PreprintPDF Available

Entangling four logical qubits beyond break-even in a nonlocal code

Authors:
Preprints and early-stage research may not have been peer reviewed yet.

Abstract

Quantum error correction protects logical quantum information against environmental decoherence by encoding logical qubits into entangled states of physical qubits. One of the most important near-term challenges in building a scalable quantum computer is to reach the break-even point, where logical quantum circuits on error-corrected qubits achieve higher fidelity than equivalent circuits on uncorrected physical qubits. Using Quantinuum's H2 trapped-ion quantum processor, we encode the GHZ state in four logical qubits with fidelity $ 99.5 \pm 0.15 \% \le F \le 99.7 \pm 0.1\% $ (after postselecting on over 98% of outcomes). Using the same quantum processor, we can prepare an uncorrected GHZ state on four physical qubits with fidelity $97.8 \pm 0.2 \% \le F\le 98.7\pm 0.2\%$. The logical qubits are encoded in a $[\![ 25,4,3 ]\!]$ Tanner-transformed long-range-enhanced surface code. Logical entangling gates are implemented using simple swap operations. Our results are a first step towards realizing fault-tolerant quantum computation with logical qubits encoded in geometrically nonlocal quantum low-density parity check codes.
Entangling four logical qubits beyond break-even in a nonlocal code
Yifan Hong,1, Elijah Durso-Sabina,2David Hayes,2and Andrew Lucas1
1Department of Physics and Center for Theory of Quantum Matter,
University of Colorado, Boulder, CO 80309, USA
2Quantinuum, Broomfield, CO 80021, USA
(Dated: June 6, 2024)
Quantum error correction protects logical quantum information against environmental decoher-
ence by encoding logical qubits into entangled states of physical qubits. One of the most important
near-term challenges in building a scalable quantum computer is to reach the break-even point,
where logical quantum circuits on error-corrected qubits achieve higher fidelity than equivalent cir-
cuits on uncorrected physical qubits. Using Quantinuum’s H2 trapped-ion quantum processor, we
encode the GHZ state in four logical qubits with fidelity 99.5±0.15% F99.7±0.1% (af-
ter postselecting on over 98% of outcomes). Using the same quantum processor, we can prepare
an uncorrected GHZ state on four physical qubits with fidelity 97.8±0.2% F98.7±0.2%.
The logical qubits are encoded in a J25,4,3KTanner-transformed long-range-enhanced surface code.
Logical entangling gates are implemented using simple swap operations. Our results are a first step
towards realizing fault-tolerant quantum computation with logical qubits encoded in geometrically
nonlocal quantum low-density parity check codes.
Introduction. It is widely believed that fault toler-
ance will be necessary for quantum computers to outper-
form classical computers on large-scale problems. De-
coherence, if left ignored, will eventually destroy entan-
glement and any quantum advantage in computation. A
critical milestone for the development of a quantum com-
puter will, therefore, be the protection of logical qubits
and the implementation of logical gates whose error rates
are below that of physical qubits and circuits which are
not error-corrected. When logical qubits are more robust
than physical qubits, we have reached break-even.
Since the discovery of topological quantum codes [1,2]
and their promise to preserve quantum information for
arbitrarily long times with a relatively high tolerance for
errors [3,4], there have been immense efforts to imple-
ment such codes in current architectures [59]. Unfor-
tunately, the resource requirements for large-scale, fault-
tolerant quantum computation using topological codes
with current hardware specifications may be enormous
[10].
More generally, it is known that these overheads can-
not be substantially improved in planar architectures
(codes) with only nearest-neighbor connectivity [11]. To
avoid this fundamental challenge, quantum low-density
parity-check (LDPC) codes [1219] have been developed
to significantly reduce the resource overheads for error
correction [20,21]. To avoid the constraints of Ref. [11],
these codes will necessarily require long-range connectiv-
ity [22,23]. Platforms with such capabilities may be able
to significantly reduce the overhead for quantum error
correction, and a number of recent proposals have been
made for how to incorporate limited long-range connec-
tivity into hardware [2427] to ameliorate the constraints
of spatial locality.
yifan.hong@colorado.edu
Thus far, we are not aware of any explicit implemen-
tation of a nonlocal LDPC code in any quantum hard-
ware. Other recent demonstrations of break-even opera-
tions we are aware of include the use of a small 7-qubit
Steane code [8,28] and the use of concatenated error
detection codes [9], neither of which achieve the high-
encoding rates of qLDPC codes. While such codes may
improve performance, to truly avoid large resource over-
heads in a scalable way, it is crucial to use a nonlocal code
which protects logical qubits in a more complex manner,
yet which is also amenable to quantum hardware.
This work provides a proof-of-principle demonstra-
tion that nonlocal codes can improve the performance
of present-day quantum computers. We have imple-
mented a J25,4,3Kquantum LDPC code (encoding 4 log-
ical qubits in 25 physical qubits, with the smallest logical
operator acting on 3 physical qubits) on Quantinuum’s
H2 trapped-ion quantum computer [29], which was con-
figured with 32 physical qubits at the time of the exper-
iments. This encoding enables us to prepare an error-
corrected, logical 4-qubit entangled GHZ state beyond
the break-even point, with mild postselection overhead,
for the first time. Using the existing hardware, it would
not have been possible to attempt such a demonstration,
even in principle, by using four surface code patches (with
36 total physical) to encode the logical qubits. This work
highlights the importance of choosing good hardware-
suited codes with efficient logical gates. The long-range-
enhanced surface code [26] underlying our methods will
be scalable to larger trapped-ion QCCD processors [30
33] as well as to neutral atom arrays [8,3437] for the
foreseeable future.
