<<

Practical with the XZZX code and Kerr-cat

Andrew S. Darmawan,1, 2 Benjamin J. Brown,3 Arne L. Grimsmo,3 David K. Tuckett,3 and Shruti Puri4, 5 1Yukawa Institute for Theoretical Physics (YITP), Kyoto University, Kitashirakawa Oiwakecho, Sakyo-ku, Kyoto 606-8502, Japan∗ 2JST, PRESTO, 4-1-8 Honcho, Kawaguchi, Saitama 332-0012, Japan 3Centre for Engineered Quantum Systems, School of Physics, University of Sydney, Sydney, New South Wales 2006, Australia 4 Department of Applied Physics, Yale University, New Haven, Connecticut 06511, USA† 5Yale Quantum Institute, Yale University, New Haven, Connecticut 06511, USA (Dated: April 21, 2021) The development of robust architectures capable of large-scale fault-tolerant quantum computa- tion should consider both their quantum error-correcting codes, and the underlying physical qubits upon which they are built, in tandem. Following this design principle we demonstrate remarkable error correction performance by concatenating the XZZX surface code with Kerr-cat qubits. We contrast several variants of fault-tolerant systems undergoing different circuit noise models that reflect the physics of Kerr-cat qubits. Our simulations show that our system is scalable below a threshold gate infidelity of pCX 6.5% within a physically reasonable parameter regime, where ∼ pCX is the infidelity of the noisiest gate of our system; the controlled-not gate. This threshold can be reached in a superconducting circuit architecture with a Kerr-nonlinearity of 10MHz, a 6.25 photon cat , single-photon lifetime of > 64µs, and thermal photon population < 8%.∼ Such parameters are routinely achieved in superconducting∼ circuits. ∼

