider how the correlations in the noise affect
this effective single qubit error threshold. Fig. 8
shows the thresholds in the single qubit error rate,
which is defined as the probability with which any
individual qubit has suffered a Pauli error at the
end of the round of noise application. The effec-
tive single qubit error thresholds for the standard
decoder (solid blue line) can be compared with
the error event thresholds (dashed blue line). The
latter decreases rapidly, since a smaller number of
events introduces more single qubit Pauli errors.
However, we see that even with the standard de-
coder, the effective error threshold decreases only
gradually as correlations increase. We also plot
the effective error threshold using the fixed weight
decoder, and specialised decoder, as described in
the previous sections. Here, we see a more grad-
ual decline in the threshold.
It is often believed that any correlations in the
noise model are highly destructive, and should be
avoided. While this may sometimes be the case,
for example with very long range correlations in
which a single error event spans the code, the
effect on the threshold depends strongly on the
exact structure of the correlations. These results
show that for a diffusive noise model resulting in
spatial correlations much smaller than the code
distance, the threshold is not strongly affected,
and using a targeted decoding strategy the loss
of logical error rate can be reduced.
5 Practical considerations
It is worth noting the practicality of the scheme
we have discussed for analysis of a real exper-
imental setup. Given that some hardware has
been developed to implement an error correcting
code, we can perform the following operations:
prepare the code in a known logical state, per-
form stabilizer measurements, and measure out
the logical qubit with the intention of implement-
ing the identity gate. We can repeat this process
many times to build up a large data set of mea-
surement outcomes. A decoder can analyse the
measurement outcomes to determine if a logical
error occurred. We can then perform the analysis
we have described in classical post-processing on
this data set, using a large suite of decoders tar-
geted at errors over a large space of possible noise
Accepted in Quantum 2019-03-26, click title to verify 11
models. The fact that this analysis is a classical
and offline makes it a very low cost way of study-
ing the performance of a logical qubit. Further-
more, since no direct interaction is needed with
the physical device, there need be no requirement
that the decoder is fast, in the sense that it can
keep up with the operating speed of the physi-
cal qubits. Given that we would expect to apply
our method offline, we remark that this may be
a practical use case for slow decoders with high
thresholds, or other useful properties, even if such
decoders are impractical for online decoding.
The results of such an analysis, giving some
characterization of the noise model, could then
be used in multiple ways to improve a device. For
example, adjustments at the hardware level may
be possible to minimize sources of noise, and as
we have shown, we can also improve the classi-
cal processing accompanying the error correcting
code to account for the known structure of the
noise.
6 Conclusions
To summarise, we have developed diagnostic
tools to analyse an unknown parameter of a cor-
related error model acting on the toric code. Our
method generalises the MWPM algorithm to de-
fine a parametrised family of decoders. We go on
to compare the logical error rates of the family
of decoders to hypothesise the types of correlated
errors a code suffers. We have shown that a code
can tolerate a higher error rate if we calibrate
a decoder under the error events that commonly
occur, and furthermore that an increased thresh-
old can be found for certain spatially correlated
noise models. We also discussed how our scheme
could be implemented in a practical setting.
In this work we have studied only a simple
noise model, but we expect that other correlated
error models where string-like errors are drawn
from some distribution, such as thermal noise,
will display a similar behaviour. Beyond spa-
tial correlations, the methods we have proposed
can be more broadly applied to other experi-
mentally relevant models of correlated noise, as
long as an appropriate family of decoders could
be found to characterize the error. For exam-
ple, errors correlated between X and Z bases,
higher probabilities of error in certain physical
locations, and temporal error correlations, per-
haps due to malfunctions in measurement appa-
ratus [68, 69, 70, 71], are all sources of concern.
Many known decoders exist which could poten-
tially be adapted to handle these types of corre-
lations [55, 56, 72, 73, 74, 75, 76, 63]. As a first
step, it will be interesting to extend the ideas we
have presented here to identify multiple unknown
parameters of an error model to enable us to use
decoding algorithms to identify features of more
realistic noise models.
Beyond the toric code, similar techniques to
those we have demonstrated could potentially be
used to probe string-like correlated errors in other
error-correction procedures that use adaptations
of the MWPM algorithm [77, 78, 79]. This read-
ily extends to decoders for the toric code in the
setting where measurements are not ideal [80].