Quantum encoding. The quantum LDPC code we
use is a Calderbank-Shor-Steane (CSS) stabilizer code
[3840], which is defined by a stabilizer group of Pauli
strings that mutually commute and act trivially on log-
ical codewords. In a CSS code, these Pauli strings are
generated by products of strictly Xor Zoperators act-
arXiv:2406.02666v1 [quant-ph] 4 Jun 2024
2
ing on nqubits. Given the Pauli (anti)commutation re-
lations, it is helpful to organize data about the stabilizer
group in a pair of matrices HX,Z with F2={0,1}coeffi-
cients. Each row of the matrix HX, which in turn has n
columns, fixes one of the stabilizers of the code to be a
product of Pauli Xs on all qubits with a 1 in the corre-
sponding column; HZis defined similarly. To be a valid
code, we require Xand Zstabilizers to commute, which
means HXHT
Z= 0 (here multiplication/addition are mod
2).
A convenient way to build a quantum code is the hy-
pergraph product [13] of a classical linear code with itself.
For an input m×nclassical parity-check matrix H, the
resulting CSS stabilizer matrices take the form
HX=H1n|1mHT(1a)
HZ=1nH|HT1m.(1b)
If the classical input code has parameters [n, k, d] with
m=nkparity checks (meaning that there are nclas-
sical bits encoding klogicals, with code distance d), then
so long as the classical code has no redundant parity
checks, the resulting quantum CSS hypergraph product
code has parameters Jn2+ (nk)2, k2, dK. The geomet-
rical layout of the quantum code can be understood as
a Cartesian graph product between those of its classical
input codes. While the hypergraph product can produce
codes with constant rate (kn), the distance scaling
is sublinear (dn). Nonetheless, this distance is pre-
served under standard syndrome extraction techniques
with circuit-level noise [41], a feature not supported by
other quantum codes such as the Steane code [40].
We now summarize the construction of our quantum
LDPC code, with a graphical depiction in Fig. 1. For
the classical input code, we first pick a [3,2,2] parity
code consisting of a single, global parity check on all
three physical bits. We then concatenate two of the
physical bits with a [2,1,2] repetition code to obtain a
[5,2,3] code. We then take the hypergraph product of
this [5,2,3] code with itself to obtain a J34,4,3KCSS
code. Finally, we perform a quantum Tanner transforma-
tion [19] to reduce the number of physical qubits from 34
to 25, while preserving both kand d. We note that the
quantum Tanner transformation does not preserve the
generic circuit-level distance property mentioned earlier;
however, we can still mitigate the effects of correlated er-
rors by choosing a clever pattern for syndrome extraction
like that for the rotated surface code [42]. The order of
CNOTs is carefully chosen so that all correlated ancilla
(hook) errors end up “against the grain” of the logical
operators, and so d= 3 errors on the ancilla qubits are
necessary to affect the codespace. This code is a minimal
example of a more general family of long-range-enhanced
surface codes [26], which possess the fewest nonlocal sta-
bilizers necessary to avoid the constraints of locality [11],
while adding O(1) logical qubits.
The standard protocol to prepare the GHZ state on 4
qubits is to initialize the qubits in |+000and then apply
(a)
(b) (c)
FIG. 1. The code construction is illustrated. Circles and
red (blue) squares denote qubits and X(Z)-type stabilizers
respectively, with edges representing their connectivity. Ma-
genta lines denote long-range interactions. (a) A quantum
Tanner transformation reduces the number of physical qubits
from 34 to 25. (b) The support of the logical Xand Zop-
erators for logical qubits 1-4 are drawn with shaded boxes,
solid lines, dashed lines, and striped boxes respectively. (c)
The order of CNOT gates for distance-preserving syndrome
extraction is depicted for a selection of stabilizers.
CNOTs between the |+qubit (control) and the rest (tar-
get); here Z|0=|0and X|+=|+. Since our code
is CSS and has a distance-preserving syndrome extrac-
tion circuit, we can perform a transversal initialization
into the logical |¯
0¯
0¯
0¯
0state by initializing all physical
qubits in |0nand then measuring all X-type stabiliz-
ers. The stabilizer measurement can be performed by
initializing ancillas in the |+state, performing CNOT
gates between the ancillas (control) and physical qubits
(target) according to the pattern in Fig. 1c, and finally
measuring all ancillas in the Xbasis. We then prepare
the third logical qubit in the |¯
+state by measuring one
of its logical ¯
X3operators three times and postselecting
on the agreement of all three outcomes in order to mit-
igate the effects of readout error. After postselection, a
logical ¯
Z3is then applied (offline) if the agreed outcome is
1. After preparing the logical |¯
0¯
0¯
+¯
0state, we perform
the desired logical CNOT gates by permuting the data
qubits a feature inherited from the [3,2,2] parity code.
This relabeling step can be done in software in our par-
ticular experiment. In particular, two combinations of
“fold-swaps” are sufficient: see Fig. 2. Finally, we per-
form a transversal measurement of all logical qubits in
either the Xor Zbasis by measuring all physical qubits
in that basis. This transversal measurement is then used
to reconstruct a final syndrome which is not subject to
3
(a)
(c) (d)
(b)
FIG. 2. The procedures for preparing the physical and logical
4-qubit GHZ states are illustrated. (a) The quantum circuit
which prepares the physical GHZ state. (b) The circuit which
prepares the logical GHZ state. (c-d) A combination of fold
swaps implements the above logical circuit while preserving
the stabilizer group.
circuit-level noise and can be used to correct errors. All
error correction is performed “in software” by tracking
the Pauli frame: the final logical measurement outcomes
are modified according to a decoding algorithm. We
use belief-propagation with ordered statistics decoding
(BP+OSD) [43,44], specifically the “minimum-sum” (10
iterations) and “combination sweep” (λ= 14) strategies
of BP and OS respectively, to infer physical errors based
on the syndrome measurement outcomes.
