You are on page 1of 19

PHYSICAL REVIEW A 101, 012316 (2020)

Editors’ Suggestion

Fault-tolerant bosonic quantum error correction with the surface–Gottesman-Kitaev-Preskill code


1,* 2,†
Kyungjoo Noh and Christopher Chamberland
1
Department of Physics, Yale University, New Haven, Connecticut 06520, USA
2
IBM T. J. Watson Research Center, Yorktown Heights, New York 10598, USA

(Received 19 August 2019; published 13 January 2020)

Bosonic quantum error correction is a viable option for realizing error-corrected quantum information
processing in continuous-variable bosonic systems. Various single-mode bosonic quantum error-correcting codes
such as cat, binomial, and Gottesman-Kitaev-Preskill (GKP) codes have been implemented experimentally
in circuit QED and trapped-ion systems. Moreover, there have been many theoretical proposals to scale up
such single-mode bosonic codes to realize large-scale fault-tolerant quantum computation. Here, we consider
the concatenation of the single-mode GKP code with the surface code, namely, the surface-GKP code. In
particular, we thoroughly investigate the performance of the surface-GKP code by assuming realistic GKP states
with a finite squeezing and noisy circuit elements due to photon losses. By using a minimum-weight perfect
matching decoding algorithm on a three-dimensional spacetime graph, we show that fault-tolerant quantum
error correction is possible with the surface-GKP code if the squeezing of the GKP states is higher than 11.2 dB
in the case where the GKP states are the only noisy elements. We also show that the squeezing threshold changes
to 18.6 dB when both the GKP states and circuit elements are comparably noisy. At this threshold, each circuit
component fails with probability 0.69%. Finally, if the GKP states are noiseless, fault-tolerant quantum error
correction with the surface-GKP code is possible if each circuit element fails with probability less than 0.81%.
We stress that our decoding scheme uses the additional information from GKP-stabilizer measurements and we
provide a simple method to compute renormalized edge weights of the matching graphs. Furthermore, our noise
model is general because it includes full circuit-level noise.

DOI: 10.1103/PhysRevA.101.012316

I. INTRODUCTION bosonic codes should, for example, be concatenated with


some other error-correcting code families such as the surface
Continuous-variable systems or bosonic modes are ubiq-
code [18–20].
uitous in many quantum computing platforms and there have
Recently, there have been proposals for scaling up the cat
been various proposals for realizing quantum computation in
codes by concatenating them with a repetition code [21] or a
continuous-variable systems [1–4]. Notably, bosonic quantum
surface code [22] which are tailored to biased noise models
error correction [5] has recently risen as a hardware-efficient
[23–25]. These schemes take advantage of the fact that the cat
route to implement quantum error correction (QEC) by tak-
code can suppress bosonic dephasing (stochastic random ro-
ing advantage of the infinite dimensionality of a bosonic
tation) errors exponentially in the size of the cat code, thereby
Hilbert space. Various bosonic quantum error-correcting
yielding a qubit with a biased noise predominated either by
codes include the Schrödinger-cat-state [6], binomial [7], and
bit-flip or phase-flip errors. These studies have shown that
Gottesman-Kitaev-Preskill (GKP) [8] codes. All these codes
the gates on the cat code needed for the concatenation can
encode a logical qubit in a physical bosonic oscillator mode
be implemented in a noise-bias-preserving way. On the other
and have been realized experimentally in circuit QED [9–14]
hand, the full concatenated error correction schemes have not
and trapped ion [15–17] systems in the past few years.
been thoroughly studied in these works.
While bosonic QEC with a single bosonic mode (and a
Meanwhile, there have also been studies on scaling up the
single ancilla qubit) can suppress relevant errors such as pho-
GKP code by concatenating it with a repetition code [26],
ton losses or phase-space shift errors in a hardware-efficient
the [[4, 2, 2]] code [26,27], and the surface code [28–30],
way, it should also be noted that logical error rates cannot
or by using cluster states and measurement-based quantum
be suppressed to an arbitrarily small value with this minimal
computation [29,31,32]. One of the recurring themes in these
architecture. For example, the experimentally realized four-
previous works is that the continuous error information gath-
component cat code and the binomial code cannot correct two
ered during the GKP code error correction protocol can boost
(or more) photon-loss events. Similarly, the GKP code cannot
the performance of the next layer of the concatenated error
correct phase-space shift errors of a size larger than a critical
correction. For example, while the surface code by itself has
value. Therefore, to further suppress the residual errors, these
the code capacity threshold ∼11% [19], the threshold can
be increased to ∼14% if the additional error information
from GKP-stabilizer measurements is incorporated in the
*
noh827@gmail.com surface code error correction protocol [28–30]. However, note

mathematicschris@gmail.com that the code capacity thresholds are obtained by assuming

2469-9926/2020/101(1)/012316(19) 012316-1 ©2020 American Physical Society


KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

that only qubits can fail, i.e., gates, state preparations, and Our paper is organized as follows: In Sec. II, we introduce
measurements are assumed perfect. Hence, the above code the surface-GKP code and describe the noise model that we
capacity threshold for the concatenated GKP code is evaluated assume for the fault-tolerance study. In Sec. III, we summarize
by assuming noiseless GKP and surface code stabilizer mea- the main results and establish fault-tolerance thresholds. A
surements or, equivalently, by assuming that ideal GKP states detailed description of our analysis is given in Appendix B.
(with an infinitely large squeezing) are used for the stabilizer In Sec. IV, we compare our results with the previous ones and
measurements. conclude the paper with an outlook.
If the error syndrome is extracted by using realistic GKP
states with a finite squeezing, the error correction proto- II. THE SURFACE–GOTTESMAN-KITAEV-PRESKILL
cols would become faulty. Nevertheless, in the framework CODE
of measurement-based quantum computation [33], it has
been shown that fault-tolerant quantum error correction with In this section, we introduce the surface-GKP code, i.e.,
finitely squeezed GKP states is possible if the strength of the GKP qubits concatenated with the surface code. The GKP
squeezing is above a certain threshold. Specifically, the recent qubits are constructed by using the standard square-lattice
works [29,32] have demonstrated that the threshold value can GKP code that encodes a single qubit into an oscillator mode
be brought down from ∼20 dB [31] to less than 10 dB by [8], which is reviewed in Sec. II A. For the next layer of the
using postselection. encoding, we use the family of rotated surface codes that
In the framework of gate-based quantum computation, requires d 2 data qubits and d 2 − 1 syndrome qubits where
several fault-tolerance thresholds have been computed for the d ∈ {2n + 1 : n ∈ N} is the distance of the code [38,39]. In
GKP code concatenated with the toric code, namely, the toric- Sec. II B, we construct the surface-GKP code and discuss its
GKP code, by assuming a phenomenological noise model implementation. In Sec. II C, we introduce the noise model
[28,30]. In these previous works, however, shift errors were that we use to simulate the full noisy error correction protocol
manually added instead of being derived from an underlying for the surface-GKP code. Readers who are familiar with the
noise model for realistic GKP states and the noisy circuits GKP code and the surface code may skip Secs. II A and II B
used for stabilizer measurements. and are referred to Sec. II C.
In our work, we thoroughly investigate the full error cor-
rection protocol for the GKP code concatenated with the A. Gottesman-Kitaev-Preskill qubit
surface code, namely, the surface-GKP code. We choose the √ √
surface code for the next level of encoding because it can Let q̂ = (↠+ â)/ 2 and p̂ = i(↠− â)/ 2 be the po-
be implemented in a geometrically local way in a planar sition and momentum operators of a bosonic mode, where
architecture. In particular, we consider a detailed circuit-level â and ↠are annihilation and creation operators satisfying
noise model and assume that every GKP state supplied to the [â, ↠] = 1. We define the GKP qubit as the two-dimensional
error correction chain is finitely squeezed, and also that every subspace of a bosonic Hilbert space that is stabilized by the
circuit element can be noisy due to photon losses and heating. two stabilizers
Unlike previous works such as in Refs. [28,30] (where noise √ √
Ŝq ≡ exp[i2 π q̂], Ŝ p ≡ exp[−i2 π p̂]. (1)
propagation was not considered), we comprehensively take
into account the propagation of such imperfections throughout Measuring these two commuting stabilizers is equivalent
the entire circuit and simulate the full surface code error cor- to measuring the √ position and momentum operators q̂
rection protocol by assuming this general circuit-level noise and p̂ modulo π . Therefore, any phase-space shift error
model. Finally, by using a simple decoding algorithm based exp[i(ξ p q̂ − ξq p̂)] acting on the ideal GKP√qubit can be de-
on a minimum-weight perfect matching (MWPM) [34,35] al- tected and corrected as long as |ξq |, |ξ p | < π /2.
gorithm applied to three-dimensional (3D) spacetime graphs, Explicitly, the computational basis states of the ideal GKP
we establish that fault-tolerant quantum error correction is qubit are given by
possible if the squeezing of the GKP states is higher than  √
11.2 dB when the GKP states are the only noisy components |0gkp  = |q̂ = 2n π ,
or higher than 18.6 dB when both the GKP states and cir- n∈Z
cuit elements are comparably noisy. In the latter case, each  √
circuit element that implements the surface-GKP code fails |1gkp  = |q̂ = (2n + 1) π . (2)
with probability 0.69%. In the case where GKP states are n∈Z
noiseless, we find that fault-tolerant quantum error correction
Also, the complementary basis states |±gkp  ≡ √1 (|0gkp  ±
with the surface-GKP code is possible if each circuit element 2
fails with probability less than 0.81%. In general, it has |1gkp ) are given by
been shown that the use of edge weights in the matching  √
graphs which are computed from the most likely error con- |+gkp  = | p̂ = 2n π ,
figurations can significantly improve the performance of a n∈Z
 √
topological code [36,37]. Our decoding algorithm provides a |−gkp  = | p̂ = (2n + 1) π . (3)
simple way to compute renormalized edge weights of the 3D n∈Z
matching graphs, tailored to our general circuit-level noise √
model, based on information obtained from GKP-stabilizer Clearly, all these basis states have q̂ = p̂ = 0 modulo π and
measurements. thus are stabilized by Ŝq and Ŝ p .

012316-2
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

(a) (b)

FIG. 1. (a) Computational basis states (|0gkp , |1gkp ) and complementary basis states (|+gkp , |−gkp ) of an approximate GKP qubit with an
average photon number n̄ = 5. (b) Circuits for measuring the Ŝq and Ŝ p stabilizers. Mq and M p represent the homodyne measurement of the
position and momentum operators, respectively. Also, the controlled-⊕ symbol represents the SUM gate and similarly the controlled- symbol
represents the inverse-SUM gate [see Eq. (6)]. Note that the size of the correction shifts exp[i p̂a zq(b) ] and exp[−iq̂a z (b)
p ] in the Ŝq and Ŝ p stabilizer
measurements are determined by the homodyne measurement outcomes zq(b) and z (b) p .