Moreover, it will be interesting to study local cor-
related error events that introduce many stabi-
lizer defects. We expect the Metropolis-based de-
coders presented in Refs. [56, 72], neural network
decoders [81], tensor-network approximants of
maximum likelihood decoders [66, 76] or perhaps
minimum-weight matching decoders that are sup-
plemented by belief propagation [57, 58, 59] may
be suitable for this purpose, but we leave this for
future work.
As hardware rapidly progresses towards the
first implementations of small error correcting
codes, it is important to consider practical meth-
ods for analysing and optimising the performance
of these systems. The methods we have proposed
here offer a simple way of studying and correct-
ing for correlated errors with offline classical post
processing. With further development, we hope
these methods will prove a useful tool for evalu-
ating the performance of the first generation of
fault-tolerant quantum processors.
Acknowledgements
The authors thank S. Benjamin, H. Bombín,
D. Browne, J. Combes, M. Kastoryano, D.
Poulin and J. Wootton for helpful and encour-
aging discussions. We are also grateful to C.
Chubb comments on an earlier draft of the
manuscript. We acknoweldge the use of the
Imperial College Research Computing Service,
DOI: 10.14469/hpc/2232. NHN is supported by
the Engineering and Physical Sciences Research
Council. BJB is supported by the Villum Foun-
Accepted in Quantum 2019-03-26, click title to verify 12
dation, the University of Sydney Fellowship Pro-
gramme and the Australian Research Council via
the Centre of Excellence in Engineered Quantum
Systems(EQUS) project number CE170100009.
A Logical Error Rates for the ballistic
error model in the path-counting regime
Here we show that a targeted decoder will have
better logical error rates in the path-counting
regime for the ballistic noise model for ξ 3. For
simplicity we assume only lattice sizes L = kξ for
even integers k.
In the path-counting regime the errors that
will most frequently cause the decoder to fail are
where k/2 = L/2ξ errors align along the same
row or column of the lattice in a suitable configu-
ration. The standard decoder will fail with prob-
ability independent of the size of the system if all
of these k/2 error events occur in a configuration
where no errors overlap. However, as discussed in
the main text, we can design a targeted decoder
that makes use of value ξ to improve logical fail-
ure rates.
There are N
st.
configurations of k/2 non-
overlapping error events that will cause the stan-
dard decoder to fail where
N
st.
=
2ξ
ξ + 1
(k(ξ + 1)/2)!
(k/2)!(kξ/2)!
. (11)
In the case of the targeted decoder however there
are only N
sp.
error configurations that will cause
logical failure with high probability. This is be-
cause the targeted decoder is designed such that
in the path-counting regime the decoder will fail
if and only if all stabilizer defects of a given error
configuration are separated from each other by
distances that are all factors of ξ. There are only
N
sp.
=
ξk!
(k/2)!(k/2)!
, (12)
of such configurations. We can use the expres-
sions given in Eqn. (11) and Eqn. (12) to bound
the ratio
R =
N
sp.
N
st.
ξ + 1
2
2
ξ
L/2ξ
, (13)
which is proportional to the ratio of the decoder
failure probabilities for the targeted decoder and
the standard decoder for the ballistic noise model.
Clearly, for small ξ 3 the logical failure proba-
bility for the targeted decoder is exponentially
smaller than the logical failure probability of
the standard decoder when dealing with ballistic
noise.
The remainder of this Appendix will prove
Eqn (11). We consider two families of functions,
the first, N
l
(t), denotes the number of configu-
rations that t error events of fixed length ξ can
be aligned on a ring of l sites such that none of
the t error events overlap. The second, M
l
(t), is
the number of configurations that t error events
of size ξ can be arranged along a line of l sites
with open boundary conditions such that none of
the error events overlap.
We have that
N
st.