We now discuss the fault tolerance (or lack thereof)
of each of the aforementioned steps in our logical GHZ
protocol. As previously shown, our syndrome extraction
circuit is distance-preserving and so we do not need to
worry about ancilla hook errors. However, since we only
perform a single round of syndrome extraction, one may
worry about the effects of ancilla measurement errors.
For example, in a large surface code, a single flipped
check in the middle of the code may cause a matching
decoder to pair it to the boundary, resulting in a large
error string. Fortunately, for our distance three code, any
single flipped X-check can be caused by a single-qubit Z
error, and so we do not suffer from this type of error
propagation. For the logical ¯
X3measurement, a single
Zerror on a physical qubit prior to the measurements
can spoil the result, and so this step is not fault tolerant.
We note that a fault-tolerant procedure of measuring ¯
X3
can be done using generalized lattice surgery techniques
[45], though this would add considerable overhead (and
requires more physical qubits than were available in H2
at the time) and so we leave it to future work. The phys-
ical qubit permutations to implement the fold-swaps can
be achieved by a simple relabeling in software and so is
fault tolerant by definition. Lastly, the final measure-
GHZ experiment ¯zmismatch (%) ¯xmismatch (%)
Physical 1.3±0.2 0.9±0.1
Logical (No QEC) 5.7±0.5 2.9±0.3
Logical (with QEC) 0.3±0.1 0.2±0.1
TABLE 1. Fidelity of GHZ state preparation in experiment.
The second column displays the percentage of runs when the
four Zmeasurement outcomes disagreed. The third column
displays the percentage of runs when the product of the four
Xmeasurements was 1. Standard errors for the data are
also shown.
ment of all physical qubits is transversal and so is also
fault tolerant.
Experimental results. The experiments described in
this work used the H2 trapped-ion processor described
in Ref. [29]. The H2 processor was configured with 32
trapped-ion qubits, all of which can be directly gated
with each other using ion-transport operations to pair
any two-qubits in a single potential well a crucial fea-
ture for our J25,4,3Kcode. Physical-level entangling op-
erations are implemented using a Mølmer-Sørensen in-
teraction [46] driven by lasers directed to the four gating
regions. Likewise, physical level single-qubit gates, reset,
and measurement all take place in these same regions
using additional lasers systems. At the time of these ex-
periments, the typical error rates in H2 were as follows:
<2×103two-qubit gate error, 3×105single-qubit
gate error, and 2×103state preparation and mea-
surement (SPAM) error.
For the physical GHZ experiment, we run the circuit in
Fig. 2a for 10,000 shots, measuring all physical qubits in
the Zbasis for 5000 shots and in the Xbasis for the other
half. For the logical GHZ experiment, we run the circuit
in Fig. 2b for 5000 shots, with 2500 shots each dedicated
to logical ¯
Xand ¯
Zreadout. Uncertainties in the data are
computed using the standard error σ=pp(1 p)/N,
where Nis the number of shots. For a perfect GHZ
state, we expect all Zmeasurement outcomes to agree
and the product of all Xmeasurement outcomes to have
even parity. The mismatch rates in the experimental
data from these theoretical predictions are tabulated in
Table 1. In addition to the error-corrected logical results,
we also give the uncorrected logical results which are ob-
tained by simply ignoring the syndrome values (i.e. no
feedback).
Since the GHZ state is a stabilizer state with gener-
ators {Z1Z2, Z2Z3, Z3Z4, X1X2X3X4}, we can compute
simple upper and lower bounds on the GHZ fidelity of
our experimental state ρ. Because |GHZis a pure state,
the fidelity [47] takes the form
FGHZ(ρ) = GHZ|ρ|GHZ ρGHZ .(2)
We now compute an upper bound for the experimental
fidelities. First, notice that the two projectors P±
zonto
whether the Zmeasurement outcomes agree (P+
z= (1+
Z1Z2)(1+Z2Z3)(1+Z3Z4)/8) or disagree (P
z=1P+
z)
4
form a projection-valued measure that splits the Hilbert
space into two orthogonal subspaces. Using the fact that
P+
zGHZ = 1, i.e. the GHZ state strictly lives in the P+
z
subspace, we can upper bound the fidelity (2) by
FGHZ(ρ)trP+
zρ,(3)
which corresponds to the second column in Table 1. For
the lower bound, observe that the GHZ state also strictly
resides in the subspace of P+
x(1+X1X2X3X4)/2 such
that its density matrix takes the form |GHZ⟩⟨GHZ|=
P+
zP+
x. We can then use the union bound from proba-
bility theory to upper bound the probabilities of being
orthogonal to GHZ, which provides a lower bound to the
fidelity (2):
FGHZ(ρ)1trP
zρtrP
xρ.(4)
The first average corresponds to the fraction of runs
where the Zoutcomes disagree and is given by the second
column in Table 1. The contribution from the XX X X
generator is given by the third column. For the physical
and (error-corrected) logical states ρ, ¯ρ, we hence obtain
fidelity bounds of
97.8±0.2% FGHZ(ρ)98.7±0.2% (5a)
99.5±0.15% FGHZ( ¯ρ)99.7±0.1% .(5b)
The statistical significance between the physical upper
bound and the logical lower bound is 3.6σ, assuming
statistically independent uncertainties.