I. INTRODUCTION qubit [22, 34, 35]. The Kerr-cat qubit is realized by para- metric pumping of a Kerr-nonlinear oscillator, as recently A scalable quantum computer will require a large num- demonstrated in a superconducting circuit platform using ber of almost perfect qubits. To this end, impressive a Superconducting Nonlinear Asymmetric Inductive eL- progress has been made in engineering high-quality quan- ement (SNAIL) [11]. This qubit is stabilized against bit tum systems in the laboratory that can perform some flips. Specifically, bit flips are exponentially suppressed with the photon number α 2, leading to a highly bi- requisite set of operations that are needed for quantum ∼ | | computation [1–13]. While we cannot expect to reduce ased noise channel for large α [35]. Surprisingly, it has qubit noise arbitrarily with laboratory engineering alone, recently been proposed that the controlled-not gate (CX) the theory of quantum error correction [14–19] has shown can be executed with Kerr-cat qubits without changing that we can produce logical qubits that experience next- this noise profile [35]. This enables us to go beyond to-no noise by using a redundancy of high-quality phys- schemes where it is assumed that only diagonal entan- ical qubits, provided the noise strength on the physical gling gates can be used without introducing bit-flip errors qubits as they perform computational operations is below at a significant rate [36, 37]. some threshold value [20]. Nevertheless, it still remains A number of error-correction protocols have been pro- a significant experimental challenge to produce scalable posed recently that are adapted specifically to biased quantum error-correcting codes using available technol- noise [30, 36–47]. In particular, the XZZX surface ogy. In order to overcome this challenge we should co- code [48, 49], a surface code variant [15, 16, 50], has design the code in unison with the underlying qubit ar- recently been shown to have a very high threshold un- chitecture. This way we can build better error-correcting der biased noise. This is because it mimics the behav- systems that specifically target the dominant sources of ior of the classical repetition code in the limit that a errors that the physical qubits experience. particular type of Pauli error is dominant. As such, arXiv:2104.09539v1 [quant-ph] 19 Apr 2021 A promising approach to realize a high-quality phys- minimum-weight perfect-matching decoders [16, 51, 52] ical qubit is to encode it in a bosonic mode [7–13, 21– have been developed by exploiting the symmetries of this 33]. Broadly speaking, encoding a qubit in the larger code [15, 16, 43, 53]. These practical decoders demon- Hilbert space of an oscillator gives more room to pro- strate exceptionally high fault-tolerant thresholds when tect the encoded information from noise, either through the dominant source of noise is dephasing [49]. active error correction or autonomous stabilization of In this work we present a scheme for error correction in the codespace. One of the simplest incarnations of this which the XZZX surface code is concatenated with Kerr- idea is the Kerr-cat qubit, where orthogonal superpo- cat qubits. Figure1 shows what a Kerr-cat XZZX chip sitions of two coherent states α define a physical |± i may look like if implemented with SNAILs. The domi- nant physical error channels for Kerr-cat qubits are pho- ton loss and gain, quantified by a single-photon loss rate κ ∗ [email protected] and the oscillator’s thermal population nth [11]. The [email protected] † noise model of the Kerr-cat qubit is a function of κ, nth, 2 as well as the “cat size” α 2 and the Kerr-nonlinearity of the oscillator K. With |α|2 = 6.25 and a thermal pop- | | 4 ulation of 8%, we find a threshold of κ/K = 2.5 10− . This corresponds to a total infidelity of 6.5%× for the noisiest logical operation in the error correction∼ circuit, CX, whose average bias is 351 where we define bias as the ratio of the probability∼ of Pauli Z errors and the probability of any other Pauli error occurring dur- ing the gate. This noise threshold can be reached, for example, using a superconducting Kerr-cat realization with K/(2π) = 10 MHz, a parametric pump strength of 1 62.5 MHz, and a single photon lifetime κ− = 63.6 µs. We remark that the experimental demonstration of the Kerr- cat qubit in Ref. 11 showed that frequency fluctuations are heavily suppressed when the SNAIL is pumped into the cat-subspace, such that requirements on flux noise = DC flux-bias for the SNAIL (characterized by the T2 time of the un-pumped SNAIL) are greatly relaxed compared to frequency tunable trans- = resonator for qubit readout mons. These parameters are routinely achieved in su- = out-of-plane interconnect = perconducting circuits [54, 55], suggesting that Kerr-cats through which all the microwave well below threshold may be realized in the near future, drives (such as the two-photon making the proposed architecture a promising approach SNAIL drives, coupling drives for CX gates and drives for single-qubit gates) to large-scale fault-tolerant . are applied While our protocol achieves exceptional performance using Kerr-cat qubits, we also show that similar improve- FIG. 1. Schematic for a possible realization of the XZZX ments may be attainable in other biased-noise architec- surface code using a 2D lattice of nearest-neighbor capaci- tures. Using the XZZX code with a generic biased circuit tively coupled SNAIL-based Kerr-cat qubits. Data qubits are noise model we observe roughly a factor of two improve- represented in magenta and ancilla qubits in green. A DC ment in the threshold error rate over the standard CSS flux line is present to bias the SNAIL at the point of desired version of the surface code. Surprisingly, we still find nonlinearities. The lattice arrangement is such that neigh- that the XZZX code has a 50% higher threshold error boring SNAILs resonate at different frequencies to minimize rate than the CSS surface code when using standard CX cross-talk. Microwave drives applied at appropriate frequen- gates rather than bias-preserving CX gates. cies through the out-of-plane interconnects parametrically ac- tivate gate operations [11, 35]. Importantly, no extra nonlin- We also perform exact simulations on small codes that ear control/coupling elements are required unlike other biased represent near-term experiments. These simulations al- noise bosonic qubits. For direct readout of ancilla and data low us to study non-Pauli noise models with optimal qubits, an additional readout resonator is available at each decoders and reveal that the high performance of our site. This readout resonator also provides the ability to para- scheme at both high and low biases is relatively insensi- metrically introduce two-photon dissipation in the SNAILs which could be used for autonomous leakage correction. tive to sub-optimal implementation of gates and decoding compared to other schemes.

This paper is structured as follows. In Sec.II we re- II. THE XZZX SURFACE CODE view the definition of the XZZX surface code and describe how it can be implemented in a Kerr-cat architecture. In Sec.III we review Kerr-cat qubits and describe the We describe the XZZX code using the stabilizer for- bias-preserving gates available to them. In Sec.IIIB we malism, i.e., a list of commuting Pauli operators whose describe the important noise processes affecting Kerr-cat common +1 eigenspace defines the codespace of the code. qubits and the methods we have used to simulate noisy We lay out the code with a single qubit on each vertex of gate execution. In Sec.IV we present results of numerical the square lattice. Up to its boundaries, all of its stabi- simulations, demonstrating that very high thresholds can lizers are the product of four Pauli terms; XZZX acting be achieved using XZZX codes and Kerr-cat qubits. The on the four qubits at the corners of each of the faces of performance of low-distance codes under biased noise is the square lattice, see Fig.2(a). We denote the standard presented in Sec.V. A discussion of the results along Pauli matrices by Xj, Yj and Zj acting on the qubit in- dexed j with 1 j n, and we write the stabilizer with potential future research directions are provided in ≤ ≤ Sec.VI. Details of the surface code simulation methods generator associated to face f as Sf . and some advantages of Kerr-cat qubits over purely dis- The XZZX code is locally equivalent [56, 57] to the con- sipative cat qubits are included in AppendicesA andB ventional CSS surface code [15, 16]. As such it inherits respectively. its code properties. We choose a lattice of size d d ∼ x × z 3

small n = 9 codes in Sec.V where we use the rotated A0 A1 A2 A3 A4 surface code layout. Here, on every second face the or- D1 der of application of the two CX gates is swapped. For