= N
l=L
(k/2), N
l
(0) = 1. (14)
We show Eqn. (11) recursively by finding a re-
lationship between N
l
(t) and N
l
(t 1). To do so,
we derive some facts relating the function N
l
(t)
to M
l
(t). Specifically, we have that
N
l
(t) =
l
t
M
lξ
(t 1). (15)
We find N
l
(t) of Eqn. (15) by first considering
a single error event of size ξ that can be placed
on any of the l possible positions along a one-
dimensional lattice with periodic boundary con-
ditions. The remaining t 1 non-overlapping er-
ror events can be placed on the vacant l ξ lat-
tice sites that are not occupied by the first er-
ror event. We thus count lM
lξ
(t 1) differ-
ent non-overlapping configurations. With this
expression, each error configuration counted by
N
l
(t) is counted t times. The right-hand side of
the expression thus includes an additional factor
of 1/t to account for the degeneracy. We also
require the relationship
M
l
(t) = N
l
(t) (ξ 1)M
lξ
(t 1). (16)
This expression is found by realising that the
number of configurations included in N
l
(t) that
are not included in M
l
(t) are those where a sin-
gle error event crosses one particular fixed point
of the ring of l sites. There are ξ 1 po-
sitions that one error event may lie to cross
this particular fixed point, and the remaining
t 1 non-overlapping error events may lie in
any of M
lξ
(t 1) configurations. We substitute
Accepted in Quantum 2019-03-26, click title to verify 13
Eqn. (16) into Eqn. (15) to find
M
l
(t) =
1
t(ξ 1)
l
N
l
(t). (17)
We finally combine Eqn. (15) and Eqn. (17) to
obtain
N
l
(t) =
l
t
1
(t 1)(ξ 1)
l ξ
N
lξ
(t1). (18)
We recursively apply Eqn. (18) to Eqn. (14) to
obtain the result of Eqn. (11).
References
[1] J. Chiaverini, D. Leibfried, T. Schaetz, M. D.
Barrett, R. B. Blakestad, J. Britton, W. M.
Itano, J. D. Jost, E. Knill, C. Langer, R. Oz-
eri, and D. J. Wineland. Realization of quan-
tum error correction. Nature, 432:602 EP
–, 12 2004. URL http://dx.doi.org/10.
1038/nature03074.
[2] 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 correction with supercon-
ducting circuits. Nature, 482:382 EP –, 02
2012. URL http://dx.doi.org/10.1038/
nature10786.
[3] R. Barends, J. Kelly, A. Megrant, A. Veitia,
D. Sank, E. Jeffrey, T. C. White, J. Mu-
tus, 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 John M. Martinis. Superconducting
quantum circuits at the surface code thresh-
old for fault tolerance. Nature, 508:500 EP
–, 04 2014. URL http://dx.doi.org/10.
1038/nature13171.
[4] D. Nigg, M. Müller, E. A. Martinez,
P. Schindler, M. Hennrich, T. Monz,
M. A. Martin-Delgado, and R. Blatt.
Quantum computations on a topolog-
ically encoded qubit. Science, 345
(6194):302–305, 2014. DOI: 10.1126/sci-
ence.1253742. URL http://science.
sciencemag.org/content/345/6194/302.
[5] A. D. Córcoles, Easwar Magesan, Srikanth J.
Srinivasan, Andrew W. Cross, M. Steffen,
Jay M. Gambetta, and Jerry M. Chow.
Demonstration of a quantum error detec-
tion code using a square lattice of four su-
perconducting qubits. Nature Communica-
tions, 6:6979 EP –, 04 2015. URL http:
//dx.doi.org/10.1038/ncomms7979.
[6] J. Kelly, R. Barends, A. G. Fowler,
A. Megrant, E. Jeffrey, T. C. White,
D. Sank, J. Y. Mutus, B. Campbell,
Yu Chen, Z. Chen, B. Chiaro, A. Dunsworth,
I. C. Hoi, C. Neill, P. J. J. O’Malley, C. Quin-
tana, P. Roushan, A. Vainsencher, J. Wen-
ner, A. N. Cleland, and John M. Martinis.
State preservation by repetitive error detec-
tion in a superconducting quantum circuit.
Nature, 519:66 EP –, 03 2015. URL http:
//dx.doi.org/10.1038/nature14270.
[7] Maika Takita, A. D. Córcoles, Easwar
Magesan, Baleegh Abdo, Markus Brink,
Andrew Cross, Jerry M. Chow, and Jay M.
Gambetta. Demonstration of weight-
four parity measurements in the surface
code architecture. Physical Review Let-
ters, 117(21):210505–, 11 2016. DOI:
10.1103/PhysRevLett.117.210505. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.117.210505.
[8] Peter W. Shor. Scheme for reducing
decoherence in quantum computer mem-
ory. Physical Review A, 52(4):R2493–
R2496, 10 1995. DOI: 10.1103/Phys-
RevA.52.R2493. URL https://link.aps.
org/doi/10.1103/PhysRevA.52.R2493.
[9] A. M. Steane. Error correcting codes in
quantum theory. Physical Review Letters, 77
(5):793–797, 07 1996. DOI: 10.1103/Phys-
RevLett.77.793. URL https://link.aps.
org/doi/10.1103/PhysRevLett.77.793.