The ideal GKP qubit states consist of infinitely many given by


infinitely squeezed states and thus are unrealistic. Realistic  
GKP qubit states can be obtained by applying a Gaussian q̂2
envelope operator exp[−n̂] to the ideal GKP states, i.e., Ŝgkp = exp i ,
2

|ψgkp  ∝ exp[−n̂]|ψgkp  and have a finite average photon  π 
number or finite squeezing. Here, n̂ = ↠â is the excitation Ĥgkp = exp i ↠â ,
2
number operator and  characterizes the width of each j→k
peak in the Wigner function of a realistic GKP state. In CNOTgkp = SUM j→k ≡ exp[−iq̂ j p̂k ], (6)
Fig. 1(a), we plot the Wigner functions of the basis states
of an approximate GKP qubit with n̄ = 5. There are many and one can similarly check that
proposals for realizing approximate GKP states in various
experimental platforms [8,40–51]. Notably, approximate GKP Ŝgkp |0gkp  = |0gkp , Ŝgkp |1gkp  = i|1gkp ,
states have been realized experimentally in circuit QED [13] Ĥgkp |0gkp  = |+gkp , Ĥgkp |1gkp  = |−gkp , (7)
and trapped ion systems [15–17]. In Sec. II C, we discuss the
adverse effects of the finite photon number in more detail. In and
this section, we instead focus on the properties of an ideal

j→k  ( j)  (k) 
 ( j)  
GKP qubit. CNOTgkp μgkp νgkp = μgkp (μ ⊕ ν)(k) ,
gkp (8)
Pauli operators of the GKP qubit are given by the square
root of the stabilizers, i.e., ( j) √
for all μ, ν ∈ Z2 , where |μgkp  ≡ n∈Z |q̂ j = (2n + μ) π 
1 √ is the GKP state in the jth mode and ⊕ is the addition
Ẑgkp = (Ŝq ) 2 = exp[i π q̂],
1 √ modulo 2.
X̂gkp = (Ŝ p ) 2 = exp[−i π p̂]. (4) Recall that measuring the stabilizers of the GKP qubit Ŝq
and Ŝ p is equivalent to measuring √the position and the momen-
Indeed, one can readily check that these Pauli operators act on tum operators q̂ and p̂ modulo π . These measurements can
the computational basis states as desired: be performed respectively by preparing an ancilla GKP state
|+gkp  or |0gkp , and then applying the SUMD→A or SUM†A→D
Ẑgkp |0gkp  = |0gkp , Ẑgkp |1gkp  = −|1gkp ,
gate, and finally measuring the position or the momentum
X̂gkp |0gkp  = |1gkp , X̂gkp |1gkp  = |0gkp . (5) operator of the ancilla mode via a homodyne detection [see
Fig. 1(b)]. Here, D refers to the data mode and A refers to
Clifford operations [52] on the GKP qubits can be imple- the ancilla mode. Note that the only non-Gaussian resources
mented by using only Gaussian operations. More explicitly, required for the GKP-stabilizer measurements are the ancilla
j→k
generators of the Clifford group, Ŝgkp , Ĥgkp , and CNOTgkp are GKP states |0gkp  and |+gkp .

012316-3
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

Now, consider the Gaussian random displacement error


channel N [σ ] defined as

2  
d α |α|2
N [σ ](ρ̂) ≡ exp − 2 D̂(α)ρ̂ D̂† (α), (9)
πσ2 σ
where D̂(α) ≡ exp[α ↠− α ∗ â] is the displacement operator
and α ∈ C is the amplitude of the displacement. In the
Heisenberg picture, the error channel N [σ ] adds shift errors to
the position and momentum quadratures; that is, q̂ → q̂ + ξq
and p̂ → p̂ + ξ p , where ξq and ξ p follow a Gaussian random
distribution with zero mean and standard deviation σ : ξq , ξ p ∼
N (0, σ ). If, for example,
√ the size of the√random position shift
ξq is smaller than π /2 (i.e., |ξq | < π /2), the shift can
be correctly identified by measuring √ the GKP √ stabilizer Ŝq .
However, if ξq lies in the range |ξq − π | < √π /2, the shift
is incorrectly identified as a smaller shift ξq − π . Then, FIG. 2. p[σ ](z) for σ = 0.2, 0.5, and 1. p[σ ](z) is defined in Eq.
√ such
a misidentification results in a residual shift exp[−i π p̂] = (11) and represents the conditional probability of having a Pauli X

(or Z) error, given the measurement outcome ξq = z + n π (or ξ p =
X̂gkp and thus causes a Pauli X error on the GKP qubit.√ √
z + n π) for some integer n.
√ In general, if ξ√ ) lies in the range |ξq − n π | <
q (or ξ p√
π /2 (or |ξ p − n π | < π /2) for an odd integer n, the
GKP error correction protocol results in a Pauli X (or Z) error
on the GKP qubit and this happens with probability perr (σ ), surements (on a vacuum state) is a distillable GKP-magic state
where perr (σ ) is defined as and therefore postselection is not necessary. Since Clifford

(2n+ 3 )√π   operations (necessary for magic state distillation) on GKP
 1 2 ξ2 qubits can be implemented by using only Gaussian operations,
perr (σ ) ≡ √ dξ exp − . (10)
n∈Z
2π σ 2 (2n+ 21 )√π 2σ 2 the ability to prepare GKP states is the only non-Gaussian
resource needed for universal quantum computation using
Now, consider a specific instance where, for example, the Ŝq GKP qubits.
stabilizer
√ measurement (i.e., the position measurement
√ mod-
ulo π) informs us√ that ξq is given by ξq = z + n π for some
integer n and |z| < π /2. Then, since odd n corresponds to a B. The surface code with Gottesman-Kitaev-Preskill qubits

Pauli X error and even n corresponds to the no-error case, we Recall that shift errors of size larger than π/2 cannot
can infer that, given the measured value z, there is a Pauli X be corrected by the single-mode GKP code. Here, to correct
error with probability p[σ ](z) where p[σ ](z) is defined as arbitrarily large shift errors, we consider the concatenation
√ of the GKP code with the surface code [18–20], namely, the
exp{−[z − (2n + 1) π ]2 /(2σ 2 )}
p[σ ](z) ≡ n∈Z √ 2 . (11) surface-GKP code. Specifically, we use the family of rotated
n∈Z exp[−(z − n π ) /(2σ )]
2
surface codes [38,39] that only requires d 2 data qubits and
As shown in Fig. 2, the conditional probability p[σ ](z) d 2 − 1 syndrome qubits to get a distance-d code. Note that
becomes larger as |z| gets closer to the decision boundary the distance-d surface code can correct arbitrary qubit errors
√ √ of weight less than or equal to d−1 .
π /2. Therefore,
√ if the measured shift value modulo π is 2
close to ± π /2, we know that this specific instance of the The layout for the data and ancilla qubits of the surface-
GKP error correction is less reliable. This way, the GKP error GKP code is given in Fig. 3. Each of the d 2 data qubits
correction protocol not only corrects the small shift errors (white circles in Fig. 3) corresponds to a GKP qubit as defined
but also informs us how reliable the correction is. Various in Sec. II A. That is, the distance-d surface-GKP code is
ways of incorporating this additional information in the next stabilized by the following 2d 2 GKP stabilizers:
level of concatenated error correction have been studied in √ √
Refs. [26,27,29,30,32]. In Appendix B, we explain in detail Ŝq(k) ≡ exp[i2 π q̂k ], Ŝ p(k) ≡ exp[−i2 π p̂k ], (12)
how the additional information from GKP-stabilizer measure-
ments can be used to compute renormalized edge weights of for k ∈ {1, . . . , d 2 }. These GKP stabilizers are measured by
the matching graphs used in the surface code error correction d 2 ancilla GKP qubits (gray circles in Fig. 3) using the circuits
protocol. given in Fig. 1(b). Moreover, the data GKP qubits are further
Lastly, although not relevant to the purpose of our work, stabilized by the d 2 − 1 surface code stabilizers. For example,
it has been shown that an H-type GKP-magic state |Hgkp  = in the d = 3 case, the 8 surface code stabilizers are given
cos( π8 )|0gkp  + sin( π8 )|1gkp  can be prepared by performing explicitly by
GKP-stabilizer measurements on a vacuum state and √ then
postselecting the Ŝq = Ŝ p = 1 (or q̂ = p̂ = 0 modulo π )
event [44] (see Ref. [53] for more details on the magic states). ŜZ(1) = Ẑgkp
(1) (4)
Ẑgkp , ŜZ(2) = Ẑgkp
(2) (3) (5) (6)
Ẑgkp Ẑgkp Ẑgkp ,
Notably, a more recent study [54] has quantitatively shown ŜZ(3) = Ẑgkp
(4) (5) (7) (8)
Ẑgkp Ẑgkp Ẑgkp , ŜZ(4) = Ẑgkp
(6) (9)
Ẑgkp , (13)
that any postmeasurement state after the GKP-stabilizer mea-

012316-4
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

FIG. 3. The surface-GKP codes with d = 3 and d = 5. White circles represent the data GKP qubits and gray circles represent the ancilla
GKP qubits that are used to measure GKP stabilizers of each data GKP qubit. Green and orange circles represent the syndrome GKP qubits
that are used to measure the Z-type and X -type surface code stabilizers of the data GKP qubits, respectively. In general, there are d 2 data GKP
qubits and (d 2 − 1)/2 Z-type and X -type syndrome GKP qubits. See also Fig. 5 for the reason behind our choice of inverse-SUM gates in the
X -type stabilizer measurements.

and (k) †
Therefore, it does make a difference to choose (X̂gkp ) instead
(1) † (2) (4) (5) † (7) † (8) (k)
of X̂gkp in the noisy measurement case, since shift errors
ŜX(1) = X̂gkp X̂gkp X̂gkp X̂gkp , ŜX(2) = X̂gkp X̂gkp ,
propagate differently depending on the choice. For example,

(3) †
(5) † (6) (8) (9) †
ŜX(3) = X̂gkp
(2)
X̂gkp , ŜX(4) = X̂gkp X̂gkp X̂gkp X̂gkp , (14) we illustrate in Fig. 5 how the initial position shift error in the
fourth X -type syndrome GKP qubit (X 4 qubit) propagates to
(k) √ (k) √ the second Z-type syndrome GKP qubit (Z2 qubit) through
where Ẑgkp ≡ exp[i π q̂k ] and X̂gkp ≡ exp[−i π p̂k ] (see
Fig. 3). the fifth and the sixth data GKP qubits (D5 and D6 qubits).
As shown in Fig. 4, the Z-type surface code stabilizers
are measured by the Z-type GKP syndrome qubits (green
circles in Fig. 3) by using the SUM gates SUMa→e , . . . , SUMd→e
and the position homodyne measurement Mq . Similarly, the
X -type surface code stabilizers are measured by the X -type
GKP syndrome qubits (orange circles in Fig. 3) by using the
SUM and the inverse-SUM gates SUM†e→a , SUMe→b , SUMe→c ,
SUMe→d , and the momentum homodyne measurement M p .

Note that all the Z-type and X -type surface code stabilizers
can be measured in parallel without conflicting with each
other, if the SUM and the inverse-SUM gates are executed in
an order that is specified in Figs. 3 and 4.
We remark that, in the usual case where the surface
code is implemented with bare qubits (such as transmons
[55,56]), it makes no difference to replace, for example,
ŜX(1) = (X̂ (1) )† X̂ (2) X̂ (4) (X̂ (5) )† by ŜX(1) = X̂ (1) X̂ (2) X̂ (4) X̂ (5)
since the Pauli operators are Hermitian. Similarly, the action
(k) †
of (X̂gkp ) on the GKP qubit subspace is identical to that of
(k)
X̂gkp and therefore measuring ŜX(1) = (X̂gkp
(1) † (2) (4) (5) †
) X̂gkp X̂gkp (X̂gkp ) is
equivalent to measuring ŜX(1) = X̂gkp
(1) (2) (4) (5)
X̂gkp X̂gkp X̂gkp in the case
of the surface-GKP code if the syndrome measurements are
noiseless.
(k) †
It is important to note, however, that the actions of (X̂gkp )
(k)
and X̂gkp are not the same outside of the GKP qubit subspace. FIG. 4. Circuits for surface code stabilizer measurements.