D2 that particular layout, this order was shown in Ref. 59 to mitigate the effect of hook errors, in which a single error D3 on an ancilla can propagate to two data qubits. D4 In a realistic device, errors may occur during gate ex- ecution, state preparation, measurements and on idle qubits. Repeated measurement of stabilizer generators provides data, called the syndrome, that allows us to FIG. 2. (a) An XZZX surface code check as a product of two X and two Z Pauli operators on qubits arranged on the correct for all of these types of errors, provided we adopt vertices of a square lattice. Every face corresponds to a mea- an appropriate decoding strategy. sured check operator and all checks (except for the three-qubit checks on the boundary) have the same form. The order in which data qubits of a check are coupled to the ancilla (at B. Decoding the center of each face) is specified by the arrows. We have used the same order for every check. (b) Check measurement circuit for the XZZX surface code. The four data qubits are A decoder is a classical algorithm that uses the error coupled to an ancilla qubit prepared in the + state before syndrome as an input to estimate the best choice of cor- measuring the ancilla in the X basis. Distinct| errori locations rection that can be applied to recover the encoded state. on the ancilla qubits are labelled A0-4, and on data qubits We use a minimum-weight perfect-matching (MWPM) are labelled D1-4. The error probabilities at these locations decoding algorithm to correct for all types of errors [16]. determine the weights of the syndrome graph for decoding. Note that correctly identifying the error, up to stabilizer equivalence, is equivalent to finding a correction for that error. We format the error syndrome as a list of defects, using n = O(dxdz) qubits, where dx and dz correspond where we say that a defect appears at face f and at time to the lengths of the smallest X and Z logical operators, t if S (t 1)S (t) = 1, where S (t) denotes the out- f − f − f which have a string-like support and traverse homolog- come of the stabilizer measurement Sf made at time t. ically non-trivial paths between opposite boundaries of Describing the error syndrome as a list of defects means the code. With open boundary conditions, the XZZX that errors can be interpreted as strings where defects lie code encodes k = 1 qubit. at their endpoints [16]. This interpretation enables us to employ MWPM to estimate the error that has occurred. The MWPM algorithm takes a graph with weighted A. Stabilizer measurements edges and returns a matching such that the sum of the weights of the matching is minimal. Given the error We measure stabilizer operators to identify the loca- model and the syndrome, the graph input to MWPM tions of errors that occur on the qubits of the code. As is constructed such that each node corresponds to a de- the codespace is defined as the +1 eigenspace of the sta- fect, and the weight of an edge connecting two defects is bilizers, measuring a stabilizer in the 1 state indicates log p, where p is the probability of the most likely error − an error has occurred that takes the− system out of its that can give rise to that pair of defects. This choice of codespace. weights implies that the edges returned by the MWPM We can read out all of the stabilizer generators simul- algorithm correspond globally to the most likely error taneously using the circuit shown in Fig.2(b). We place consistent with the observed syndrome. a single ancilla in the + state on each face. We subse- This practical decoder is particularly well suited for de- quently apply two-qubit| entanglingi gates between the an- coding biased noise using the XZZX code. For the XZZX cilla and its four nearest-neighbor data qubits such that code, a Z error on any data qubit will produce a pair of the ancilla collects the parity information for the stabi- defects displaced in the same direction, independent of lizer Sf on their respective face f. We finally measure the qubit, and this is perpendicular to the displacement the ancilla in the Pauli-X basis to learn the value of the produced by an X error. Thus, when the noise is strongly check. biased towards Z errors, the edge weights in the syn- In Fig.2(a), we have indicated the order in which drome graph have much higher weight in one direction. qubits should be coupled to the ancilla. With the ex- This allows errors to be identified less ambiguously us- ception of qubits on the boundary, every data qubit is ing MWPM, compared to when edges in both directions coupled with an ancilla qubit at each time step. The are more equally weighted, and results in a high thresh- same ordering is used for checks on the boundary, except old under biased noise. This was demonstrated using a when there is no neighbouring ancilla or data qubit to phenomenological noise model in Ref. 49. couple with (at which time the qubit is left idle). This In realistic circuit level noise, single error events can ordering enables parallel readout of all the stabilizer op- give rise to pairs of defects separated ‘diagonally’ i.e. in erators [58]. The exception to this ordering is for the both space and time as well as hook errors, which oc- 4 cur when the ancilla qubit experiences a X error halfway (a) through the stabilizer readout circuit [an X error at A2 in Fig.2(b)]. This error is copied onto two of the data qubits by the entangling gates. However, as the hook error is caused by a low rate event, the effect of these errors is relatively benign for biased noise. Our decoder takes into account all of these types of errors by appro- priately assigning edge weights in the construction of the syndrome graph.