[10] A. Yu. Kitaev. Fault-tolerant quan-
tum computation by anyons. Annals of
Physics, 303:2, 2003. DOI: 10.1016/S0003-
4916(02)00018-0. URL https://doi.org/
10.1016/S0003-4916(02)00018-0.
[11] Eric Dennis, Alexei Kitaev, Andrew Lan-
dahl, and John Preskill. Topological quan-
tum memory. Journal of Mathematical
Physics, 43(9):4452–4505, 2018/10/10 2002.
DOI: 10.1063/1.1499754. URL https://
doi.org/10.1063/1.1499754.
[12] Barbara M. Terhal. Quantum error cor-
rection for quantum memories. Reviews of
Modern Physics, 87(2):307–346, 04 2015.
Accepted in Quantum 2019-03-26, click title to verify 14
DOI: 10.1103/RevModPhys.87.307. URL
https://link.aps.org/doi/10.1103/
RevModPhys.87.307.
[13] Benjamin J. Brown, Daniel Loss, Jian-
nis K. Pachos, Chris N. Self, and James R.
Wootton. Quantum memories at fi-
nite temperature. Reviews of Modern
Physics, 88(4):045005–, 11 2016. DOI:
10.1103/RevModPhys.88.045005. URL
https://link.aps.org/doi/10.1103/
RevModPhys.88.045005.
[14] Earl T. Campbell, Barbara M. Terhal, and
Christophe Vuillot. Roads towards fault-
tolerant universal quantum computation.
Nature, 549:172 EP –, 09 2017. URL http:
//dx.doi.org/10.1038/nature23460.
[15] David P. DiVincenzo. Fault-tolerant ar-
chitectures for superconducting qubits.
Phys. Scr., T137:014020, 2009. URL
https://doi.org/10.1088/0031-8949/
2009/T137/014020.
[16] Martin Wosnitzka, Fabio L. Pedrocchi, and
David P. DiVincenzo. Methodology for bus
layout for topological quantum error correct-
ing codes. EPJ Quantum Technology, 3(1):4,
2016. DOI: 10.1140/epjqt/s40507-016-0042-
8. URL https://doi.org/10.1140/epjqt/
s40507-016-0042-8.
[17] Sergey Bravyi and Alexander Vargo. Sim-
ulation of rare events in quantum er-
ror correction. Physical Review A, 88(6):
062308–, 12 2013. DOI: 10.1103/Phys-
RevA.88.062308. URL https://link.aps.
org/doi/10.1103/PhysRevA.88.062308.
[18] Fern H. E. Watson and Sean D. Bar-
rett. Logical error rate scaling of the
toric code. New J. Phys., 16:093045,
2014. URL https://doi.org/10.1088/
1367-2630/16/9/093045.
[19] Benjamin J. Brown, Naomi H. Nickerson,
and Dan E. Browne. Fault-tolerant er-
ror correction with the gauge color code.
Nature Communications, 7:12302 EP –, 07
2016. URL http://dx.doi.org/10.1038/
ncomms12302.
[20] Andrew J. Landahl, Jonas T. Ander-
son, and Patrick R. Rice. Fault-tolerant
quantum computing with color codes.
arXiv:1108.5738, 2011. URL https://
arxiv.org/abs/1108.5738.
[21] H. G. Katzgraber and R. S. Andrist. Sta-
bility of topologically-protected quantum
computing proposals as seen through spin
glasses. J. Phys.: Conf. Ser., 473:012019,
2013. URL https://doi.org/10.1088/
1742-6596/473/1/012019.
[22] Andrew S. Darmawan and David Poulin.
Tensor-network simulations of the surface
code under realistic noise. Physical Review
Letters, 119(4):040502–, 07 2017. DOI:
10.1103/PhysRevLett.119.040502. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.119.040502.
[23] Sergey Bravyi, Matthias Englbrecht, Robert
König, and Nolan Peard. Correcting co-
herent errors with surface codes. npj
Quantum Information, 4:55, 2018. DOI:
10.1038/s41534-018-0106-y. URL https://
doi.org/10.1038/s41534-018-0106-y.
[24] Pavithran S. Iyer and David Poulin. A small
quantum computer is needed to optimize
fault-tolerant protocols. arXiv:1711.04736,
2017. URL https://arxiv.org/abs/1711.
04736.