012316-5
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

Let us recall that realistic GKP states have a finite average


photon number, or finite squeezing. As discussed in Sec. II A,
a finite-size GKP state can be modeled by applying a Gaussian
envelope operator exp[−n̂] to an ideal GKP state, i.e.,

|ψgkp  ∝ exp[−n̂]|ψgkp . Expanding the envelope operator
in terms of displacement operators [61], we can write

2
   d α

gkp ∝ Tr[exp [−n̂]D̂† (α)]D̂(α)|ψgkp 
π


|α|2
∝ d α exp − 2 D̂(α)|ψgkp ,
2
(15)
2σgkp
FIG. 5. Noise propagation from the X 4 qubit to the Z2 qubit
during surface code stabilizer measurements. The red lightening 1
symbol represents the initial position of a shift error on the qubit
where σgkp 2
= (1 − e− )/(1 + e− ) −−→ /2 [see Eq.
X 4. During the propagation of the shift error to the qubit Z2, the sign (A1)]. That is, an approximate GKP state can be understood
of the shift error is flipped by the inverse-SUM gate SUM†X 4→D5 . This as the state that results from applying coherent superpositions
sign flip then results in cancellations of the propagated shift errors on of displacement operations with a Gaussian envelope to an
the qubit Z2 (empty lightening symbol). ideal GKP state. More details about the approximate GKP
codes can be found in Refs. [44,50,62–64].
To simplify our analysis of the surface-GKP code, we con-
Note that an initial random position shift in the X 4 qubit sider noisy GKP states corrupted by an incoherent mixture of
(represented by the red lightning symbol) is propagated to the displacement operations, instead of the coherent superposition
D6 qubit via the SUM gate SUMX 4→D6 and then to the Z2 qubit as in Eq. (15). That is, whenever a fresh GKP state |0gkp  or
via SUMD6→Z2 . Additionally, it is also propagated to the D5 |+gkp  is supplied to the error correction chain, we assume that
qubit via the inverse-SUM gate SUM†X 4→D5 with its sign flipped a noisy GKP state
and then the flipped shift is further propagated to the Z2 qubit
via SUMD5→Z2 . Thus, the propagated shift errors eventually |0gkp  → N [σgkp ](|0gkp 0gkp |), or
cancel out each other at the Z2 qubit (visualized by the empty |+gkp  → N [σgkp ](|+gkp +gkp |) (16)
lightning symbol) due to the sign flip during the inverse-SUM
gate. is supplied where the Gaussian random displacement error
Note that if the SUM gate SUMX 4→D5 were used instead of N [σ ] is defined in Eq. (9). Note that N [σ ] models an inco-
the inverse-SUM gate SUM†X 4→D5 , the propagated shift errors herent mixture of random displacement errors. We remark that
would add together and therefore be amplified by a factor of the noisy GKP states corrupted by an incoherent displacement
2. In this regard, we emphasize that we have carefully chosen error [as in Eq. (16)] are noisier than the noisy GKP states
the specific pattern of the SUM and the inverse-SUM gates in corrupted by a coherent displacement error [as in Eq. (15)],
Fig. 3 to avoid such noise amplifications. because the former can be obtained from the latter by applying
a technique similar to Pauli twirling [65] (see Appendix A).
In this sense, by adopting the incoherent noise model, we
C. Noise model make a conservative assumption about the GKP noise while
In this section we discuss the noise model that we use to simplifying the analysis.
simulate the full error correction protocol with the surface- We define the squeezing sgkp of a noisy GKP state
GKP code. To be more specific, the surface-GKP error cor- N [σgkp ](|ψgkp ψgkp |) as sgkp ≡ −10 log10 (2σgkp
2
) (aligning
rection protocol is implemented by repeatingly measuring the our notation with that in Refs. [29,31,32]), where the units
Ŝq and Ŝ p GKP stabilizers for each data GKP qubit by using of sgkp are dB. We also assume that idling modes are undergo-
the circuits in Fig. 1(b) and then measuring the surface code ing independent Gaussian random displacement errors N [σ p ]
stabilizers shown in Figs. 3 and 4. Note that the required with variance σ p2 = κt p during the GKP state preparation,
resources for these measurements are as follows: where κ is the photon loss and heating rate (see below) and
(i) Preparation of the GKP states |0gkp  and |+gkp ; t p is the time needed to prepare the GKP states.
(ii) SUM and inverse-SUM gates; Second, we assume that photon loss errors occur contin-
(iii) Position and momentum homodyne measurements; uously during the execution of the SUM or the inverse-SUM
(iv) Displacement operations for error correction. gates. To be more specific, we assume that SUM gates are
We assume that all these components can be noisy except implemented by letting the system evolve under the Hamil-
for the displacement operations since in most experimental tonian Ĥ = gq̂1 p̂2 for t = 1/g (the first mode is the control
platforms, the errors associated with the displacement opera- mode and the second mode is the target mode), during which
tions are negligible compared with the other errors. Moreover, independent photon loss errors occur continuously in both the
note that displacement operations are only needed for error control and the target mode. That is, we replace the unitary
correction. Thus, they need not be implemented physically SUM gate SUM1→2 = exp[−iq̂1 p̂2 ] (or the inverse-SUM gate
SUM1→2 = exp[iq̂1 p̂2 ]) by a completely positive and trace-

in practice since they can be kept track of by using a Pauli
frame [57–60]. Below, we describe the noise model for each preserving (CPTP) map [66] exp[L+ t] (or exp[L− t])
component in more detail. with t = 1/g, where g is the coupling strength and the

012316-6
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

Lindbladian generator L± is given by Then, we find the threshold values for σgkp (Case I), σ
(Case II), and σgkp = σ (Case III) under which fault-tolerant
L± (ρ̂) = ∓ig[q̂1 p̂2 , ρ̂] + κ (D[â1 ] + D[â2 ])ρ̂. (17)
quantum error correction is possible with the surface-GKP
Here, D[Â](ρ̂) ≡ Âρ̂ † − 21 {† Â, ρ̂}, and κ is the photon loss code. Specifically, we take the distance-d surface-GKP code
rate. and repeat the (noisy) stabilizer measurements d times. Then,
In a similar spirit as above, we make a more conservative we construct 3D spacetime graphs based on the stabilizer
assumption about the gate error to make the analysis more measurement outcomes and apply a minimum-weight perfect
tractable. That is, we make the noisy gate exp[L± t] noisier matching decoding algorithm Refs. [34] and [35] to perform
by adding heating errors κ (D[â1† ] + D[â2† ]) to the Lindbladian error correction. Specifically, we use a simple method to
L± , i.e., compute the renormalized edge weights of the 3D match-
ing graphs, based on the information obtained during GKP-
L± ≡ L± + κ (D[â1† ] + D[â2† ]), (18) stabilizer measurements. Such graphs are then used to perform
where the heating rate κ is the same as the photon loss rate. MWPM. A detailed description of our method is given in
This is to convert the loss errors into random displacement Appendix B. Below, we report the logical X error rates, which
errors (see Refs. [5,67]). Indeed, the noisy SUM or the inverse- are the same as the logical Z error rates. Logical Y error rates
SUM gate exp[L± t] is equivalent to the ideal SUM or the
are not shown since they are much smaller than the logical X
inverse-SUM gate followed by a correlated Gaussian random and Z error rates.
displacement error q̂k → q̂k + ξq(k) and p̂k → p̂k + ξ p(k) for In Fig. 6(a), we consider the case where GKP states are
k ∈ {1, 2}, where the additive shift errors are drawn from the only noisy components in the scheme, i.e., σ = 0 (Case I).
bivariate Gaussian distributions (ξq(1) , ξq(2) ) ∼ N (0, N ± We show the performance of the surface-GKP code when both
q ) and
± the additional information from GKP-stabilizer measurements
(ξ p , ξ p ) ∼ N (0, N p ) with the noise covariance matrices
(1) (2)
is incorporated and when it is ignored. When the additional
    information is incorporated, the logical X error rate (same
1 ±1/2 2 4/3 ∓1/2
N±q = σ 2
c ±1/2 , N ±
p = σ c ∓1/2 . as the logical Z error rate) decreases as we increase the
4/3 1
(19) code distance d if σgkp is smaller than the threshold value

σgkp = 0.194 (or if the squeezing of the noisy GKP state sgkp

Here, the variance σc2 is given by σc2 = κt = κ/g. The is higher than the threshold value sgkp = 11.2 dB). That is, in
noise covariance matrices N + +
q and N p are used for the SUM
this case, fault-tolerant error correction is possible with the
gate and N − −
q and N p are used for the inverse-SUM gate. If
surface-GKP code if the squeezing of the GKP states is above
there are idling modes during the application of the SUM or 11.2 dB. Note that if the additional information from GKP-
the inverse-SUM gates on some other pairs of modes, then we stabilizer measurements is ignored, the threshold squeezing
assume that the idling modes undergo independent Gaussian value decreases and logical error rates can range from one to
random displacement errors N [σc ] of the same variance σc2 = several orders of magnitude larger for a given σgkp .
κt = κ/g because they should wait for the same amount of In Fig. 6(b), we consider the case where GKP states are
time until the gates are completed. noiseless but the other circuit elements are noisy, i.e., σgkp = 0
Lastly, we model errors in position and momentum homo- (Case II). In this case, if the additional information from
dyne measurements by adding independent Gaussian random the GKP error correction protocol is incorporated, we can
displacement errors N [σm ] of the variance σm2 = κtm before suppress the logical X error rate (same as the logical Z
the ideal homodyne measurements. Here, tm is the time error rate) to any desired small value by choosing a suffi-
needed to implement the homodyne measurements. Also, ciently large code distance d as long as σ is smaller than
during the homodyne measurements, we assume that idling the threshold value σ = 0.09. Note that since σ 2 = κ/g,
modes are undergoing independent Gaussian random dis- the threshold value σ = 0.09 corresponds to (κ/g) = 8.1 ×
placement errors of the same variance σm2 = κtm . 10−3 = 0.81%, where κ is the photon loss rate and g is
the coupling strength of the SUM or the inverse-SUM gates.
That is, fault-tolerant error correction with the surface-GKP
III. MAIN RESULTS
code is possible if the SUM or the inverse-SUM gates can
In this section, we rigorously analyze the performance of be implemented roughly 120 times faster than the photon
the surface-GKP code by simulating the full error correction loss processes. Note that, if the additional information from
protocol assuming the noise model described in Sec. II C. GKP-stabilizer measurements is ignored, the threshold value
We focus on the case σ p = σc = σm ≡ σ where all circuit becomes smaller and logical error rates can range from one to
elements are comparably noisy. However, we assume that several orders of magnitude larger for a given σ .
the noise afflicting GKP states, σgkp , is independent of the Finally, in Fig. 6(c), we consider the case where the GKP
circuit noise. Since we have two independent noise parameters states and the other circuit elements are comparably noisy,
σgkp and σ , the fault-tolerance thresholds would form a curve i.e., σ = σgkp (Case III). In this case, fault-tolerant error
instead of a single number. Therefore, instead of exhaustively correction is possible if σ = σgkp is smaller than the threshold
investigating the entire parameter space, we consider the value σ = σgkp

= 0.083. This threshold value corresponds to

following three representative scenarios: the GKP squeezing sgkp = 18.6 dB and κ/g = 6.9 × 10−3 =
Case I: σgkp = 0 and σ = 0; 0.69%. Similarly, as in the previous cases, if the additional
Case II: σgkp = 0 and σ = 0; information from GKP-stabilizer measurements is ignored,
Case III: σgkp = σ = 0. the threshold value becomes smaller and logical error rates