Outlook. We have prepared a logical GHZ state on
four qubits with a higher fidelity than its physical coun-
terpart. The key ingredient is a quantum LDPC code
whose long-range interactions enable (i) a compact en-
coding of logical qubits and (ii) efficient logical entan-
gling gates via qubit permutations. Having demonstrated
the ability to prepare specific quantum entangled logical
states beyond break-even, an immediate near-term ex-
perimental goal will be to use our encoding (or others)
to demonstrate break-even on larger circuits involving
multiple logical qubits, eventually including circuits with
non-Clifford gates.
Our work highlights an advantage of using quantum
LDPC codes in existing hardware over conventional topo-
logical codes like the surface and color codes. Of course,
a major feature of the topological codes is that they are
equipped with transversal implementations of the Clif-
ford group. A big challenge in the field is to construct
nonlocal quantum LDPC codes with hardware-efficient
logical gates; some progress using hypergraph product
codes has been previously made [48,49]. We hope that
this work stimulates further progress in this direction,
especially since it is already known that LDPC codes
can significantly reduce the overhead for fault-tolerant
quantum computation when functioning just as a mem-
ory block [45]. On a separate front, nonlocal codes de-
rived from concatenating quantum Hamming codes also
show promise for low-overhead fault tolerance [50,51].
Since these codes are not LDPC, they require more com-
plicated gadgets for fault tolerance. A worthwhile future
direction is to perform a detailed cost-benefit analysis be-
tween LDPC and concatenated architectures on different
hardware.
Acknowledgements. This work was supported in part
by the Alfred P. Sloan Foundation under Grant FG-2020-
13795 (AL), by the Air Force Office of Scientific Re-
search under Grant FA9550-24-1-0120 (YH, AL), and by
the Office of Naval Research via Grant N00014-23-1-2533
(YH, AL). Additionally, we thank the hardware team at
Quantinuum for making these experiments possible.
[1] A.Yu. Kitaev, “Fault-tolerant quantum computation by
anyons,” Annals of Physics 303, 2–30 (2003).
[2] H. Bombin and M. A. Martin-Delgado, “Topological
quantum distillation,” Phys. Rev. Lett. 97, 180501
(2006).
[3] Eric Dennis, Alexei Kitaev, Andrew Landahl, and
John Preskill, “Topological quantum memory,” Journal
of Mathematical Physics 43, 4452–4505 (2002).
[4] Andrew J. Landahl, Jonas T. Anderson, and Patrick R.
Rice, “Fault-tolerant quantum computing with color
codes,” (2011), arXiv:1108.5738 [quant-ph].
[5] Lukas Postler, Sascha Heußen, Ivan Pogorelov, Manuel
Rispler, Thomas Feldker, Michael Meth, Christian D.
Marciniak, Roman Stricker, Martin Ringbauer, Rainer
Blatt, Philipp Schindler, Markus uller, and Thomas
Monz, “Demonstration of fault-tolerant universal quan-
tum gate operations,” Nature 605, 675–680 (2022),
arXiv:2111.12654 [quant-ph].
[6] C. Ryan-Anderson, J. G. Bohnet, K. Lee, D. Gresh,
A. Hankin, J. P. Gaebler, D. Francois, A. Chernoguzov,
D. Lucchetti, N. C. Brown, T. M. Gatterman, S. K. Halit,
K. Gilmore, J. A. Gerber, B. Neyenhuis, D. Hayes, and
R. P. Stutz, “Realization of real-time fault-tolerant quan-
tum error correction,” Phys. Rev. X 11, 041058 (2021).
[7] Google Quantum AI, “Suppressing quantum errors by
scaling a surface code logical qubit,” Nature 614, 676–
681 (2023).
[8] Dolev Bluvstein, Simon J. Evered, Alexandra A. Geim,
Sophie H. Li, Hengyun Zhou, Tom Manovitz, Sepehr
Ebadi, Madelyn Cain, Marcin Kalinowski, Dominik
Hangleiter, J. Pablo Bonilla Ataides, Nishad Maskara,
Iris Cong, Xun Gao, Pedro Sales Rodriguez, Thomas
Karolyshyn, Giulia Semeghini, Michael J. Gullans,
Markus Greiner, Vladan Vuleti´c, and Mikhail D. Lukin,
“Logical quantum processor based on reconfigurable
atom arrays,” Nature 626, 58–65 (2023).
[9] M. P. da Silva, C. Ryan-Anderson, J. M. Bello-Rivas,
A. Chernoguzov, J. M. Dreiling, C. Foltz, F. Frachon,
J. P. Gaebler, T. M. Gatterman, L. Grans-Samuelsson,
D. Hayes, N. Hewitt, J. Johansen, D. Lucchetti, M. Mills,
5
S. A. Moses, B. Neyenhuis, A. Paz, J. Pino, P. Siegfried,
J. Strabley, A. Sundaram, D. Tom, S. J. Wernli, M. Zan-
ner, R. P. Stutz, and K. M. Svore, “Demonstration of
logical qubits and repeated error correction with better-
than-physical error rates,” (2024), arXiv:2404.02280
[quant-ph].
[10] Craig Gidney and Martin Eker˚a, “How to factor 2048 bit
rsa integers in 8 hours using 20 million noisy qubits,”
Quantum 5, 433 (2021).