[25] Dorit Aharonov, Alexei Kitaev, and John
Preskill. Fault-tolerant quantum computa-
tion with long-range correlated noise. Phys-
ical Review Letters, 96(5):050504–, 02 2006.
DOI: 10.1103/PhysRevLett.96.050504. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.96.050504.
[26] Hui Khoon Ng and John Preskill. Fault-
tolerant quantum computation versus gaus-
sian noise. Physical Review A, 79(3):
032318–, 03 2009. DOI: 10.1103/Phys-
RevA.79.032318. URL https://link.aps.
org/doi/10.1103/PhysRevA.79.032318.
[27] John Preskill. Sufficient condition on noise
correlations for scalable quantum comput-
ing. Quant. Inf. Comp., 13:181, 2013. URL
https://doi.org/10.26421/QIC13.3-4.
[28] Pejman Jouzdani, E. Novais, I. S. Tupit-
syn, and Eduardo R. Mucciolo. Fidelity
threshold of the surface code beyond single-
qubit error models. Physical Review A, 90
(4):042315–, 10 2014. DOI: 10.1103/Phys-
RevA.90.042315. URL https://link.aps.
org/doi/10.1103/PhysRevA.90.042315.
[29] Adrian Hutter and Daniel Loss. Breakdown
of surface-code error correction due to cou-
pling to a bosonic bath. Physical Review A,
89(4):042334–, 04 2014. DOI: 10.1103/Phys-
Accepted in Quantum 2019-03-26, click title to verify 15
RevA.89.042334. URL https://link.aps.
org/doi/10.1103/PhysRevA.89.042334.
[30] Austin G. Fowler and John M. Martinis.
Quantifying the effects of local many-qubit
errors and nonlocal two-qubit errors on the
surface code. Physical Review A, 89(3):
032316–, 03 2014. DOI: 10.1103/Phys-
RevA.89.032316. URL https://link.aps.
org/doi/10.1103/PhysRevA.89.032316.
[31] E. Novais, A. J. Stanforth, and Eduardo R.
Mucciolo. Surface code fidelity at finite
temperatures. Physical Review A, 95(4):
042339–, 04 2017. DOI: 10.1103/Phys-
RevA.95.042339. URL https://link.aps.
org/doi/10.1103/PhysRevA.95.042339.
[32] Joe O’Gorman, Naomi H Nickerson, Philipp
Ross, John JL Morton, and Simon C Ben-
jamin. A silicon-based surface code quan-
tum computer. Npj Quantum Information,
2:15019 EP –, 02 2016. URL https://doi.
org/10.1038/npjqi.2015.19.
[33] R Alicki, M Fannes, and M Horodecki. On
thermalization in kitaev’s 2d model. Jour-
nal of Physics A: Mathematical and Theoret-
ical, 42(6):065303, 2009. DOI: 10.1088/1751-
8113/42/6/065303. URL https://doi.
org/10.1088/1751-8113/42/6/065303.
[34] Stefano Chesi, Beat Röthlisberger, and
Daniel Loss. Self-correcting quantum
memory in a thermal environment. Phys-
ical Review A, 82(2):022305–, 08 2010.
DOI: 10.1103/PhysRevA.82.022305. URL
https://link.aps.org/doi/10.1103/
PhysRevA.82.022305.
[35] E. Novais and Eduardo R. Mucciolo.
Surface code threshold in the presence
of correlated errors. Physical Review
Letters, 110(1):010502–, 01 2013. DOI:
10.1103/PhysRevLett.110.010502. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.110.010502.
[36] P. Jouzdani, E. Novais, and E. R. Mucciolo.
Fidelity of the surface code in the presence
of a bosonic bath. Physical Review A, 88
(1):012336–, 07 2013. DOI: 10.1103/Phys-
RevA.88.012336. URL https://link.aps.
org/doi/10.1103/PhysRevA.88.012336.
[37] C. Daniel Freeman, C. M. Herdman, D. J.
Gorman, and K. B. Whaley. Relaxation
dynamics of the toric code in contact with
a thermal reservoir: Finite-size scaling
in a low-temperature regime. Physi-
cal Review B, 90(13):134302–, 10 2014.
DOI: 10.1103/PhysRevB.90.134302. URL
https://link.aps.org/doi/10.1103/
PhysRevB.90.134302.
[38] D. A. López-Delgado, E. Novais, E. R. Muc-
ciolo, and A. O. Caldeira. Long-time efficacy
of the surface code in the presence of a super-
ohmic environment. Physical Review A, 95
(6):062328–, 06 2017. DOI: 10.1103/Phys-
RevA.95.062328. URL https://link.aps.
org/doi/10.1103/PhysRevA.95.062328.