012316-7
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

(a) negligible Kerr nonlinearities and thus may not be compatible


with the GKP qubits which should be operated in the regime
where Kerr nonlinearities are negligible [13]. On the other
hand, by using three-wave mixing elements [70], it would
be possible to implement the SUM or the inverse-SUM gates
between two high-Q cavity modes in a way that is not signifi-
cantly limited by Kerr nonlinearities.
Let us now compare the performance of the surface-GKP
code with the usual rotated surface code implemented by bare
qubits such as transmon qubits. Assuming a full circuit-level
depolarizing noise (both for single- and two-qubit gates),
(b)
it was numerically demonstrated that fault-tolerant quantum
error correction is possible with the rotated surface code
if the physical error rate is below the threshold p = 1.2%
[36]. Note that such a high threshold value was obtained by
introducing 3D spacetime correlated edges (see Figs. 3 and
4 in Ref. [36]) and fully optimizing the renormalized edge
weights based on the noise parameters.
Our circuit-level noise model (in terms of shift errors) is
quite different from the depolarizing noise model considered
in typical qubit-based fault-tolerant error correction schemes.
(c) Moreover, we also introduce non-Gaussian resources, i.e.,
GKP states in our scheme. Therefore, our results cannot be
directly compared with the results in Ref. [36]. We never-
theless point out that we obtain comparable threshold values
(κ/g) = 0.81% (Case II) and (κ/g) = 0.69% (Case III)
where κ is the photon loss rate and g is the coupling strength
of the two-mode gates. We stress that we do not introduce
3D spacetime correlated edges and provide a simple method
for computing the renormalized edge weights. In particular,
3D spacetime correlated edges are not necessary in our case
with the surface-GKP code. This is because any shift errors
FIG. 6. The logical X error rate of the surface-GKP code for that are correlated due to two-mode gates will not cause
various d when (a) σ = 0 (Case I), (b) σgkp = 0 (Case II), and (c) any Pauli errors to GKP qubits nor trigger syndrome GKP
σ = σgkp (Case III), which is the same as the logical Z error rate. The qubits incorrectly,
solid lines represent logical error rates when information from the √as long as the size of the correlated shifts
is smaller than π /2, which is the case below the fault-
GKP-stabilizer measurements is used to renormalize edge weights tolerance thresholds computed above.
in the matching graphs. The dotted lines correspond to the case We also point out that in general, topological codes without
when information from GKP-stabilizer measurements is ignored. In
leakage reduction units [71] are not robust against leakage
all cases, given that σgkp and σ are below certain fault-tolerance
errors that occur when a bare qubit state is excited and
thresholds, the logical X or Z error rates are suppressed to an
falls out of its desired two-level subspace [71–74]. In the
arbitrarily small value as we increase the code distance d.
case of the surface-GKP code, leakage errors do occur as
well because each bosonic mode may not be in the desired
can range from one to several orders of magnitude larger for a two-level GKP code subspace. However, the surface-GKP
given noise parameter σ = σgkp . code is inherently resilient to such leakage errors (and thus
For all three cases, we clearly observe that fault-tolerant does not require leakage reduction units) since GKP-stabilizer
quantum error correction with the surface-GKP code is possi- measurements will detect and correct such events. Indeed,
ble despite noisy GKP states and noisy circuit elements, given in our simulation of the surface-GKP code, leakage errors
that the noise parameters are below certain fault-tolerance continuously occur due to shift errors, but the established
thresholds. Recent state-of-the-art experiments have demon- fault-tolerance thresholds are nevertheless still favorable since
strated the capability to prepare GKP states of squeezing be- GKP-stabilizer measurements prevent the leakage errors from
tween 5.5 and 9.5 dB [13,15–17], approaching the established propagating further.

squeezing threshold values sgkp  11.2 dB. We lastly remark that the logical X or Z error rates in
In circuit QED systems, beam-splitter interactions between Fig. 6 decrease very rapidly as σgkp and σ approach zero in
two high-Q cavity modes have been implemented exper- the case of the surface-GKP code. This is again because √ the
imentally with κ/g ∼ 10−2 , where g is the relevant cou- GKP code can correct any shift errors of size less than π /2
pling strength and κ is the photon loss rate [68]. While the and therefore the probability that a Pauli error occurs in a
same scheme (based on four-wave mixing processes) may be GKP qubit (at the end of GKP-stabilizer measurements) be-
adapted to realize the SUM or the inverse-SUM gates between comes exponentially small as σgkp and σ approach zero. More
two high-Q cavity modes [69], this scheme will induce non- precisely, at the end of each GKP-stabilizer measurement, a

012316-8
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

these works describe faulty syndrome extraction procedures


(due to finitely squeezed ancilla GKP states) in a way that
does not take into account the propagation of the relevant shift
errors. More specifically, in Figs. 1, 2, and 7 of Ref. [30], shift
errors are manually added in the beginning of each stabilizer
measurement and right before each homodyne measurement.
Therefore, this phenomenological noise model can be un-
derstood as a model for homodyne detection inefficiencies
while assuming ideal ancilla GKP states. In other words,
the fault-tolerance threshold values established in Ref. [30]
(i.e., σ0 = 0.235 and σ0 = 0.243; see Fig. 12 therein) do not
accurately represent the tolerable noise in the ancilla GKP
states since the noise propagation was not thoroughly taken
into account. Thus, these threshold values can only be taken as

FIG. 7. Visualization of√ the function perr (σ ) (blue). The asymp- a rough upper bound on σgkp and cannot be directly compared
totic expression pasy (σ ) = ( 8σ 2 /π ) exp[−π /(8σ 2 )] is represented with the threshold values obtained in our work. Note also
by the yellow dashed line. perr (σ ) and pasy (σ ) agree well with each that the threshold values in Ref. [30] were computed for the
other in the σ  1 limit. toric code which has a different threshold compared with the
rotated surface code [20].
On the other hand, in our work we assume that every GKP
bulk data GKP qubit undergoes a Pauli X or Z error with state supplied to the error correction chain has a finite squeez-
probability ing and we comprehensively take into account the propagation
 2 of such shift errors through the entire error correction circuit.
perr 5σgkp + 59
3
σ2 , (20) By doing so, we accurately estimate the tolerable noise in

where perr (σ ) is defined in Eq. (10). Here, the variance the finitely squeezed ancilla GKP states by computing σgkp .
2
5σgkp + (59/3)σ 2 was carefully determined by thoroughly Related, we stress that, when the noise propagation is taken
keeping track of how circuit-level noise propagates during into account, detailed scheduling and design of the syndrome
stabilizer measurements (see also Appendix B). As can be extraction circuits become very crucial and we carefully de-
seen from Fig. 7, perr √ (σ ) agrees well with the asymptotic signed the circuits in a way that mitigates the adverse effects
expression pasy (σ ) = ( 8σ 2 /π ) exp[−π /(8σ 2 )] in the σ  of the noise propagation (see Fig. 5).
1 limit. Thus, perr (σ ) decreases exponentially as σ goes to Moreover, we also consider photon loss and heating er-
zero. rors occurring continuously during the implementation of
Similarly, the probability that a bulk surface code stabilizer the SUM and inverse-SUM gates. Thus, we establish fault-
measurement yields an incorrect measurement outcome is tolerance thresholds for the strength of the two-mode cou-
given by pling relative to the photon loss rate and demonstrate that
fault-tolerant quantum error correction with the surface-GKP
 2 code is possible in more general scenarios. We also remark
perr 7σgkp + 116
3
σ2 (21)
that Ref. [30] used a minimum-energy decoder based on
and decays exponentially as σgkp and σ approach zero. There- statistical-mechanical methods in the noisy regime whereas
fore, if the circuit-level noise of the physical bosonic modes we provide a simple method for computing renormalized edge
is very small to begin with, GKP codes will locally provide a weights to be used in a MWPM decoder.
significant noise reduction. In this case, the overall resource Second, Refs. [29,31,32] considered measurement-based
overhead associated with the next level of global encoding quantum computing with GKP qubits and did establish fault-
will be modest since a small-distance surface code would tolerance thresholds for the squeezing of the GKP states.
suffice. Therefore in this regime, the surface-GKP code may Assuming that GKP states are the only noisy components
be able to achieve the same target logical error rate in a more (i.e., Case I), Ref. [31] found the squeezing threshold value

hardware-efficient way than the usual surface code. How- sgkp = 20.5 dB, and Refs. [29] and [32] later brought the

ever, since this regime requires high-quality GKP states, the value down to sgkp = 9.8 dB and sgkp = 8.3 dB, respectively.
additional resource overhead associated with the preparation Notably, the squeezing thresholds found in Refs. [29,32] are
of such high-quality GKP states should also be taken into more favorable than the squeezing threshold found in the

account for a comprehensive resource estimate. We leave such present work, i.e., sgkp = 11.2 dB [see Fig. 6(a)]. In this
an analysis to future work. regard, we remark that the favorable threshold values obtained
in Refs. [29,32] rely on the use of postselection. That is,
each GKP measurement succeeds with probability strictly less
IV. DISCUSSION AND OUTLOOK
than unity and thus the overall success probability would
Here, we compare the results obtained in this paper with decrease exponentially as the system size d increases. On the
previous works in Refs. [28–32]. First, Refs. [28,30] consid- other hand, we do not discard any measurement outcomes
ered the toric-GKP code and computed fault-tolerance thresh- and thus our scheme succeeds with unit probability for any
olds for both code capacity and phenomenological noise mod- distance d. Therefore, our scheme with the surface-GKP code
els. In particular, the phenomenological noise models used in deterministically suppresses errors exponentially with the

012316-9
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

TABLE I. Threshold values for the squeezing of GKP states for fault-tolerant quantum error correction. Here, we compare the established
threshold values obtained by assuming that GKP states are the only noisy components in the error correction circuit (i.e., Case I). “MB” stands

for measurement-based and “GB” stands for gate-based. d is the distance of the code. For Refs. [28,30], σgkp and sgkp are not available because
they assumed phenomenological noise models that do not take into account the propagation of shift errors through the entire error correction
circuit. That is, the threshold values established in Ref. [30] using the toric-GKP code (i.e., σ0 = 0.235 and σ0 = 0.243; see Fig. 12 therein)
do not accurately quantify the tolerable noise in the ancilla GKP states. Instead, σ0 can only be taken as a rough upper bound on σgkp
(see main
text for more details).


Case I (σ = 0) Method σgkp sgkp Post-selection Success probability

Ref. [31] Concatenated codes (MB) 0.067 20.5 dB No 1


Ref. [29] 3D cluster state (MB) 0.228 9.8 dB Yes Decreases exponentially with d 3
Ref. [32] 3D cluster state (MB) 0.273 8.3 dB Yes Decreases exponentially with d 3
Refs. [28,30] Toric-GKP code (GB) NA NA No 1
Present work Surface-GKP code (GB) 0.194 11.2 dB No 1

code distance as long as σgkp and σ are below the threshold decoding on a 3D spacetime graph (with a simple method
values. The differences between our work and the previous for computing renormalized edge weights), we numerically
works are summarized in Table I. demonstrated that fault-tolerant quantum error correction is
Let us now consider the number of bosonic modes needed possible with the surface-GKP code if the squeezing of the
to implement the distance-d surface-GKP code: Recall GKP states and the circuit noise are below certain fault-
Fig. 3 and note that we use d 2 data modes (white circles tolerance thresholds. Since our scheme does not require any
in Fig. 3), d 2 ancilla modes (gray circles in Fig. 3), and postselection and thus succeeds with unit probability, our
d 2 − 1 syndrome modes (green and orange circles in Fig. 3). scheme is clearly scalable. We also described our methods in
Although we introduced the d 2 ancilla modes to describe our great detail such that our results can easily be reproduced.
scheme in a simpler way, the d 2 ancilla modes can in fact
be replaced by the d 2 − 1 syndrome qubits plus one more
ACKNOWLEDGMENTS
additional mode. Thus, we only need a total of 2d 2 modes and
geometrically local two-mode couplings to implement the We thank Andrew Cross, Christophe Vuillot, Barbara
distance-d surface-GKP code. For example, 18 modes would Terhal, Alec Eickbusch, Steven Touzard, and Philippe
suffice to realize the smallest nontrivial case with d = 3. Campagne-Ibarcq for helpful discussions and for providing
We finally emphasize that we modeled noisy GKP states by comments on the paper. K.N. is grateful for the hospitality of
applying an incoherent random displacement error N [σgkp ] the IBM T. J. Watson research center where this work was
to the ideal GKP states, similarly as in Refs. [29,31,32]. conceived and completed.
While we use this noise model for theoretical convenience
and justify it by using a twirling argument (see Appendix A),
similar to the justification of a depolarizing error model in APPENDIX A: SUPPLEMENTARY MATERIAL FOR
qubit-based QEC, we remark that it is not practical to use THE NOISE MODEL
the twirling operation in realistic situations. This is because To derive Eq. (15), we used the following identity:
the twirling operation increases the average photon number
of the GKP states, whereas in practice it is desirable to keep Tr[exp [−n̂]D̂† (α)]
the photon number bounded below a certain cutoff. Therefore,


an interesting direction for future work would be to see if one
can implement the stabilizer measurements in Figs. 1, 3, and = e−n n|D̂† (α)|n
n=0
4 in a manner that prevents the average photon number from
  ∞
diverging as we repeat the stabilizer measurements. It will be |α|2  −n
especially crucial to keep the average photon number under = exp − e Ln (|α|2 )
2 n=0
control when each bosonic mode suffers from dephasing
   
errors and/or undesired nonlinear interactions such as Kerr |α|2 1 e−
nonlinearities. = exp − exp − |α| 2
2 1 − e− 1 − e−
Related, we remark that, in the recent experimental realiza-  
tion of the GKP code in a circuit QED system, an envelope- 1 1 + e−
= exp − |α| ,
2
(A1)
trimming technique was used to constrain the average photon 1 − e− 2(1 − e− )
number of the system [13]. Whether a similar technique can be
incorporated in a large-scale surface-GKP code architecture where Ln (x) is the Laguerre polynomial. Furthermore, going
would be an interesting future research direction. from the third to fourth line, we used the generating
function
To summarize, we have thoroughly investigated the perfor- for the Laguerre polynomials which satisfies ∞ n=0 t Ln (x) =
n
1 −tx/(1−t )
mance of the surface-GKP code assuming a detailed circuit- 1−t
e .
level noise model. By simulating the full noisy error correc- Now, we explain how one can transform the noisy GKP
tion protocol and using a minimum-weight perfect matching state corrupted by coherent superpositions of displacement