III. CAT QUBITS AND BIAS PRESERVING GATES

(b) Here we describe the Kerr-cat [11, 34] which, in our proposal, forms the elementary qubit from which the XZZX code is built. The Kerr-cat qubit is realized in a two-photon driven Kerr nonlinear oscillator. The Hamil- tonian of such an oscillator in a frame rotating at the oscillator frequency ωt is given by

2 2 2 2 2 H = Ka† a + Kα (a† + a ). (1) cat − Here, K > 0 is the strength of the nonlinearity while Kα2 is the amplitude of the two-photon drive. The orthogonal even- and odd-parity cat states α± = ( α α ) are degenerate eigenstates of this|C Hamiltonian.i N± | i ± As|− il-i lustrated in Fig.3, the degenerate cat-subspace = C span α± is separated from the rest of the Hilbert space by{|C ani} energy gap, which in the limit of large α is well C⊥ 2 FIG. 3. (a) Eigenspectrum of the two-photon driven nonlin- approximated by ωgap 4Kα . We will work in the | | ∼ + ear oscillator in the rotating frame. The cat states α± are basis where the cat states α + and α− |C i are aligned along the X-axis|C ofi ≡ the | Blochi sphere,|C i ≡ and |−i exactly degenerate [34]. The eigenspectrum can be divided + √ into an even and an odd parity manifold. The cat subspace, their superpositions ( α + α− )/ 2 0 α and highlighted in green, is separated from the first excited state + |C i |C i ≡ | i' | i ( − )/√2 1 α are aligned along the Z- 2 |Cα i − |Cα i ≡ | i' |− i by an energy gap ωgap 4Kα . (b) Bloch sphere of the cat axis (where the approximations hold for large α ). qubit. The figure also| shows| ∼ cartoons of the Wigner functions | | The Kerr-cat qubit has been recently realized exper- corresponding to the eigenstates of Pauli-X,Y ,Z operators. imentally with a capacitively shunted SNAIL [11]. The SNAIL is a flux-biased nonlinear device comprising of four Josephson junctions (Fig.1). Both third- and the setup in Fig.4 which is the building block of the sur- fourth-order (Kerr) nonlinearities exist in the SNAIL at face code architecture. Suppose the resonance frequency an appropriate value of the external magnetic flux. Due of one of the data qubit SNAILs is ωt. Each SNAIL is to the third-order non-linearity, a single microwave drive capacitively coupled to a readout resonator of frequency applied at twice the resonance frequency generates the ωr. A microwave drive applied to the SNAIL at 2ωt ωr two-photon term in the Hamiltonian in Eq. (1). can induce the two-photon dissipation as follows.− Due It has been shown that the error channel of the Kerr- to the Kerr-nonlinearity in the SNAIL, the drive pho- cat qubit is dominated by phase-flip errors while X and Y tons are consumed to convert two photons in the SNAIL errors are exponentially suppressed with the average pho- to one photon in the readout mode. The photon in the 2 ton number a†a α [35]. Some noise sources, such low-Q readout is rapidly lost to the environment, thus ef- as thermal excitation,h i ∼ | frequency-fluctuations,| etc., may fectively inducing two-photon dissipation in the SNAIL. introduce leakage errors. Leakage can, to some extent, be Depending on the drive amplitude, mixing between the autonomously corrected or “cooled” by introducing any readout and SNAIL mode, and lifetime of the readout, additional photon dissipation channel without adversely the strength of the two-photon dissipation achieved can affecting the bias [35]. A two photon dissipation source be easily engineered to be K/100 K/10. − is preferred for such autonomous leakage correction. Single-photon loss and thermal noise were observed The inherent nonlinearity in the Kerr-cat enables a to be the two main sources of errors in the experiment straight-forward realization of two-photon dissipation for in Ref. 11. Consequently, we also consider these noise autonomous leakage correction. For example, consider sources for the simulations in the present work. Impor- 5 tantly, the experiment found that non-phase-flip errors decreased with increasing cat size, confirming the theo- retically predicted biased-noise property of the qubit.