[11] Sergey Bravyi, David Poulin, and Barbara Terhal,
“Tradeoffs for reliable quantum information storage in
2d systems,” Phys. Rev. Lett. 104, 050503 (2010).
[12] Nikolas P. Breuckmann and Jens Niklas Eberhardt,
“Quantum low-density parity-check codes,” PRX Quan-
tum 2(2021), 10.1103/prxquantum.2.040101.
[13] Jean-Pierre Tillich and Gilles Zemor, “Quantum ldpc
codes with positive rate and minimum distance pro-
portional to the square root of the blocklength,” IEEE
Transactions on Information Theory 60, 1193–1202
(2014).
[14] Matthew B. Hastings, Jeongwan Haah, and Ryan
O’Donnell, “Fiber bundle codes: Breaking the
n1/2polylog(n) barrier for quantum ldpc codes,”
(2020), arXiv:2009.03921 [quant-ph].
[15] Pavel Panteleev and Gleb Kalachev, “Quantum ldpc
codes with almost linear minimum distance,” IEEE
Transactions on Information Theory 68, 213–229 (2022).
[16] Nikolas P. Breuckmann and Jens N. Eberhardt, “Bal-
anced product quantum codes,” IEEE Transactions on
Information Theory 67, 6653–6674 (2021).
[17] Pavel Panteleev and Gleb Kalachev, “Asymptotically
good quantum and locally testable classical ldpc codes,”
(2022), arXiv:2111.03654 [cs.IT].
[18] Irit Dinur, Min-Hsiu Hsieh, Ting-Chun Lin, and Thomas
Vidick, “Good quantum ldpc codes with linear time de-
coders,” (2022), arXiv:2206.07750 [quant-ph].
[19] Anthony Leverrier and Gilles emor, “Quantum tanner
codes,” (2022), arXiv:2202.13641 [quant-ph].
[20] Daniel Gottesman, “Fault-tolerant quantum computa-
tion with constant overhead,” (2014), arXiv:1310.2984
[quant-ph].
[21] Omar Fawzi, Antoine Grospellier, and Anthony Lever-
rier, “Constant overhead quantum fault-tolerance with
quantum expander codes,” in 2018 IEEE 59th An-
nual Symposium on Foundations of Computer Science
(FOCS) (IEEE, 2018).
[22] Nou´edyn Baspin and Anirudh Krishna, “Quantifying
nonlocality: How outperforming local quantum codes is
expensive,” Phys. Rev. Lett. 129, 050505 (2022).
[23] Nou´edyn Baspin, Omar Fawzi, and Ala Shayeghi, “A
lower bound on the overhead of quantum error correction
in low dimensions,” (2023), arXiv:2302.04317 [quant-ph].
[24] Sergey Bravyi, Andrew W. Cross, Jay M. Gambetta,
Dmitri Maslov, Patrick Rall, and Theodore J. Yo-
der, “High-threshold and low-overhead fault-tolerant
quantum memory,” Nature 627, 778–782 (2024),
arXiv:2308.07915 [quant-ph].
[25] Qian Xu, J. Pablo Bonilla Ataides, Christopher A. Pat-
tison, Nithin Raveendran, Dolev Bluvstein, Jonathan
Wurtz, Bane Vasi´c, Mikhail D. Lukin, Liang Jiang, and
Hengyun Zhou, “Constant-overhead fault-tolerant quan-
tum computation with reconfigurable atom arrays,” Na-
ture Physics (2024), 10.1038/s41567-024-02479-z.
[26] Yifan Hong, Matteo Marinelli, Adam M. Kaufman, and
Andrew Lucas, “Long-range-enhanced surface codes,”
(2024), arXiv:2309.11719 [quant-ph].
[27] Joshua Viszlai, Willers Yang, Sophia Fuhui Lin, Junyu
Liu, Natalia Nottingham, Jonathan M. Baker, and
Frederic T. Chong, “Matching Generalized-Bicycle Codes
to Neutral Atoms for Low-Overhead Fault-Tolerance,”
(2023), arXiv:2311.16980 [quant-ph].
[28] C. Ryan-Anderson, N. C. Brown, M. S. Allman, B. Arkin,
G. Asa-Attuah, C. Baldwin, J. Berg, J. G. Bohnet,
S. Braxton, N. Burdick, J. P. Campora, A. Chernoguzov,
J. Esposito, B. Evans, D. Francois, J. P. Gaebler, T. M.
Gatterman, J. Gerber, K. Gilmore, D. Gresh, A. Hall,
A. Hankin, J. Hostetter, D. Lucchetti, K. Mayer, J. My-
ers, B. Neyenhuis, J. Santiago, J. Sedlacek, T. Skripka,
A. Slattery, R. P. Stutz, J. Tait, R. Tobey, G. Vittorini,
J. Walker, and D. Hayes, “Implementing fault-tolerant
entangling gates on the five-qubit code and the color
code,” arXiv:2208.01863 (2022).
[29] S. A. Moses, C. H. Baldwin, M. S. Allman, R. An-
cona, L. Ascarrunz, C. Barnes, J. Bartolotta, B. Bjork,
P. Blanchard, M. Bohn, J. G. Bohnet, N. C. Brown,
N. Q. Burdick, W. C. Burton, S. L. Campbell, J. P.
Campora, C. Carron, J. Chambers, J. W. Chan, Y. H.
Chen, A. Chernoguzov, E. Chertkov, J. Colina, J. P.