012316-10
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

errors [see Eq. (15)] into a noisy GKP state corrupted √ by an incoherent mixture of displacement errors [see Eq. (16)]. To do
so, we apply random shifts of integer multiples 2 π in both the position and the momentum directions to the noisy GKP state
 
|ψgkp  ∝ exp[−n̂]|ψgkp . Then, |ψgkp  is transformed into
   

ψ̂gkp ∝ (Ŝq )n1 (Ŝ p )n2 ψgkp
   † n2 † n1
ψgkp (Ŝ p ) (Ŝq )
n1 ,n2 ∈Z


|α| 2
+ |β| 2
∝ d 2 αd 2 β exp − 2
(Ŝq )n1 (Ŝ p )n2 D̂(α)|ψgkp ψgkp |D̂† (β )(Ŝ †p )n2 (Ŝq† )n1
2σgkp
n1 ,n2 ∈Z


|α|2 + |β|2 √ √
∝ d αd β exp −
2 2
2
exp[i 2π (αR − βR )n1 − i 2π (αI − βI )n2 ]
2σgkp
n1 ,n2 ∈Z

× D̂(α)(Ŝq )n1 (Ŝ p )n2 |ψgkp ψgkp |(Ŝ †p )n2 (Ŝq† )n1 D̂† (β )


|α|2 + |β|2 √ √
∝ d αd β exp −
2 2
2
exp[i 2π (αR − βR )n1 − i 2π (αI − βI )n2 ]D̂(α)|ψgkp ψgkp |D̂† (β ), (A2)
2σgkp
n1 ,n2 ∈Z
∗ ∗
αβ −α β
where we√used the identity D̂(α) √ D̂(β ) = D̂(β )D̂(α)e and the fact that GKP states are stabilized by the GKP stabilizers
Ŝq =
D̂(i 2π ) and Ŝ p = D̂( 2π ), i.e., Ŝq |ψgkp  = Ŝ p |ψgkp  = |ψgkp . Using the Poisson summation formula, n∈Z e
ian
=
2π k∈Z δ(a − 2π k) we can further simplify Eq. (A2) as


|α|2 + |β|2 √ √

ψ̂gkp ∝ d αd β exp −
2 2
2
δ(αR − βR − 2π k1 )δ(αI − βI − 2π k2 )D̂(α)|ψgkp ψgkp |D̂† (β )
2σgkp
k1 ,k2 ∈Z
√ 

|α|2 + |α − 2π (k1 + ik2 )|2 √
= d α exp −
2
2
D̂(α)|ψgkp ψgkp |D̂† (α − 2π (k1 + ik2 ))
2σgkp
k1 ,k2 ∈Z

   
 π |k1 + ik2 |2 α − π (k1 + ik2 )2 √
= exp − d α exp −
2 2
D̂(α)|ψgkp ψgkp |D̂† (α − 2π (k1 + ik2 )). (A3)
2
2σgkp σgkp
2
k1 ,k2 ∈Z


Lastly, if σgkp  π (which is the case below the fault- yields
tolerance threshold σgkp
 0.194), we can neglect all the     
t t N
(k1 , k2 ) = (0, 0) terms due to the exponentially decaying pref- exp[L± t]= lim exp V± exp Lerr . (A7)
N→∞ N N
actor exp[− π|k2σ 1 +ik2 |
2
2 ] and get the noise model in Eq. (16):
gkp Note that both exp[V± t/N] and exp[Lerr t/N] are Gaussian
 channels with the characterization matrices

 d 2α |α|2 ⎡ ⎤
ψ̂gkp ∝ exp − 2 D̂(α)|ψgkp ψgkp |D̂† (α) 1 0 0 0
π σgkp
2 σgkp ⎢ 0 1 0 ∓1/N ⎥
T± = ⎣ , N ± = 0,
= N [σgkp ](|ψgkp ψgkp |). (A4) ±1/N 0 1 0 ⎦
0 0 0 1
Let us now derive the gate error model given in Eq. (19). ⎡ ⎤ ⎡ ⎤
Recall that L± is given by 1 0 0 0 1 0 0 0
⎢0 1 0 0⎥ κt ⎢0 1 0 0⎥
=⎣ , = ,
0⎦ ⎣ 0⎦
T err N err
0 0 1 N 0 0 1
L± = V± + Lerr , (A5) 0 0 0 1 0 0 0 1
(A8)
where V± and Lerr are defined as
respectively (see, for example, Ref. [76] for the definition of
V± (ρ̂) ≡ ∓ig[q̂1 p̂2 , ρ̂], Gaussian channels and their characterization matrices). Thus,
the quadrature operator x̂ = (q̂1 , p̂1 , q̂2 , p̂2 )T is transformed

2
via the noisy SUM or the inverse-SUM gate as
Lerr (ρ̂) ≡ κ (D[âk ] + D[âk† ])ρ̂. (A6) ⎡ ⎤ ⎡ ⎤
k=1 1 0 0 0 q̂1
⎢ 0 1 0 ∓1⎥ ⎢ p̂ ∓ p̂2 ⎥
x̂ → (T ± )N x̂ = ⎣ x̂ = ⎣ 1 ,
The noisy SUM or the inverse-SUM gates is then given by ±1 0 1 0⎦ q̂2 ± q̂1 ⎦
exp[L± t] with t = 1/g. Note that Trotter’s formula [75] 0 0 0 1 p̂2
(A9)
012316-11
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

as desired. Also, the covariance matrix V is transformed as



N
V → (T ± )N V ((T ± )N )T + (T ± )k N err (T T± )k
k=1
⎡ ⎤
1 0 ± Nk 0
2
 κt ⎢ ⎥
N
⎢ 0 1 + Nk 0 ∓ Nk ⎥
= (T ± ) V ((T ± ) ) +
N N T
⎢ k k 2 ⎥
k=1
N ⎣± N 0 1+ N 0 ⎦
0 ∓ Nk 0 1
⎡ ⎤
1 0 ±1/2 0
⎢ 0 4/3 0 ∓1/2⎥
= (T ± ) V ((T ± ) ) + κt ⎣
N N T
.
0 ⎦
(A10)
±1/2 0 4/3
0 ∓1/2 0 1
Therefore, the noisy SUM or the inverse-SUM gate can be understood as the ideal SUM or the inverse-SUM gate followed by a
correlated Gaussian random displacement error with the noise covariance matrices N ± ±
q and N p as given in Eq. (19).

APPENDIX B: SIMULATION DETAILS


Here, we describe in detail how we simulate the syndrome extraction protocol for the surface-GKP code and how we decode
the obtained syndrome measurement outcome.

1. GKP-stabilizer measurements
Consider the distance-d surface-GKP √ code consisting of √ d 2 data GKP qubits. Each data GKP qubit is stabilized by the
two GKP stabilizers Ŝq = exp[i2 π q̂k ] and Ŝ p = exp[−i2 π p̂k ] where k ∈ {1, . . . , d 2 }. In the first step of GKP-stabilizer
(k) (k)

measurements (left in Fig. 8), Ŝq(k) (Ŝ p ) stabilizers are measured for odd (even) k. In the second step (right in Fig. 8), on the
other hand, Ŝ p(k) (Ŝq(k) ) stabilizers are measured for odd (even) k. Note that we alternate between Ŝq and Ŝ p measurements in a
checkerboard pattern in order to balance the position and momentum quadrature noise.
Let ξqD and ξ pD (ξqA and ξ pA ) be the data (ancilla) position and momentum quadrature noise, where
2 2
ξqD = ξq(D1) , . . . , ξq(Dd ) , ξ pD = ξ p(D1) , . . . , ξ p(Dd ) ,
2 2
ξqA = ξq(A1) , . . . , ξq(Ad ) , ξ pA = ξ p(A1) , . . . , ξ p(Ad ) . (B1)
In step 1, we add random shift errors occurring during the GKP state preparation as follows:
ξq(Dk) ← ξq(Dk) + randG(σ 2 ),
ξ p(Dk) ← ξ p(Dk) + randG(σ 2 ),
2
ξq(Ak) ← randG σgkp ,

ξ p(Ak) ← randG σgkp 2
, (B2)
for k ∈ {1, . . . , d 2 } where randG(V ) generates a random vector sampled from a multivariate Gaussian distribution N (0, V ) with
zero mean and the covariance matrix V . Then, due to the SUM and the inverse-SUM gates, the quadrature noise vectors are updated
as follows:
  
(Dk) (Ak) (Dk) (Ak) 1 1/2
ξq , ξq ← ξ q , ξq + ξq (Dk)
+ randG σ 2
,
1/2 4/3
  
(Dk) (Ak) 4/3 −1/2
ξp , ξp ← ξ p(Dk) − ξ p(Ak) , ξ p(Ak) + randG σ 2 , (B3)
−1/2 1
for odd k (Ŝq(k) stabilizer measurement) and
  
(Dk) (Ak) 4/3 −1/2
ξ q , ξq ← ξq(Dk) − ξq(Ak) , ξq(Ak) + randG σ 2 ,
−1/2 1
  
(Dk) (Ak) (Dk) (Ak) 1 1/2
ξp , ξp ← ξp , ξp + ξp (Dk)
+ randG σ 2
, (B4)
1/2 4/3

012316-12
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

FIG. 8. Measurement of the GKP stabilizers for d = 3. See also


Fig. 1(b) for the graphical notation.

for even k (Ŝ p(k) stabilizer measurement). Due to the noise


before (or during) the homodyne measurement, the noise
vectors are updated as

ξq(Dk) ← ξq(Dk) + randG(σ 2 ),


FIG. 9. Measurement of the surface code stabilizers for d = 3.
ξ p(Dk) ← ξ p(Dk) + randG(σ ),
2