A. Kerr-cats and dissipative cats

In our work, the choice of the Kerr-cat as the biased- noise system used for the numerical analysis is moti- vated by several characteristics that makes it an excellent candidate for large-scale quantum computation. Firstly, FIG. 4. A drive at 2ωt applied to the SNAIL implements the the large nonlinearities inherently present in supercon- two-photon drive required to realize the Kerr-cat Hamiltonian ducting circuits ensures a large energy gap ωgap , typ- in Eq. (1). Here ωt is the resonance frequency of the SNAIL. ically in the range of 10 300 MHz. As| further| de- It is also possible to activate the two-photon dissipation for scribed in the next section,− this large gap allows fast autonomous leakage correction by applying a microwave drive at 2ωt ωr. Here ωr is the resonance frequency of the readout gates (10 100 ns). The typical lifetimes in these cir- − cuits, 10 −100µs [54, 55, 60, 61], are 100 1000 times resonators coupled to the SNAIL. longer than− the gate times ensuring high-fidelity∼ − gate op- erations [11, 35]. Consequently, it becomes much easier to reach the threshold parameters estimated in our work the Kerr-cat readout scheme, this scheme is slow and its with current experimental techniques. The inherent non- fidelity is effectively independent of cat-size. linearity can also be exploited to implement all the single- Another challenge with dissipative cats is that ad- and multi-qubit gate operations and measurements using ditional nonlinear coupling elements required to imple- microwave pulses, thus eliminating any need for external ment quantum operations may open up new sources of coupling elements and providing a hardware efficient so- dissipation in the system which can reduce the noise lution for scaling up [11, 35]. bias [62, 63]. Finally, during gate operations, dissipative We now contrast the Kerr-cat qubit with the purely cats may suffer from non-adiabatic phase-flip errors (see dissipative cat qubit which also has a strongly biased AppendixB)[30]. Because this extra source of phase- noise channel [24]. The dissipative cat is realized by en- flips is absent in Kerr-cats, it is easier to achieve the gineering a two-photon drive and two-photon dissipation threshold parameters in practice. in a linear resonator via external nonlinear modes, such Future theoretical and experimental research in opti- as a [12, 30, 62]. So far, the strength of the mizing the hardware and control architecture will cer- two-photon dissipation induced this way has been lim- tainly improve both biased noise cat platforms and may ited to 10 100 KHz, leading to slow gate operations even lead to development of new biased-noise qubits [46]. (> 1µs)[63−]. But because of the reasons outlined above and the cur- ∼In addition, the lack of self-Kerr nonlinearity in purely rent state of technology, this paper focuses on the XZZX dissipative cats makes implementation of gate operations, code implementation with Kerr-cat qubits. preparation, and measurements more challenging. This is evident, for example, from the scheme for X-basis mea- surement. The Kerr-cat readout (see Sec.IIIB for more B. Logical Cat Operations and Noise Channels details) proceeds by effectively rotating the qubit around the Y axis, which maps the X basis states to the Z basis. We assume that at the beginning of the computation, The Z basis states are subsequently mapped onto a lin- each SNAIL (data and ancilla) has been pumped into ear readout resonator and a homodyne measurement of the cat-qubit basis by the adiabatic pumping technique its field reveals the state of the Kerr-cat qubit. A useful described in Refs. [11, 34]. Once in the logical cat sub- feature of this scheme is that the strength of the ho- space, the set of bias-preserving operations in the Kerr- modyne field, and hence measurement fidelity, increases cat platform, relevant for the XZZX code implementa- with the size of the Kerr-cat. The Y axis rotation is tion, are the projective X-basis measurement, X , and M possible by freely evolving the qubit under the self-Kerr the gates S, S†, CZ, CX . We use the projective mea- { } nonlinearity in the Kerr-cat mode. This rotation is not surement X for state preparation + as well. We will M P| i possible in purely dissipative cats and hence this simple now describe how these operations can be implemented. measurement scheme cannot be employed. In the case of dissipative cats, X-basis measurement proceeds by adiabatic deflation of the even and odd par- 1. S, S† gates ity cat states to vacuum and single-photon Fock states respectively. The Fock states are subsequently read out Consider the Hamiltonian of a Kerr-cat qubit when, via coupling to other nonlinear modes like a transmon or in addition to the two-photon drive, an additional res- asymmetrically threaded SQUID [12, 30]. Compared to onant microwave drive of amplitude ε is applied HZ = 6