Curtis, R. Daniel, M. DeCross, D. Deen, C. Delaney,
J. M. Dreiling, C. T. Ertsgaard, J. Esposito, B. Estey,
M. Fabrikant, C. Figgatt, C. Foltz, M. Foss-Feig, D. Fran-
cois, J. P. Gaebler, T. M. Gatterman, C. N. Gilbreth,
J. Giles, E. Glynn, A. Hall, A. M. Hankin, A. Hansen,
D. Hayes, B. Higashi, I. M. Hoffman, B. Horning, J. J.
Hout, R. Jacobs, J. Johansen, L. Jones, J. Karcz,
T. Klein, P. Lauria, P. Lee, D. Liefer, S. T. Lu, D. Luc-
chetti, C. Lytle, A. Malm, M. Matheny, B. Mathew-
son, K. Mayer, D. B. Miller, M. Mills, B. Neyenhuis,
L. Nugent, S. Olson, J. Parks, G. N. Price, Z. Price,
M. Pugh, A. Ransford, A. P. Reed, C. Roman, M. Rowe,
C. Ryan-Anderson, S. Sanders, J. Sedlacek, P. Shevchuk,
P. Siegfried, T. Skripka, B. Spaun, R. T. Sprenkle, R. P.
Stutz, M. Swallows, R. I. Tobey, A. Tran, T. Tran,
E. Vogt, C. Volin, J. Walker, A. M. Zolot, and J. M.
Pino, “A race-track trapped-ion quantum processor,”
Phys. Rev. X 13, 041052 (2023).
[30] D. Kielpinski, C. Monroe, and D. J. Wineland, “Ar-
chitecture for a large-scale ion-trap quantum computer,”
Nature 417, 709–711 (2002).
[31] Jonathan P. Home, David Hanneke, John D.
Jost, Jason M. Amini, Dietrich Leibfried, and
David J. Wineland, “Complete methods set
for scalable ion trap quantum information
processing,” Science 325, 1227–1230 (2009),
https://www.science.org/doi/pdf/10.1126/science.1177077.
[32] J. M. Pino, J. M. Dreiling, C. Figgatt, J. P. Gaebler,
S. A. Moses, M. S. Allman, C. H. Baldwin, M. Foss-Feig,
D. Hayes, K. Mayer, C. Ryan-Anderson, and B. Neyen-
huis, “Demonstration of the trapped-ion quantum-ccd
computer architecture,” Nature (2020), 10.1038/s41586-
021-03318-4,arXiv:2003.01293 [quant-ph].
[33] V. Kaushal, B. Lekitsch, A. Stahl, J. Hilder, D. Pijn,
C. Schmiegelow, A. Bermudez, M. uller, F. Schmidt-
Kaler, and U. Poschinger, “Shuttling-based trapped-ion
quantum information processing,” AVS Quantum Science
2, 014101 (2020),https://pubs.aip.org/avs/aqs/article-
pdf/doi/10.1116/1.5126186/19738817/014101 1 online.pdf.
[34] M Saffman, “Quantum computing with atomic qubits
6
and rydberg interactions: progress and challenges,” Jour-
nal of Physics B: Atomic, Molecular and Optical Physics
49, 202001 (2016).
[35] Antoine Browaeys and Thierry Lahaye, “Many-body
physics with individually controlled Rydberg atoms,” Na-
ture Physics 16, 132–142 (2020).
[36] Adam M Kaufman and Kang-Kuen Ni, “Quantum sci-
ence with optical tweezer arrays of ultracold atoms and
molecules,” Nature Physics 17, 1324–1333 (2021).
[37] Simon J. Evered, Dolev Bluvstein, Marcin Kalinowski,
Sepehr Ebadi, Tom Manovitz, Hengyun Zhou, Sophie H.
Li, Alexandra A. Geim, Tout T. Wang, Nishad Maskara,
Harry Levine, Giulia Semeghini, Markus Greiner, Vladan
Vuleti´c, and Mikhail D. Lukin, “High-fidelity parallel
entangling gates on a neutral-atom quantum computer,”
Nature 622, 268–272 (2023).
[38] Daniel Gottesman, “Stabilizer codes and quantum error
correction,” (1997), arXiv:quant-ph/9705052 [quant-ph].
[39] A. R. Calderbank and Peter W. Shor, “Good quantum
error-correcting codes exist,” Phys. Rev. A 54, 1098–1105
(1996).
[40] Andrew Steane, “Multiple-particle interference and
quantum error correction,” Proceedings of the Royal So-
ciety of London. Series A: Mathematical, Physical and
Engineering Sciences 452, 2551–2577 (1996).
[41] Argyris Giannisis Manes and Jahan Claes, “Distance-
preserving stabilizer measurements in hypergraph prod-
uct codes,” (2023), arXiv:2308.15520 [quant-ph].
[42] Yu Tomita and Krysta M. Svore, “Low-distance surface
codes under realistic quantum noise,” Phys. Rev. A 90,
062320 (2014).
[43] Joschka Roffe, David R. White, Simon Burton, and Earl
Campbell, “Decoding across the quantum low-density
parity-check code landscape,” Phys. Rev. Res. 2, 043423
(2020).
[44] Joschka Roffe, “LDPC: Python tools for low density par-
ity check codes,” (2022).
[45] Lawrence Z. Cohen, Isaac H. Kim, Stephen D. Bartlett,
and Benjamin J. Brown, “Low-overhead fault-tolerant
quantum computing using long-range connectivity,” Sci-
ence Advances 8(2022), 10.1126/sciadv.abn1717.