ξq(Ak) ← ξq(Ak) + randG(σ 2 ), GKP qubits, where


ξ p(Ak) ← ξ p(Ak) + randG(σ 2 ), (B5)  
ξqZ = ξq(Z1) , . . . , ξq(Zd ) , ξ pZ = ξ p(Z1) , . . . , ξ p(Zd ) ,
for all k ∈ {1, . . . , d 2 }. Then, through the homodyne measure-  
ξqX = ξq(X 1) , . . . , ξq(X d ) , ξ pX = ξ p(X 1) , . . . , ξ p(X d ) . (B9)
ment and the error correction process, the data noise vectors
are transformed as
Note that the SUM and the inverse-SUM gates for the syndrome
extraction are executed in four time steps (see steps 3–6 in
ξq(Dk) ← ξq(Dk) − R√π ξq(Ak) , (B6) Fig. 9). Let Z1 (k), . . . , Z4 (k)[X1 (k), . . . , X4 (k)] be the label
of the data GKP qubit that the k th Z-type (X -type) syndrome
GKP qubit is coupled with in steps 3, . . . , 6 (if the syndrome
ξ p(Dk) ← ξ p(Dk) − R√π ξ p(Ak) , (B7)
GKP qubit is idling, the value is set to be zero). For example,
when d = 3, Z1 (k) and X1 (k) are given by
for odd and even k, respectively. Rs (z) is defined as
 
z 1 Z1 (1) = 1, Z1 (2) = 3, Z1 (3) = 5, Z1 (4) = 0,
Rs (z) ≡ z − s + . (B8)
s 2 X1 (1) = 2, X1 (2) = 0, X1 (3) = 8, X1 (4) = 6,
(B10)
In step 2, Ŝ p(k) (Ŝq(k) ) stabilizers are measured for odd (even)
k instead of Ŝq(k) (Ŝ p(k) ). Thus, the noise vectors are updated
representing the connectivity between the syndrome and the
similarly as in Eqs. (B2) to (B7), except that Eqs. (B3) and
data GKP qubits in step 3.
(B6) [Eqs. (B4) and (B7)] are applied when k is even (odd)
Due to the shift errors occurring during the preparation of
instead of when k is odd (even).
GKP states, the noise vectors are updated as follows:

2. Surface code stabilizer measurements ξq(Dk) ← ξq(Dk) + randG(σ 2 ), ξ p(Dk) ← ξ p(Dk) + randG(σ 2 ),

Recall that there are d ≡ (d − 1)/2 Z-type and X -type
2 2 2
ξq(Z) ← randG σgkp , ξ p(Z) ← randG σgkp ,
syndrome GKP qubits that are used to measure the surface 2 2
code stabilizers. Let ξqZ and ξ pZ (ξqX and ξ pX ) be the position and (X ) (X )
ξq ← randG σgkp , ξ p ← randG σgkp , (B11)
momentum noise vectors of the Z-type (X -type) syndrome

012316-13
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

for k ∈ {1, . . . , d 2 } and  ∈ {1, . . . , d  }. In step 3, the SUM gates transform the noise vectors as
  
(DZ1 ()) (Z) (DZ1 ()) (Z) 1 1/2
ξq , ξq ← ξq , ξq + ξq (DZ1 ())
+ randG σ 2
,
1/2 4/3
  
(DZ1 ()) (Z) 4/3 −1/2
ξp , ξp ← ξ p(DZ1 ()) − ξ p(Z) , ξ p(Z) + randG σ 2 , (B12)
−1/2 1
for all  ∈ {1, . . . , d  } if Z1 () = 0 and
ξq(Z) ← ξq(Z) + randG(σ 2 ),
ξ p(Z) ← ξ p(Z) + randG(σ 2 ), (B13)
if Z1 () = 0. Similarly,
  
(DX1 ()) (X ) (DX1 ()) (X ) (X )
2 4/3 1/2
ξq , ξq ← ξq + ξq , ξq + randG σ ,
1/2 1
  
(DX1 ()) (X ) 1 −1/2
ξp , ξp ← ξ p(DX1 ()) , ξ p(X ) − ξ p(DX1 ()) + randG σ 2 , (B14)
−1/2 4/3
for all  ∈ {1, . . . , d  } if X1 () = 0 and
ξq(X ) ← ξq(X ) + randG(σ 2 ),
ξ p(X ) ← ξ p(X ) + randG(σ 2 ), (B15)
if X1 () = 0. Since there are idling data GKP qubits, the data noise vectors are updated as
ξq(Dk) ← ξq(Dk) + randG(σ 2 ),
ξ p(Dk) ← ξ p(Dk) + randG(σ 2 ), (B16)

only for k such that Z1 () = k and X1 () = k for all  ∈ {1, . . . , d }.
In step 4, the SUM gates between the Z-type syndrome GKP qubits and data GKP qubits transform the noise vectors in the
same way as in Eqs. (B12) and (B13) except that Z1 () is replaced by Z2 (). However, since the X -type syndrome GKP qubits
are coupled with the data GKP qubits through inverse-SUM gates instead of SUM gates, the noise vectors are then updated as
  
(DX2 ()) (X ) (DX2 ()) (X ) (X )
2 4/3 −1/2
ξq , ξq ← ξq − ξq , ξq + randG σ ,
−1/2 1
  
(DX2 ()) (X ) 1 1/2
ξp , ξp ← ξ p(DX2 ()) , ξ p(X ) + ξ p(DX2 ()) + randG σ 2 , (B17)
1/2 4/3

for all  ∈ {1, . . . , d  } if X2 () = 0 and ξq(X ) ← ξq(X ) + randG(σ 2 ), ξ p(X ) ← ξ p(X ) + randG(σ 2 ),
ξq(X ) ← ξq(X ) + randG(σ 2 ), (B19)
ξ p(X ) ← ξ p(X ) + randG(σ 2 ), (B18) for all k ∈ {1, . . . , d 2 } and  ∈ {1, . . . , d  }. Then, through the
homodyne measurement, we measure ξq(Z) and ξ p(X ) modulo
if X2 () = 0, instead of as in Eqs. (B14) and (B15). Due to the √
idling data GKP qubits, the noise vectors are further updated 2 π and assign stabilizer values as
  
as in Eq. (B16) only for k such that Z2 () = k and X2 () = k +1 R√ ξ (Z)   √π /2
2π q
for all  ∈ {1, . . . , d  }. ()
ŜZ ←   √
Note that, in steps 5 and 6, the X -type syndrome GKP −1 R√2π ξq(Z)  > π /2,
   √
qubits are coupled with the data GKP qubits via inverse-SUM
()
+1 R√2π ξ p(X )   π /2
gates and SUM gates, respectively. Therefore, in step 5, the ŜX ←   √ (B20)
noise vectors are updated in the same way as in step 4, except −1 R√2π ξ p(X )  > π /2,
that Z2 () and Z2 () are replaced by Z3 () and X3 (). On the
other hand, in step 6, the noise vectors are updated in the same for all  ∈ {1, . . . , d  }. Rs (z) is defined in Eq. (B8).
way as in step 3, except that Z1 () and X1 () are replaced
by Z4 () and X4 (). Due to the noise before (or during) the 3. Construction of three-dimensional spacetime graphs
homodyne measurement, the noise vectors are updated as Now we construct 3D spacetime graphs to which we will
ξq(Dk) ← ξq(Dk) + randG(σ ),2
ξ p(Dk) ← ξ p(Dk) + randG(σ ),2 apply a minimum-weight perfect matching decoding algo-
rithm. The overall structure is as follows: Since each stabilizer
ξq(Z) ← ξq(Z) + randG(σ 2 ), ξ p(Z) ← ξ p(Z) + randG(σ 2 ), measurement can be faulty, we repeat the noisy stabilizer

012316-14
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

measurement cycle d times. Then, we perform another round


of ideal stabilizer measurement cycle assuming that all circuit
elements and supplied GKP states are noiseless. The reason
for adding the extra noiseless measurement cycle is to ensure
that the noisy states are restored back to the code space so
that we can later conveniently determine whether the error
correction succeeds. Then, the Z-type and the X -type 3D
spacetime graphs are constructed to represent the outcomes
of d + 1 rounds of stabilizer measurement cycles. These
spacetime graphs will then be used to decode the Z-type and
the X -type syndrome measurement outcomes.
We first construct the Z-type and X -type two-dimensional
(2D) space graphs as in Fig. 10. Each bulk vertex of the FIG. 10. Z-type and X -type 2D space graphs for the surface-
2D space graph corresponds to a syndrome GKP qubit GKP code with d = 5. These 2D graphs will be stacked up to
and each bulk edge corresponds to a data GKP qubit. construct Z-type and X -type 3D spacetime graphs.
Note also that there are boundary vertices (squares in
Fig. 10) that do not correspond to any syndrome GKP We start by initializing the data position and momentum
qubits and the corresponding boundary edges (blue lines in noise vectors to a zero vector:
2
Fig. 10) that are not associated with any data GKP qubits. ξqD = ξq(D1) , . . . , ξq(Dd ) = (0, . . . , 0),
Therefore, the boundary edge weights are always set to 2

be zero. ξ pD = ξ p(D1) , . . . , ξ p(Dd ) = (0, . . . , 0). (B21)


Then, we associate each 2D space graph with one round These data noise vectors are fed into step 1 of the GKP-
of stabilizer measurement cycle. So, there are d + 1 2D space stabilizer measurement as described in Eqs. (B2) to (B5). Let
graphs and these 2D space graphs are stacked up together by wZH (k) and wZH (k) be the horizontal edge weights of the Z-type
introducing vertical edges that connect the same vertices in and X -type graphs corresponding to the kth data GKP qubit
two adjacent 2D space graphs (corresponding to two adjacent (k ∈ {1, . . . , d 2 }). Then, while updating the data position and
stabilizer measurement rounds). Below, we discuss in detail momentum noise vectors as prescribed in Eqs. (B6) and (B7),
how the bulk edge weights are assigned. we assign the horizontal edge weights as

⎧ !   "

⎪− log p σ 2 + 10 σ 2 R√ ξ (Ak) round 1

⎪ 2 gkp 3 π q
⎨ # H $ (Ak)
wZ (k) ← − log2 p σZ (k; d ) R√π ξq
H
round 2 to round d (B22)

⎪ 

⎩− log p# σ H (k; d ) 2 − σ 2 − 10 σ 2 $ R√ ξ (Ak)

round d + 1,
2 Z gkp 3 π q

for odd k and


⎧ # 2 $

⎪ − log p σgkp + 10 σ 2 R√π ξ p(Ak) round 1


2
# H
3
$ (Ak)
wX (k) ← − log2 p σX (k; d ) R π ξ p
H √ round 2 to round d (B23)

⎪ 

⎩ # H 2 $
− log2 p σX (k; d ) − σgkp2 − 10 σ 2 R√ ξ (Ak)
3 π p round d + 1,

for even k if the additional GKP information is used. Here, we use ξq(Ak) and ξ p(Ak) that are obtained after applying Eq. (B5).
perr (σ ) and p[σ ](z) are defined in Eqs. (10) and (11) and Rs (z) is defined in Eq. (B8). On the other hand, if the additional GKP
information is not used, we assign the horizontal edge weights as
⎧  2

⎪ − log 2 perr σgkp + 10 σ2 round 1

⎨ H
3

wZH (k) ← − log2 perr σZ (k; d ) round 2 to round d (B24)

⎪ #

⎩ $2
− log2 perr σZH (k; d ) − σgkp
2 − 10 σ 2
3
round d + 1,

for odd k and


⎧  2

⎪ − log perr σgkp + 10 σ2 round 1

⎪ 2 3

wXH (k) ← − log2 perr σXH (k; d ) round 2 to round d (B25)

⎪ 

⎩− log perr #σ H (k; d )$2 − σ 2 −

σ
10 2

round d + 1,
2 X gkp 3

012316-15
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

for even k. Here, σZH (k; d ) and σXH (k; d ) are defined as
⎧⎧

⎪⎨ 4σgkp2 + 52 σ 2 , k−1
∈ 2Z

⎪ 3 d

⎪  k ∈ dZ + 1

⎪⎩

⎪ 4σ + 3 σ 2 ,
2 58 k−1
∈ 2Z + 1
⎨⎧ gkp d

σZH (k; d ) ≡ ⎨ 4σgkp2 + 55 σ 2 , k


∈ 2Z + 1

⎪ 
3 d
k ∈ dZ

⎪⎩

⎪ 4σgkp + 3 σ ,
2 49 2 k
∈ 2Z

⎪ d


⎩ 5σ 2 + 59 σ 2 , otherwise,
gkp 3

⎧⎧

⎪⎨ 4σgkp2 + 49 σ 2 , k ∈ 2Z + 1

⎪ 3

⎪  k ∈ {1, . . . , d}

⎪⎩

⎪ 4σ + 3 σ 2 ,
2 55
k ∈ 2Z
⎨⎧ gkp
σXH (k; d ) ≡ ⎨ 4σgkp2 + 58 σ 2 , k ∈ 2Z + 1 (B26)