[39] Dara P. S. McCutcheon, Netanel H.
Lindner, and Terry Rudolph. Error dis-
tributions on large entangled states with
non-markovian dynamics. Physical Review
Letters, 113(26):260503–, 12 2014. DOI:
10.1103/PhysRevLett.113.260503. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.113.260503.
[40] R. Raussendorf, J. Harrington, and
K. Goyal. A fault-tolerant one-
way quantum computer. Annals of
Physics, 321(9):2242–2270, 2006. DOI:
10.1016/j.aop.2006.01.012. URL https:
//doi.org/10.1016/j.aop.2006.01.012.
[41] Austin G. Fowler, Ashley M. Stephens, and
Peter Groszkowski. High-threshold uni-
versal quantum computation on the sur-
face code. Physical Review A, 80(5):
052312–, 11 2009. DOI: 10.1103/Phys-
RevA.80.052312. URL https://link.aps.
org/doi/10.1103/PhysRevA.80.052312.
[42] Austin G. Fowler, Adam C. White-
side, Angus L. McInnes, and Alimo-
hammad Rabbani. Topological code
autotune. Physical Review X, 2(4):
041003–, 10 2012. DOI: 10.1103/Phys-
RevX.2.041003. URL https://link.aps.
org/doi/10.1103/PhysRevX.2.041003.
[43] Naomi H. Nickerson, Ying Li, and Simon C.
Benjamin. Topological quantum computing
with a very noisy network and local error
rates approaching one percent. Nature Com-
munications, 4:1756 EP –, 04 2013. URL
https://doi.org/10.1038/ncomms2773.
[44] Yu Tomita and Krysta M. Svore. Low-
distance surface codes under realistic quan-
tum noise. Physical Review A, 90(6):
062320–, 12 2014. DOI: 10.1103/Phys-
Accepted in Quantum 2019-03-26, click title to verify 16
RevA.90.062320. URL https://link.aps.
org/doi/10.1103/PhysRevA.90.062320.
[45] Sergey Bravyi and Andrew Cross. Doubled
color codes. arXiv:1509.03239, 2015. URL
https://arxiv.org/abs/1509.03239.
[46] Christopher T. Chubb and Steven T.
Flammia. Statistical mechanical models
for quantum codes with correlated noise.
arXiv:1809.10704, 2018. URL https://
arxiv.org/abs/1809.10704.
[47] Jack Edmonds. Paths, trees and flowers.
Canad. J. Math., 17:449, 1965. URL https:
//doi.org/10.4153/CJM-1965-045-4.
[48] Vladimir Kolmogorov. Blossom v: a new
implementation of a minimum cost perfect
matching algorithm. Mathematical Program-
ming Computation, 1(1):43–67, 2009. DOI:
10.1007/s12532-009-0002-8. URL https://
doi.org/10.1007/s12532-009-0002-8.
[49] Joshua Combes, Christopher Ferrie, Chris
Cesare, Markus Tiersch, Gerard J. Milburn,
Hans J. Briegel, and Carlton M. Caves.
In-situ characterization of quantum devices
with error correction. arXiv:1405.5656,
2014. URL https://arxiv.org/abs/1405.
5656.
[50] David P. DiVincenzo and Peter W. Shor.
Fault-tolerant error correction with effi-
cient quantum codes. Physical Review
Letters, 77(15):3260–3263, 10 1996. DOI:
10.1103/PhysRevLett.77.3260. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.77.3260.
[51] David P. DiVincenzo, Peter W. Shor, and
John A. Smolin. Quantum-channel capacity
of very noisy channels. Physical Review A,
57(2):830–839, 02 1998. DOI: 10.1103/Phys-
RevA.57.830. URL https://link.aps.
org/doi/10.1103/PhysRevA.57.830.
[52] Ming-Xia Huo and Ying Li. Learning
time-dependent noise to reduce logical er-
rors: real time error rate estimation in
quantum error correction. New Jour-
nal of Physics, 19:123032, 2017. DOI:
10.1088/1367-2630/aa916e. URL https://
doi.org/10.1088/1367-2630/aa916e.