Hcat + (aε∗ + a†ε). Projecting the interaction onto the (a) 2α2 logical subspace gives P HZ P = α(ε+ε∗)Z+ ∗(e− ), C C + + O where = 0 0 + 1 1 = + − − and Z PC | i h | | i h | |Cα ihCα | |Cα ihCα | is Pauli-Z in the logical cat basis (we use ∗[f(x)] nota- O tion to suppress polynomial factors in x [64]). S and S† gates are implemented by applying the microwave drive, (b) in phase (ε = ε∗ > 0) and π out-of-phase (ε = ε∗ < 0) with the two-photon drive respectively for time TS = π/8αε, see Fig.5(a). Clearly, by increasing the cat size and strength of the microwave drive ε it is possible to | | decrease TS, thereby decreasing the probability of errors during the gate operation. Note that the drive can cause (c) virtual transitions out of the computational cat subspace. Consequently, we must have ε ωgap so that these virtual transitions and leakage| | outside  | the| qubit mani- fold ε/ω 2 remain negligible. ∼ | gap| (d)

2. CZ gate

The implementation of a CZ gate between two Kerr- cats requires a Hamiltonian interaction of the form

HZZ = Hcat,c + Hcat,t J(ac†at + acat†) + J(ac + ac†), see Fig.5(b) for the corresponding− drive scheme. Here FIG. 5. Schematic showing parametrically driven (a) S, S†, (b) CZ, (c) CX, and (d) CD operations required to realize the subscripts c, t refer to the control and target qubits syndrome measurement circuits for the surface codes. The respectively. Evolution under this Hamiltonian for time X(π/4) gate required for ancilla readout is not shown here. 2 TCZ = π/8Jα realizes the CZ gate. Clearly, larger α It simply requires turning off the two-photon drive on the and J lead to a faster gate. To minimize virtual transi- ancilla Kerr-cat for time π/2K. The speed of the homo- tions out of the computational cat subspace we must have dyne measurement after the CD operation is characterized 2 Jα ω [34]. The term (a a + a a†) is a beam- by the rate Rmeas = 2κr β , where β = 2gα/κr is the max- gap c† t c t | | | | splitter | interaction| between the∝ two qubits and is realized imum of readout-resonator displacement caused by the Kerr- by applying a drive to either of the two qubits at their cat qubit [65]. frequency difference. The third-order nonlinearity in the SNAIL consumes the drive photons to convert photons in one SNAIL to photons in the other, effective realizing using a total Hamiltonian the beamsplitter Hamiltonian. The term (at + at†) is 2 2 2 2 ∝ H = K(a† β )(a β ) realized by applying a resonant microwave drive to the CX − c − c −      target SNAIL. 2 2 i2φ(t) β ac† 2 β + ac† K a† α e− − α − t − 2β − 2β  β a  β + a  a2 α2ei2φ(t) − c α2 c 3. CX gate × t − 2β − 2β φ˙(t) The diagonal S, S†, and CZ gates are trivially bias at†at(2β ac† ac). preserving. A more challenging gate to realize, how- − 4β − − (2) ever, is the bias-preserving CX gate. Such a gate can The interactions needed for the CX gate can be acti- be achieved by exploiting the continuous variable na- vated parametrically as shown in Fig.5(c) (adapted from ture of the cat encoding, and temporarily leaving the Ref.35). The CX gate time is limited by the adiabaticity cat-code subspace during the gate. The intuition for the requirement φ˙(t) ω . CX gate is best understood by first considering a bias  | gap| preserving X (or NOT) gate. An X gate can be exe- cuted by adiabatically changing the phase of the two co- iφ(t) 4. X , + herent states α αe such that φ(0) = 0 and M P| i φ(T ) = π. At|± timeiT →we |± have thusi mapped α α and implemented an X gate (in the large α|±limiti → where |∓ i The remaining components required for the surface 0/1 α ). Similarly, a CX gate can be executed code implementation are the projective X-basis measure- | i ' |± i if we can condition the phase φ(T ) on the state of the ment, X , and preparation + . The projective mea- control qubit. As shown in Ref. 35 this can be achieved surementM proceeds in two steps.P| Thei first step comprises 7 of rotating the ancilla states to the computational basis states 0, 1|±irespectively, using an S gate followed | i by X(π/4) = (1 + iX)/√2 operation (see Fig.6). The X(π/4) gate is a non-bias preserving gate but since this gate is being applied to the ancilla during mea- surement the effect of a non-dephasing error is the same To Homodyne as a dephasing error [36]. It is implemented by turning off the parametric drive and evolving the cats freely under the self-Kerr nonlinearity for time π/2K [11, 66]. FIG. 6. Circuit for projective measurement and subsequent preparation of basis states in the X basis. The top line is Once the states are rotated to 0, 1 , a controlled- |±i | i the data qubit and the botton is the readout resonator. The displacement (CD) operation between the Kerr-cat qubit controlled-displacement or CD gate