⎪ 
3
k ∈ {d 2 − d + 1, . . . , d 2 }

⎪⎩

⎪ 4σgkp + 3 σ ,
2 52 2
k ∈ 2Z

⎪


⎩ 5σ 2 + 59 σ 2 , otherwise.
gkp 3

We remark that we have carefully determined σZH (k; d ) and σXH (k; d ) by thoroughly keeping track of how the circuit-level noise
propagates.
Then, moving on to step 2 of the GKP-stabilizer measurement, we update the noise vectors as described in Eqs. (B2) to (B5),
except that Eqs. (B3) and (B4) are applied for even and odd k (instead of odd and even k), respectively. Similarly as above,
while updating the data position and momentum noise vectors as prescribed in Eqs. (B6) and (B7), we assign the horizontal edge
weights as
⎧ # 2 $

⎪− log 2 p σ2gkp + 20 σ 2 R√π ξq(Ak) round 1

⎨ # H $
3
(Ak)
wZH (k) ← − log2 p σZ (k; d ) R√π ξq round 2 to round d (B27)

⎪ 

⎩ # H 2 $
− log2 p σZ (k; d ) − 2σgkp 2 − 20 σ 2 R√ ξ (Ak)
3 π q round d + 1,

for even k and


⎧ # 2 $

⎪− log 2 p 2σgkp + 20 σ 2 R√π ξ p(Ak) round 1

⎨ # H $
3
(Ak)
wXH (k) ← − log2 p σX (k; d ) R√π ξ p round 2 to round d (B28)

⎪ 

⎩ # H 2 $
− log2 p σX (k; d ) − 2σgkp 2 − 20 σ 2 R√ ξ (Ak)
3 π p round d + 1,

for odd k if the additional GKP information is used. Here, we use ξq(Ak) and ξ p(Ak) that are obtained after applying Eq. (B5). If on
the other hand the additional GKP information is not used, we assign the horizontal edge weights as
⎧  2

⎪ − log p 2σ + σ
20 2
round 1


2 err
H
gkp

3

wZH (k) ← − log2 perr σZ (k; d ) round 2 to round d (B29)



⎪ 

⎩ 2
− log2 perr σZH (k; d ) − 2σgkp
2 − 20 σ 2
3
round d + 1,
for even k and
⎧  2

⎪ − log 2 perr 2σgkp + 20 σ2 round 1

⎨ H
3

wXH (k) ← − log2 perr σX (k; d ) round 2 to round d (B30)





⎩ # H $2
− log2 perr σX (k; d ) − 2σgkp 2 − σ
20 2
3
round d + 1
for odd k. This way, all the horizontal edge weights are assigned.

012316-16
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

Vertical edge weights are assigned during surface code 4. Minimum-weight perfect matching decoding
stabilizer measurements: We follow steps 3–6 of surface Now, given the 3D spacetime graphs, the correction is
code stabilizer measurements and update the noise vectors as determined by using a minimum-weight perfect matching
described in Eqs. (B11) to (B19). Let wVZ () and wVX () be the decoding algorithm. More specifically, we do the following:
vertical edge weights of the Z-type and X -type 3D spacetime (1) Simulate d rounds of noisy stabilizer measurements
graphs corresponding to the th Z-type and X -type syn- followed by one round of ideal stabilizer measurements and
drome qubit. Then, after assigning the stabilizer values as in construct the Z-type and X -type 3D spacetime graphs as
Eq. (B20), we further assign the vertical edge weights as described above.
follows: (2) Highlight all vertices whose assigned stabilizer value
# $ is changed from the previous round. If the number of high-
wVZ () ← − log2 p σZV (; d ) R√π ξq(Zk) ,
lighted vertices is odd, highlight a boundary vertex. Thus, the
# $
wVX () ← − log2 p σXV (; d ) R√π ξ p(X k) , (B31) number of highlighted vertices is always even.
(3) For all pairs of highlighted Z-type (X -type) vertices,
while in rounds 1 to d for all  ∈ {1, . . . , d  = (d 2 − 1)/2}, if find the path with the minimum total weight. Then, save the
the additional GKP information is used. Here, we use ξq(Zk) minimum total weight and all edges in the path. Then, we
are left with a Z-type (X -type) complete graph of highlighted
and ξ p(X k) that are obtained after applying Eq. (B19) and
vertices, where the weight of the edge (v, w) is given by the
σZV (; d ) and σXV (; d ) are defined as minimum total weight of the path that connects v and w.
⎧ (4) Apply the minimum-weight perfect matching algo-


2 + 56 σ 2 ,
4σgkp  ∈ 2d  Z + 1 rithm [34,35] on the Z-type (X -type) complete graph of

⎪ 3

⎪  highlighted vertices. For all matched pairs of Z-type (X -type)

⎨ 7σgkp 2 + 107 σ 2 ,  ∈ 2d  Z + d  + 1
3 vertices, highlight all the Z-type (X -type) edges contained in
σZ (; d ) = 
V

⎪ 2 + 73
σ 2 ,  ∈ 2d  Z
the path that connects the matched vertices.

⎪ 4σgkp (5) Suppress all vertical edges and project the Z-type

⎪ 
3

⎩ 7σ 2 + 116 σ 2 , otherwise, (X -type) 3D spacetime graph onto the 2D plane. For each
gkp 3 Z-type (X -type) horizontal edge, count how many times it
⎧ was highlighted. If it is highlighted even times, do nothing.


2 +
4σgkp σ ,
56 2
 ∈ 2d  Z + d  Otherwise, apply the Pauli correction operator X̂gkp (Ẑgkp ) to

⎪ 3

⎪  the corresponding data GKP qubit. Equivalently, √ update the
⎪ 4σ 2 +
⎨ σ ,
73 2
 ∈ 2d  Z + d  + 1 quadrature noise as ξq(Dk) ← ξq(Dk) + π (ξ p(Dk) ← ξ p(Dk) +
σX (; d ) = 
V gkp 3 √
⎪ π ).


2 +
7σgkp σ ,
107 2
 ∈ 2d  Z

⎪ 3 Once the correction is done, we are left with

⎪  2
⎩ 7σ 2 + the data noise vectors ξqD = (ξq(D1) , . . . , ξq(Dd ) ) and
gkp 3
σ ,
116 2
otherwise. 2
ξ pD = (ξ p(D1) , . . . , ξ p(Dd ) ). Define total (ξqD ) ≡ dk=1 ξq(Dk)
2

(B32) 2
and total (ξ pD ) ≡ dk=1 ξ p(Dk) . Then, we determine that there is
Similarly as above, we have carefully determined σZV (; d )
and σXV (; d ) by thoroughly keeping track of how the circuit- logical X total ξqD = odd & total ξ pD = even

level noise propagates. If on the other hand the additional logical Z total ξqD = even & total ξ pD = odd (B34)
GKP information is not used, we assign the vertical edge
weights as logical Y total ξqD = odd & total ξ pD = odd
error. Otherwise if both total (ξqD ) and total (ξ pD ) are even,
wVZ () ← − log2 perr σZV (; d ) ,
there is no logical error.
wVX () ← − log2 perr σXV (; d ) . (B33) We use the Monte Carlo method to compute the logical
X, Y, Z error probability. In Fig. 6, we plot the logical X error
This way, all the vertical edge weights are assigned and probability obtained from 10 000–100 000 samples, which
thus we are left with the complete Z-type and X -type 3D is the same as the logical Z error probability. The number
spacetime graphs with all the horizontal and vertical edge of samples is determined such that statistical fluctuations are
weights assigned. negligible.

[1] S. Lloyd and S. L. Braunstein, Quantum Computation Over [4] A. P. Lund, T. C. Ralph, and H. L. Haselgrove, Fault-Tolerant
Continuous Variables, Phys. Rev. Lett. 82, 1784 (1999). Linear Optical Quantum Computing with Small-Amplitude
[2] H. Jeong and M. S. Kim, Efficient quantum computation using Coherent States, Phys. Rev. Lett. 100, 030503 (2008).
coherent states, Phys. Rev. A 65, 042305 (2002). [5] V. V. Albert, K. Noh, K. Duivenvoorden, D. J. Young, R. T.
[3] T. C. Ralph, A. Gilchrist, G. J. Milburn, W. J. Munro, and Brierley, P. Reinhold, C. Vuillot, L. Li, C. Shen, S. M. Girvin,
S. Glancy, Quantum computation with optical coherent states, B. M. Terhal, and L. Jiang, Performance and structure of single-
Phys. Rev. A 68, 042319 (2003). mode bosonic codes, Phys. Rev. A 97, 032346 (2018).

012316-17
KYUNGJOO NOH AND CHRISTOPHER CHAMBERLAND PHYSICAL REVIEW A 101, 012316 (2020)