[53] J. Kelly, R. Barends, A. G. Fowler,
A. Megrant, E. Jeffrey, T. C. White,
D. Sank, J. Y. Mutus, B. Campbell,
Yu Chen, Z. Chen, B. Chiaro, A. Dunsworth,
E. Lucero, M. Neeley, C. Neill, P. J. J.
O’Malley, C. Quintana, P. Roushan,
A. Vainsencher, J. Wenner, and John M.
Martinis. Scalable in situ qubit calibration
during repetitive error detection. Phys-
ical Review A, 94(3):032321–, 09 2016.
DOI: 10.1103/PhysRevA.94.032321. URL
https://link.aps.org/doi/10.1103/
PhysRevA.94.032321.
[54] S. T. Spitz, B. Tarasinkski, C. W. J.
Beenakker, and T. E. O’Brien. Adaptive
weight estimator for quantum error correc-
tion in a time-dependent environment. Adv.
Quant. Tech., 1:1800012, 2018. URL https:
//doi.org/10.1002/qute.201800012.
[55] Guillaume Duclos-Cianci and David
Poulin. Fast decoders for topological
quantum codes. Physical Review Let-
ters, 104(5):050504–, 02 2010. DOI:
10.1103/PhysRevLett.104.050504. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.104.050504.
[56] James R. Wootton and Daniel Loss.
High threshold error correction for the
surface code. Physical Review Let-
ters, 109(16):160503–, 10 2012. DOI:
10.1103/PhysRevLett.109.160503. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.109.160503.
[57] Austin G. Fowler. Optimal complexity cor-
rection of correlated errors in the surface
code. arXiv:1310.0863, 2013.
[58] N. Delfosse and J. Tillich. A decod-
ing algorithm for css codes using the x/z
correlations. In 2014 IEEE International
Symposium on Information Theory, pages
1071–1075, 2014. ISBN 2157-8117. DOI:
10.1109/ISIT.2014.6874997.
[59] Ben Criger and Imran Ashraf. Multi-
path summation for deocding2D topologi-
cal codes. Quantum, 2:102, 2017. DOI:
10.22331/q-2018-10-19-102. URL https://
doi.org/10.22331/q-2018-10-19-102.
[60] Savvas Varsamopoulos, Ben Criger, and
Koen Bertels. Decoding small sur-
face codes with feedforward neural net-
works. Quantum Science and Technol-
ogy, 3(1):015004, 2017. DOI: 10.1088/2058-
9565/aa955a. URL http://dx.doi.org/
10.1088/2058-9565/aa955a.
[61] Stefan Krastanov and Liang Jiang. Deep
neural network probabilistic decoder for
Accepted in Quantum 2019-03-26, click title to verify 17
stabilizer codes. Scientific Reports, 7(1):
11003, 2017. DOI: 10.1038/s41598-017-
11266-1. URL https://doi.org/10.1038/
s41598-017-11266-1.
[62] P. Baireuther, T. E. O’Brien, B. Tarasinkski,
and C. W. J. Beenakker. Machine-learning-
assisted correction of correlated qubit er-
rors in a topological code. Quantum, 2:
48, 2018. DOI: 10.22331/q-2018-01-29-
48. URL https://doi.org/10.22331/
q-2018-01-29-48.
[63] Mishad Maskara, Aleksander Kubica, and
Tomas Jochym-O’Connor. Advantages of
versatile neural-network decoding for topo-
logical codes. arXiv:1802.08680, 2018.
[64] Daniel Gottesman. Stabilizer Codes and
Quantum Error Correction. PhD thesis, Cal-
ifornia Institute of Technology, 1997.
[65] Thomas M. Stace and Sean D. Barrett.
Error correction and degeneracy in surface
codes suffering loss. Physical Review A, 81
(2):022317–, 02 2010. DOI: 10.1103/Phys-
RevA.81.022317. URL https://link.aps.
org/doi/10.1103/PhysRevA.81.022317.
[66] Sergey Bravyi, Martin Suchara, and Alexan-
der Vargo. Efficient algorithms for max-
imum likelihood decoding in the sur-
face code. Physical Review A, 90(3):
032326–, 09 2014. DOI: 10.1103/Phys-
RevA.90.032326. URL https://link.aps.
org/doi/10.1103/PhysRevA.90.032326.
[67] H. Nishimori. Geometry-induced phase
transition in the ±j Ising model. J.
Phys. Soc. Jpn., 55:3305, 1986. DOI:
10.1143/JPSJ.55.3305. URL https://doi.
org/10.1143/JPSJ.55.3305.
[68] Héctor Bombín. Resilience to time-
correlated noise in quantum computation.