[46] Anders Sørensen and Klaus Mølmer, “Quantum compu-
tation with ions in thermal motion,” Phys. Rev. Lett. 82,
1971–1974 (1999).
[47] Richard Jozsa, “Fidelity for Mixed Quantum States,” J.
Mod. Opt. 41, 2315–2323 (1994).
[48] Anirudh Krishna and David Poulin, “Fault-tolerant gates
on hypergraph product codes,” Phys. Rev. X 11, 011023
(2021).
[49] Armanda O. Quintavalle, Paul Webster, and Michael
Vasmer, “Partitioning qubits in hypergraph product
codes to implement logical gates,” Quantum 7, 1153
(2023).
[50] Hayata Yamasaki and Masato Koashi, “Time-efficient
constant-space-overhead fault-tolerant quantum compu-
tation,” Nature Physics 20, 247–253 (2024).
[51] Satoshi Yoshida, Shiro Tamiya, and Hayata Ya-
masaki, “Concatenate codes, save qubits,” (2024),
arXiv:2402.09606 [quant-ph].
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Quantum low-density parity-check (qLDPC) codes can achieve high encoding rates and good code distance scaling, potentially enabling low-overhead fault-tolerant quantum computing. However, implementing qLDPC codes involves nonlocal operations that require long-range connectivity between qubits. This makes their physical realization challenging in comparison to geometrically local codes, such as the surface code. Here we propose a hardware-efficient scheme for fault-tolerant quantum computation with high-rate qLDPC codes that is compatible with the recently demonstrated capabilities of reconfigurable atom arrays. Our approach utilizes the product structure inherent in many qLDPC codes to implement the nonlocal syndrome extraction circuit through atom rearrangement, resulting in an effectively constant overhead. We prove the fault tolerance of these protocols, and our simulations show that the qLDPC-based architecture starts to outperform the surface code with as few as several hundred physical qubits. We further find that quantum algorithms involving thousands of logical qubits can be performed using less than 10⁵ physical qubits. Our work suggests that low-overhead quantum computing with qLDPC codes is within reach using current experimental technologies.
Article
Full-text available
The accumulation of physical errors1–3 prevents the execution of large-scale algorithms in current quantum computers. Quantum error correction⁴ promises a solution by encoding k logical qubits onto a larger number n of physical qubits, such that the physical errors are suppressed enough to allow running a desired computation with tolerable fidelity. Quantum error correction becomes practically realizable once the physical error rate is below a threshold value that depends on the choice of quantum code, syndrome measurement circuit and decoding algorithm⁵. We present an end-to-end quantum error correction protocol that implements fault-tolerant memory on the basis of a family of low-density parity-check codes⁶. Our approach achieves an error threshold of 0.7% for the standard circuit-based noise model, on par with the surface code7–10 that for 20 years was the leading code in terms of error threshold. The syndrome measurement cycle for a length-n code in our family requires n ancillary qubits and a depth-8 circuit with CNOT gates, qubit initializations and measurements. The required qubit connectivity is a degree-6 graph composed of two edge-disjoint planar subgraphs. In particular, we show that 12 logical qubits can be preserved for nearly 1 million syndrome cycles using 288 physical qubits in total, assuming the physical error rate of 0.1%, whereas the surface code would require nearly 3,000 physical qubits to achieve said performance. Our findings bring demonstrations of a low-overhead fault-tolerant quantum memory within the reach of near-term quantum processors.
Article
Full-text available
Scaling up quantum computers to attain substantial speedups over classical computing requires fault tolerance. Conventionally, protocols for fault-tolerant quantum computation demand excessive space overheads by using many physical qubits for each logical qubit. A more recent protocol using quantum analogues of low-density parity-check codes needs only a constant space overhead that does not grow with the number of logical qubits. However, the overhead in the processing time required to implement this protocol grows polynomially with the number of computational steps. To address these problems, here we introduce an alternative approach to constant-space-overhead fault-tolerant quantum computing using a concatenation of multiple small-size quantum codes rather than a single large-size quantum low-density parity-check code. We develop techniques for concatenating different quantum Hamming codes with growing size. As a result, we construct a low-overhead protocol to achieve constant space overhead and only quasi-polylogarithmic time overhead simultaneously. Our protocol is fault tolerant even if a decoder has a non-constant runtime, unlike the existing constant-space-overhead protocol. This code concatenation approach will make possible a large class of quantum speedups with feasibly bounded space overhead yet negligibly short time overhead.
Article
Full-text available
The promise of high-rate low-density parity check (LDPC) codes to substantially reduce the overhead of fault-tolerant quantum computation depends on constructing efficient, fault-tolerant implementations of logical gates on such codes. Transversal gates are the simplest type of fault-tolerant gate, but the potential of transversal gates on LDPC codes has hitherto been largely neglected. We investigate the transversal gates that can be implemented in hypergraph product codes, a class of LDPC codes. Our analysis is aided by the construction of a symplectic canonical basis for the logical operators of hypergraph product codes, a result that may be of independent interest. We show that in these codes transversal gates can implement Hadamard (up to logical SWAP gates) and control-Z on all logical qubits. Moreover, we show that sequences of transversal operations, interleaved with error correction, allow implementation of entangling gates between arbitrary pairs of logical qubits in the same code block. We thereby demonstrate that transversal gates can be used as the basis for universal quantum computing on LDPC codes, when supplemented with state injection.