[6] P. T. Cochrane, G. J. Milburn, and W. J. Munro, Macroscop- and S. M. Girvin, Bias-preserving gates with stabilized cat
ically distinct quantum-superposition states as a bosonic code qubits, arXiv:1905.00450.
for amplitude damping, Phys. Rev. A 59, 2631 (1999). [23] D. K. Tuckett, S. D. Bartlett, and S. T. Flammia, Ultrahigh Error
[7] M. H. Michael, M. Silveri, R. T. Brierley, V. V. Albert, J. Threshold for Surface Codes with Biased Noise, Phys. Rev.
Salmilehto, L. Jiang, and S. M. Girvin, New Class of Quantum Lett. 120, 050505 (2018).
Error-Correcting Codes for A Bosonic Mode, Phys. Rev. X 6, [24] D. K. Tuckett, C. T. Chubb, S. Bravyi, S. D. Bartlett, and S. T.
031006 (2016). Flammia, Tailoring Surface Codes for Highly Biased Noise,
[8] D. Gottesman, A. Kitaev, and J. Preskill, Encoding a qubit in an Phys. Rev. X 9, 041031 (2019).
oscillator, Phys. Rev. A 64, 012310 (2001). [25] D. K. Tuckett, S. D. Bartlett, S. T. Flammia, and B. J. Brown,
[9] Z. Leghtas, S. Touzard, I. M. Pop, A. Kou, B. Vlastakis, A. Fault-tolerant thresholds for the surface code in excess of 5%
Petrenko, K. M. Sliwa, A. Narla, S. Shankar, M. J. Hatridge, M. under biased noise, arXiv:1907.02554.
Reagor, L. Frunzio, R. J. Schoelkopf, M. Mirrahimi, and M. H. [26] K. Fukui, A. Tomita, and A. Okamoto, Analog Quantum Error
Devoret, Confining the state of light to a quantum manifold by Correction with Encoding a Qubit into an Oscillator, Phys. Rev.
engineered two-photon loss, Science 347, 853 (2015). Lett. 119, 180507 (2017).
[10] N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas, [27] K. Fukui, A. Tomita, and A. Okamoto, Tracking quantum error
B. Vlastakis, Y. Liu, L. Frunzio, S. M. Girvin, L. Jiang, M. correction, Phys. Rev. A 98, 022326 (2018).
Mirrahimi, M. H. Devoret, and R. J. Schoelkopf, Extending the [28] Y. Wang, Ph.D. thesis, Master’s thesis, RWTH Aachen Univer-
lifetime of a quantum bit with error correction in superconduct- sity, 2017 (unpublished).
ing circuits, Nature (London) 536, 441 (2016). [29] K. Fukui, A. Tomita, A. Okamoto, and K. Fujii, High-Threshold
[11] S. Touzard, A. Grimm, Z. Leghtas, S. O. Mundhada, P. Fault-Tolerant Quantum Computation with Analog Quantum
Reinhold, C. Axline, M. Reagor, K. Chou, J. Blumoff, K. M. Error Correction, Phys. Rev. X 8, 021054 (2018).
Sliwa, S. Shankar, L. Frunzio, R. J. Schoelkopf, M. Mirrahimi, [30] C. Vuillot, H. Asasi, Y. Wang, L. P. Pryadko, and B. M. Terhal,
and M. H. Devoret, Coherent Oscillations Inside a Quantum Quantum error correction with the toric Gottesman-Kitaev-
Manifold Stabilized by Dissipation, Phys. Rev. X 8, 021005 Preskill code, Phys. Rev. A 99, 032344 (2019).
(2018). [31] N. C. Menicucci, Fault-Tolerant Measurement-Based Quantum
[12] L. Hu, Y. Ma, W. Cai, X. Mu, Y. Xu, W. Wang, Y. Wu, H. Wang, Computing with Continuous-Variable Cluster States, Phys. Rev.
Y. P. Song, C. L. Zou, S. M. Girvin, L-M. Duan, and L. Sun, Lett. 112, 120504 (2014).
Quantum error correction and universal gate set operation on a [32] K. Fukui, High-threshold fault-tolerant quantum computa-
binomial bosonic logical qubit, Nat. Phys. 15, 503 (2019). tion with the GKP qubit and realistically noisy devices,
[13] P. Campagne-Ibarcq, A. Eickbusch, S. Touzard, E. Zalys- arXiv:1906.09767.
Geller, N. E. Frattini, V. V. Sivak, P. Reinhold, S. Puri, S. [33] R. Raussendorf, D. E. Browne, and H. J. Briegel, Measurement-
Shankar, R. J. Schoelkopf, L. Frunzio, M. Mirrahimi, and M. H. based quantum computation on cluster states, Phys. Rev. A 68,
Devoret, A stabilized logical quantum bit encoded in grid states 022312 (2003).
of a superconducting cavity, arXiv:1907.12487. [34] J. Edmonds, Paths, trees, and flowers, Can. J. Math. 17, 449
[14] A. Grimm, N. E. Frattini, S. Puri, S. O. Mundhada, S. (1965).
Touzard, M. Mirrahimi, S. M. Girvin, S. Shankar, and M. H. [35] J. Edmonds, Maximum matching and a polyhedron with 0,1
Devoret, The Kerr-cat qubit: Stabilization, readout, and gates, vertices, J. Res. Natl. Bur. Stand., Sect. B 69 B, 125 (1965).
arXiv:1907.12131. [36] D. S. Wang, A. G. Fowler, and L. C. L. Hollenberg, Surface
[15] C. Flühmann, V. Negnevitsky, M. Marinelli, and J. P. Home, code quantum computing with error rates over 1%, Phys. Rev.
Sequential Modular Position and Momentum Measurements of A 83, 020302(R) (2011).
a Trapped ion Mechanical Oscillator, Phys. Rev. X 8, 021001 [37] C. Chamberland, G. Zhu, T. J. Yoder, J. B. Hertzberg, and A. W.
(2018). Cross, Topological and subsystem codes on low-degree graphs
[16] C. Flühmann, T. L. Nguyen, M. Marinelli, V. Negnevitsky, with flag qubits, arXiv:1907.09528.
K. Mehta, and J. P. Home, Encoding a qubit in a trapped-ion [38] H. Bombin and M. A. Martin-Delgado, Optimal resources
mechanical oscillator, Nature (London) 566, 513 (2019). for topological two-dimensional stabilizer codes: Comparative
[17] C. Flühmann and J. P. Home, Direct characteristic-function study, Phys. Rev. A 76, 012305 (2007).
tomography of quantum states of the trapped-ion motional [39] Y. Tomita and K. M. Svore, Low-distance surface codes under
oscillator, arXiv:1907.06478. realistic quantum noise, Phys. Rev. A 90, 062320 (2014).
[18] S. B. Bravyi and A. Yu. Kitaev, Quantum codes on a lattice with [40] B. C. Travaglione and G. J. Milburn, Preparing encoded states
boundary, arXiv:quant-ph/9811052. in an oscillator, Phys. Rev. A 66, 052322 (2002).
[19] E. Dennis, A. Kitaev, A. Landahl, and J. Preskill, Topological [41] S. Pirandola, S. Mancini, D. Vitali, and P. Tombesi, Construct-
quantum memory, J. Math. Phys. 43, 4452 (2002). ing finite-dimensional codes with optical continuous variables,
[20] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N. Cleland, Europhys. Lett. 68, 323 (2004).
Surface codes: Towards practical large-scale quantum computa- [42] S. Pirandola, S. Mancini, D. Vitali, and P. Tombesi, Generat-
tion, Phys. Rev. A 86, 032324 (2012). ing continuous variable quantum codewords in the near-field
[21] J. Guillaud and M. Mirrahimi, Repetition Cat-Qubits: atomic lithography, J. Phys. B: At., Mol. Opt. Phys. 39, 997
Fault-Tolerant Quantum Computation with Highly Reduced (2006).
Overhead, Phys. Rev. X 9, 041053 (2019). [43] H. M. Vasconcelos, L. Sanz, and S. Glancy, All-optical genera-
[22] S. Puri, L. St-Jean, J. A. Gross, A. Grimm, N. E. Frattini, P. S. tion of states for “encoding a qubit in an oscillator”, Opt. Lett.
Iyer, A. Krishna, S. Touzard, L. Jiang, A. Blais, S. T. Flammia, 35, 3261 (2010).

012316-18
FAULT-TOLERANT BOSONIC QUANTUM ERROR … PHYSICAL REVIEW A 101, 012316 (2020)

[44] B. M. Terhal and D. Weigand, Encoding a qubit into a cavity [59] B. M. Terhal, Quantum error correction for quantum memories,
mode in circuit QED using phase estimation, Phys. Rev. A 93, Rev. Mod. Phys. 87, 307 (2015).
012315 (2016). [60] C. Chamberland, P. Iyer, and D. Poulin, Fault-tolerant quantum
[45] K. R. Motes, B. Q. Baragiola, A. Gilchrist, and N. C. computing in the Pauli or Clifford frame with slow error diag-
Menicucci, Encoding qubits into oscillators with atomic ensem- nostics, Quantum 2, 43 (2018).
bles and squeezed light, Phys. Rev. A 95, 053819 (2017). [61] K. E. Cahill and R. J. Glauber, Ordered expansions in boson
[46] D. J. Weigand and B. M. Terhal, Generating grid states from amplitude operators, Phys. Rev. 177, 1857 (1969).
Schrödinger-cat-states without postselection, Phys. Rev. A 97, [62] G. Pantaleoni, B. Q. Baragiola, and N. C. Menicucci, Modular
022341 (2018). bosonic subsystem codes, arXiv:1907.08210.
[47] J. M. Arrazola, T. R. Bromley, J. Izaac, C. R. Myers, K. Brádler, [63] I. Tzitrin, J. E. Bourassa, N. C. Menicucci, and K. K. Sabapathy,
and N. Killoran, Machine learning method for state preparation Towards practical qubit computation using approximate error-
and gate synthesis on photonic quantum computers, Quantum correcting grid states, arXiv:1910.03673.
Sci. Technol. 4, 024004 (2019). [64] T. Matsuura, H. Yamasaki, and M. Koashi, On the
[48] D. Su, C. R. Myers, and K. K. Sabapathy, Conversion of equivalence of approximate Gottesman-Kitaev-Preskill
Gaussian states to non-Gaussian states using photon number- codes, arXiv:1910.08301.
resolving detectors, Phys. Rev. A 100, 052301 (2019). [65] J. Emerson, M. Silva, O. Moussa, C. Ryan, M. Laforest, J.
[49] M. Eaton, R. Nehra, and O. Pfister, Non-Gaussian and Baugh, D. G. Cory, and R. Laflamme, Symmetrized characteri-
Gottesman–Kitaev–Preskill state preparation by photon catal- zation of noisy quantum processes, Science 317, 1893 (2007).
ysis, New J. Phys. 21, 113034 (2019). [66] M.-D. Choi, Completely positive linear maps on complex ma-
[50] Y. Shi, C. Chamberland, and A. Cross, Fault-tolerant prepa- trices, Lin. Alg. Appl. 10, 285 (1975).
ration of approximate GKP states, New J. Phys. 21, 093007 [67] K. Noh, V. V. Albert, and L. Jiang, Quantum capacity bounds
(2019). of Gaussian thermal loss channels and achievable rates with
[51] D. J. Weigand and B. M. Terhal, Realizing modular quadrature Gottesman-Kitaev-Preskill codes, IEEE Trans. Inf. Theory 65,
measurements via a tunable photon-pressure coupling in circuit- 2563 (2019).
QED, arXiv:1909.10075. [68] Y. Y. Gao, B. J. Lester, Y. Zhang, C. Wang, S. Rosenblum,
[52] D. Gottesman and I. L. Chuang, Demonstrating the viability of L. Frunzio, L. Jiang, S. M. Girvin, and R. J. Schoelkopf,
universal quantum computation using teleportation and single- Programmable Interference between Two Microwave Quantum
qubit operations, Nature (London) 402, 390 (1999). Memories, Phys. Rev. X 8, 021073 (2018).
[53] S. Bravyi and A. Kitaev, Universal quantum computation with [69] Y. Zhang, B. J. Lester, Y. Y. Gao, L. Jiang, R. J. Schoelkopf,
ideal Clifford gates and noisy ancillas, Phys. Rev. A 71, 022316 and S. M. Girvin, Engineering bilinear mode coupling in circuit
(2005). QED: Theory and experiment, Phys. Rev. A 99, 012314 (2019).
[54] B. Q. Baragiola, G. Pantaleoni, R.l N. Alexander, A. Karanjai, [70] N. E. Frattini, U. Vool, S. Shankar, A. Narla, K. M. Sliwa,
and N. C. Menicucci, All-Gaussian Universality and Fault and M. H. Devoret, 3-wave mixing Josephson dipole element,
Tolerance with the Gottesman-Kitaev-Preskill Code, Phys. Rev. Appl. Phys. Lett. 110, 222603 (2017).
Lett. 123, 200502 (2019). [71] P. Aliferis and B. M. Terhal, Fault-tolerant quantum compu-
[55] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. Schuster, tation for local leakage faults, Quantum Info. Comput. 7, 139
J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and R. J. (2007).
Schoelkopf, Charge-insensitive qubit design derived from the [72] A. G. Fowler, Coping with qubit leakage in topological codes,
Cooper pair box, Phys. Rev. A 76, 042319 (2007). Phys. Rev. A 88, 042308 (2013).
[56] J. A. Schreier, A. A. Houck, J. Koch, D. I. Schuster, B. R. [73] M. Suchara, A. W. Cross, and J. M. Gambetta, Leakage suppres-
Johnson, J. M. Chow, J. M. Gambetta, J. Majer, L. Frunzio, sion in the toric code, in 2015 IEEE International Symposium
M. H. Devoret, S. M. Girvin, and R. J. Schoelkopf, Suppressing on Information Theory (ISIT) (IEEE, 2015), pp. 1119–1123.
charge noise decoherence in superconducting charge qubits, [74] N. C. Brown, M. Newman, and K. R. Brown, Handling leakage
Phys. Rev. B 77, 180502(R) (2008). with subsystem codes, New J. Phys. 21, 073055 (2019).
[57] E. Knill, Quantum computing with realistically noisy devices, [75] H. F. Trotter, On the product of semi-groups of operators, Proc.
Nature 434, 39 (2005). Am. Math. Soc. 10, 545 (1959).
[58] D. P. DiVincenzo and P. Aliferis, Effective Fault-Tolerant Quan- [76] C. Weedbrook, S. Pirandola, R. García-Patrón, N. J. Cerf,
tum Computation with Slow Measurements, Phys. Rev. Lett. 98, T. C. Ralph, J. H. Shapiro, and S. Lloyd, Gaussian quantum
020501 (2007). information, Rev. Mod. Phys. 84, 621 (2012).

012316-19

You might also like