Physical Review X, 6(4):041034–, 11 2016.
DOI: 10.1103/PhysRevX.6.041034. URL
https://link.aps.org/doi/10.1103/
PhysRevX.6.041034.
[69] Héctor Bombín. Single-shot fault-tolerant
quantum error correction. Phys. Rev.
X, 5:031043, 2015. DOI: 10.1103/Phys-
RevX.5.031043. URL https://doi.org/
10.1103/PhysRevX.5.031043.
[70] Shota Nagayama, Austin G Fowler, Dominic
Horsman, Simon J Devitt, and Rodney Van
Meter. Surface code error correction on a
defective lattice. New Journal of Physics,
19(2):023050, 2017. DOI: 10.1088/1367-
2630/aa5918. URL http://dx.doi.org/
10.1088/1367-2630/aa5918.
[71] James M. Auger, Hussain Anwar, Mer-
cedes Gimeno-Segovia, Thomas M. Stace,
and Dan E. Browne. Fault-tolerance thresh-
olds for the surface code with fabrica-
tion errors. Physical Review A, 96(4):
042316–, 10 2017. DOI: 10.1103/Phys-
RevA.96.042316. URL https://link.aps.
org/doi/10.1103/PhysRevA.96.042316.
[72] Adrian Hutter, James R. Wootton, and
Daniel Loss. Efficient markov chain monte
carlo algorithm for the surface code. Phys-
ical Review A, 89(2):022326–, 02 2014.
DOI: 10.1103/PhysRevA.89.022326. URL
https://link.aps.org/doi/10.1103/
PhysRevA.89.022326.
[73] Hussain Anwar, Benjamin J Brown, Earl T
Campbell, and Dan E Browne. Fast decoders
for qudit topological codes. New Journal
of Physics, 16(6):063038, 2014. DOI:
10.1088/1367-2630/16/6/063038. URL
http://dx.doi.org/10.1088/1367-2630/
16/6/063038.
[74] Adrian Hutter, Daniel Loss, and James R
Wootton. Improved hdrg decoders for qu-
dit and non-abelian quantum error cor-
rection. New Journal of Physics, 17
(3):035017, 2015. DOI: 10.1088/1367-
2630/17/3/035017. URL http://dx.doi.
org/10.1088/1367-2630/17/3/035017.
[75] Nicolas Delfosse and Naomi H. Nickerson.
Almost-linear time decoding algorithm for
topological codes. arXiv:1709.06218, 2017.
[76] David K. Tuckett, Stephen D. Bartlett, and
Steven T. Flammia. Ultrahigh error thresh-
old for surface codes with biased noise. Phys-
ical Review Letters, 120(5):050505–, 01 2018.
DOI: 10.1103/PhysRevLett.120.050505.
URL https://link.aps.org/doi/10.
1103/PhysRevLett.120.050505.
[77] David S. Wang, Austin G. Fowler,
Charles D. Hill, and Lloyd C. L. Hol-
lenberg. Graphical algorithms and
threshold error rates for the 2D color
code. Quant. Inf. Comp., 10:0780,
2010. DOI: 10.26421/QIC10.9-10. URL
https://doi.org/10.26421/QIC10.9-10.
[78] Nicolas Delfosse. Decoding color codes
by projection onto surface codes. Phys-
Accepted in Quantum 2019-03-26, click title to verify 18
ical Review A, 89(1):012317–, 01 2014.
DOI: 10.1103/PhysRevA.89.012317. URL
https://link.aps.org/doi/10.1103/
PhysRevA.89.012317.
[79] Ashley M. Stephens. Efficient fault-
tolerant decoding of topological color codes.
arXiv:1402.3037, 2014.
[80] Chenyang Wang, Jim Harrington, and
John Preskill. Confinement-higgs tran-
sition in a disordered gauge theory and
the accuracy threshold for quantum mem-
ory. Annals of Physics, 303(1):31–58,
2003. DOI: https://doi.org/10.1016/S0003-
4916(02)00019-2. URL http://www.
sciencedirect.com/science/article/
pii/S0003491602000192.
[81] Giacomo Torlai and Roger G. Melko. Neural
decoder for topological codes. Physical Re-
view Letters, 119(3):030501–, 07 2017. DOI:
10.1103/PhysRevLett.119.030501. URL
https://link.aps.org/doi/10.1103/
PhysRevLett.119.030501.
Accepted in Quantum 2019-03-26, click title to verify 19