Article
Full-text available
The ability to perform entangling quantum operations with low error rates in a scalable fashion is a central element of useful quantum information processing¹. Neutral-atom arrays have recently emerged as a promising quantum computing platform, featuring coherent control over hundreds of qubits2,3 and any-to-any gate connectivity in a flexible, dynamically reconfigurable architecture⁴. The main outstanding challenge has been to reduce errors in entangling operations mediated through Rydberg interactions⁵. Here we report the realization of two-qubit entangling gates with 99.5% fidelity on up to 60 atoms in parallel, surpassing the surface-code threshold for error correction6,7. Our method uses fast, single-pulse gates based on optimal control⁸, atomic dark states to reduce scattering⁹ and improvements to Rydberg excitation and atom cooling. We benchmark fidelity using several methods based on repeated gate applications10,11, characterize the physical error sources and outline future improvements. Finally, we generalize our method to design entangling gates involving a higher number of qubits, which we demonstrate by realizing low-error three-qubit gates12,13. By enabling high-fidelity operation in a scalable, highly connected system, these advances lay the groundwork for large-scale implementation of quantum algorithms¹⁴, error-corrected circuits⁷ and digital simulations¹⁵.
Article
Full-text available
Practical quantum computing will require error rates well below those achievable with physical qubits. Quantum error correction1,2 offers a path to algorithmically relevant error rates by encoding logical qubits within many physical qubits, for which increasing the number of physical qubits enhances protection against physical errors. However, introducing more qubits also increases the number of error sources, so the density of errors must be sufficiently low for logical performance to improve with increasing code size. Here we report the measurement of logical qubit performance scaling across several code sizes, and demonstrate that our system of superconducting qubits has sufficient performance to overcome the additional errors from increasing qubit number. We find that our distance-5 surface code logical qubit modestly outperforms an ensemble of distance-3 logical qubits on average, in terms of both logical error probability over 25 cycles and logical error per cycle ((2.914 ± 0.016)% compared to (3.028 ± 0.023)%). To investigate damaging, low-probability error sources, we run a distance-25 repetition code and observe a 1.7 × 10−6 logical error per cycle floor set by a single high-energy event (1.6 × 10−7 excluding this event). We accurately model our experiment, extracting error budgets that highlight the biggest challenges for future systems. These results mark an experimental demonstration in which quantum error correction begins to improve performance with increasing qubit number, illuminating the path to reaching the logical error rates required for computation. A study demonstrating increasing error suppression with larger surface code logical qubits, implemented on a superconducting quantum processor.
Article
Full-text available
Quantum computers can be protected from noise by encoding the logical quantum information redundantly into multiple qubits using error-correcting codes1,2. When manipulating the logical quantum states, it is imperative that errors caused by imperfect operations do not spread uncontrollably through the quantum register. This requires that all operations on the quantum register obey a fault-tolerant circuit design3–5, which, in general, increases the complexity of the implementation. Here we demonstrate a fault-tolerant universal set of gates on two logical qubits in a trapped-ion quantum computer. In particular, we make use of the recently introduced paradigm of flag fault tolerance, where the absence or presence of dangerous errors is heralded by the use of auxiliary flag qubits6–10. We perform a logical two-qubit controlled-NOT gate between two instances of the seven-qubit colour code11,12, and fault-tolerantly prepare a logical magic state8,13. We then realize a fault-tolerant logical T gate by injecting the magic state by teleportation from one logical qubit onto the other14. We observe the hallmark feature of fault tolerance—a superior performance compared with a non-fault-tolerant implementation. In combination with recently demonstrated repeated quantum error-correction cycles15,16, these results provide a route towards error-corrected universal quantum computation. A fault-tolerant, universal set of single- and two-qubit quantum gates is demonstrated between two instances of the seven-qubit colour code in a trapped-ion quantum computer.
Article
We construct a new explicit family of good quantum low-density parity-check codes which additionally have linear time decoders. Our codes are based on a three-term chain (F₂(m×m))ⱽ −→^(δ0) (F₂ᵐ)ᴱ −→^(δ¹) F₂^F where V (X-checks) are the vertices, E (qubits) are the edges, and F (Z-checks) are the squares of a left-right Cayley complex, and where the maps are defined based on a pair of constant-size random codes C_A,C_B : F₂ᵐ → F₂^Δ where Δ is the regularity of the underlying Cayley graphs. One of the main ingredients in the analysis is a proof of an essentially-optimal robustness property for the tensor product of two random codes.
Article
Quantum low-density parity-check (LDPC) codes are a promising avenue to reduce the cost of constructing scalable quantum circuits. However, it is unclear how to implement these codes in practice. Seminal results of Bravyi et al. [Phys. Rev. Lett. 104, 050503 (2010)PRLTAO0031-900710.1103/PhysRevLett.104.050503] have shown that quantum LDPC codes implemented through local interactions obey restrictions on their dimension k and distance d. Here we address the complementary question of how many long-range interactions are required to implement a quantum LDPC code with parameters k and d. In particular, in 2D we show that a quantum LDPC code with distance d∝n^{1/2+ϵ} requires Ω(n^{1/2+ϵ}) interactions of length Ω[over ˜](n^{ϵ}). Further, a code satisfying k∝n with distance d∝n^{α} requires Ω[over ˜](n) interactions of length Ω[over ˜](n^{α/2}). As an application of these results, we consider a model called a stacked architecture, which has previously been considered as a potential way to implement quantum LDPC codes. In this model, although most interactions are local, a few of them are allowed to be very long. We prove that limited long-range connectivity implies quantitative bounds on the distance and code dimension.