maps the state of the data and a linear readout resonator can be used to read out in qubit to the displacement of the readout mode. A homodyne the computational basis. The required (rotating frame) measurement of the signal from the readout resonator then reveals the state of the qubit. interaction Hamiltonian is Hreadout = Hcat + g(a1† + a1)(ar† + ar)/2, where a1, ar are the annihilation opera- tors for the Kerr-cat mode and the readout mode respec- tively and g is the beam splitter coupling between the IV. THRESHOLDS two modes [11, 66]. Projecting onto the qubit subspace gives P HreadoutP = gαZ1(ar† +ar). From this projected C C A crucial figure of merit for a quantum error-correcting Hamiltonian it is clear that the readout resonator expe- code is the threshold: the maximum noise strength be- riences a displacement conditioned on the state of the low which it becomes possible to efficiently reduce the qubit. noise strength on the logical level arbitrarily by increas- If the single-photon decay rate of the readout mode ing the number of physical qubits and gates. We have is κr, then the qubit-dependent displacement amplitude performed numerical simulations to estimate the thresh- κrt/2 is (2giα/κr)(1 e− ). This displacement am- old noise strength using both a generic biased circuit level plitude∼ ∓ can be subsequently− measured via a typical ho- noise model in Sec.IVA, and a realistic Kerr-cat qubit modyne measurement and the outcome of the homodyne noise model in Sec.IVB. The former demonstrates the measurement can be used to infer the state of the ancilla generic performance increase of the XZZX code when us- qubit. Since the displacement of the readout-resonator ing bias-preserving gates and operations. The latter pro- increases with α, the subsequent homodyne measurement vides a target noise strength in a Kerr-cat architecture of its field becomes more efficient as the cat-size is in- based on a realistic noise model. creased. Consequently, higher fidelity measurements are The threshold is calculated by numerically computing possible by making the cat-size larger. logical failure rates at different sizes, and determining the maximum physical noise strength below which the logi- The interaction Hamiltonian, H , can be achieved readout cal error rate decreases exponentially with system size. In by applying two microwave drives to the SNAIL: one at each simulated run, the circuit implementing d rounds of the frequency difference of the readout resonator and the z noisy syndrome extraction is applied followed by a single SNAIL and the other at the sum of these two frequen- round of noiseless syndrome extraction. The final round cies [see Fig.5(d)]. The first drive gives rise to a beam- can be thought of as a readout, in which each data qubit splitter type coupling (a1†ar + ar†a1), while the second is measured individually, and without loss of generality drive gives the squeezing type coupling (a1†ar†+ara1). The can be considered noiseless as any measurement error in amplitudes of the two drives can be adjusted so that the this round can be equally regarded as having occurred strength of the beam-splitter and squeezing interactions in the previous (noisy) round. The decoder is run on are the same leading to the desired Hamiltonian Hreadout. the observed syndrome and a logical failure occurs if the In the limit of large photon number cat qubits, only one physical error and the correction, computed by the de- drive at the difference frequency is sufficient. In this coder, are homologically distinct. Further details of the limit, the resulting beam-splitter interaction projected Pauli simulation method are provided in AppendixA1. in the cat-basis is P (a1†ar + ar†a1)P gαZ1(ar† + ar) As is standard in most numerical studies of quantum C C 2 2∼α to a very good approximation [ ∗(e− )]. This read- error correcting codes, a Pauli noise model is used in out scheme, called quadrature readoutO has been demon- these threshold calculations, enabling efficient simula- strated experimentally [11]. After the measurement is tion of large system sizes. Although the noise model for complete, the ancilla is projected onto one of the com- Kerr-cat qubits obtained through numerically solving the putational basis states conditioned on the measurement Lindblad master equation is not a Pauli channel, we have outcome. In order to re-initialize the ancilla in + or demonstrated with exact simulations of smaller systems, for the next round of syndrome extraction, the| statesi described in Sec.VB, that a Pauli approximation to the can|−i be rotated from the computational basis by applying noise does not appear to significantly affect calculated X(π/4)† = X( π/4) followed by S†. logical error rates. − 8

The results we present in this section for the XZZX in this simplistic toy model we are assuming probability code are obtained with the conventional open-boundary of correlated Z Z error to be the same as the prob- a ⊗ d layout of the surface code, which has three-qubit checks ability of non-dephasing errors pz/ζ. This gives a total on the boundary [50]. Alternative boundary conditions error probability (or infidelity) of pz(2 + 13/ζ) per CZ (including periodic in both in space and time) were also gate. Here the subscripts d, a refer to the data and an- tested, and similar threshold values were obtained in all cilla qubits respectively. Single qubit noise including idle cases (despite large differences in logical error rates). noise on data qubits and noise during ancilla preparation As was shown in Ref. 49, with the standard layout and have a single qubit Pauli error model with X, Y and Z er- open boundary conditions, the aspect ratio of the code ror probabilities given by pz/ζ, pz/ζ and pz respectively. can be tuned to the bias. By increasing the length of the Faulty measurements are modelled by flipping the ancilla logical Z operator compared to the logical X operator, measurement outcomes with probability pz + pz/ζ. an overall lower logical error rate can be obtained for In the case of ordinary discrete two-level qubits, evo- a given number of qubits. Unless otherwise specified, lution under an interaction Hamiltonian of the form we determine the aspect ratio using small systems by χ [(1 + Za)Id/2 + (1 Za)Xd/2] for time π/2χ would be varying the length of the code relative to its width until required for the CX− operation. It is easy to show that the logical X error rate is comparable to the logical Z if the probability of a Z error during the gate is pz, error rate near threshold. This aspect ratio is then fixed then the resulting error channel is (writing only the when scaling up the code. To make more system sizes Pauli-terms for brevity) pzZaIdρZaId + pz0 IaZdρIaZd + computationally accessible, we restrict to aspect ratios pz0 ZaZdρZaZd + pz00IaYdρIaYd + pz00ZaYdρZaYd with such that the length of the code is a multiple of its width. R π/2 4 p0 = (2p /π) cos (φ)dφ 0.375p and p00 = As with boundary conditions, varying the aspect ratio z z 0 ∼ z z R π/2 2 2 does not appear to appreciably change the value of the (2pz/π) cos (φ) sin (φ)dφ 0.125pz. Consequently, 0 ∼ threshold, provided the aspect ratio is fixed as the system we use this equation for probability assignment for the is scaled up. IaZd,ZaId,ZaZd,IaYd, and ZaYd errors, while all other We also performed simulations of the standard CSS two-qubit Pauli errors are assigned the probability pz/ζ. version of the surface code and an alternative variant of The total infidelity of the CX gate in this case is pz(2 + the surface code proposed for biased noise which we call 10/ζ). the tailored surface code (TSC). The TSC is simply the For the bias-preserving CX gate in the generic noise standard surface code, but with Z checks replaced with Y model, unlike its ordinary counterpart described above, checks, such that a single phase flip error on a data qubit phase-flip errors on the data qubits may convert to creates four adjacent defects. This version of the surface Z Z but they do not convert to Z Y or I Y a ⊗ d a ⊗ d a ⊗ d code was studied in Refs. 39 and 40, where it was found errors. Consequently, in this case Ia Zd and Za Zd to have strong resilience to phase-flip noise. However, we errors are assigned the probability p⊗/2, Z I is⊗ as- z a ⊗ d found that the thresholds obtained with the TSC with signed the probability pz and all other two-qubit Pauli circuit level noise and a modified MWPM decoder were errors are assigned the probability pz/ζ. Note that total somewhat lower than those of the XZZX code, so we have infidelity of the CX gate in this case is larger than the not presented them here. Nevertheless, the TSC may infidelity of the generic non-bias-preserving CX gate by have some advantages for finite-sized systems, which we 2pz/ζ. For large ζ, for example ζ = 100 used here, this discuss in Sec.V. difference becomes negligibly small. The resulting logical error probability vs pz is plotted for the surface code with both bias preserving and non-bias preserving CX gates A. Generic biased circuit noise thresholds in Fig.7(a). For comparison, we have also performed simulations First, we estimate the threshold of the XZZX surface of the conventional CSS surface code using the bias- code using a generic biased circuit noise model. We com- preserving CX gate and the same generic biased er- pare this to the threshold of the standard CSS version ror model. In this case, we used the ‘rotated’ layout of the surface code under the same noise model and find of Ref. 67 with open boundary conditions and 2-qubit that the XZZX code has a significantly higher threshold. checks on the boundary, which has a code distance that Surprisingly, we also find that the XZZX code maintains is a factor of √2 larger than the XZZX code on the unro- a large advantage over the CSS surface code even if a tated geometry using an equal number of physical qubits. standard CX gate for two-level qubits, rather than a bias- The resulting logical error rates are shown in Fig.7(b). preserving CX gate, is used. We do not use the rotated layout for the XZZX code, The generic noise model has two parameters pz and the since the all Z logical operator in this case is oriented noise bias ζ. In the following we assume, without loss of at 45◦ to the lattice boundaries, and we therefore cannot generality, that the Pauli errors are applied after the gate. improve the resilience of the code against phase flips by On the CZ gates, the errors Z I and I Z , are each varying the code’s aspect ratio (unlike with the standard a ⊗ d a ⊗ d assigned the probability pz and all other two-qubit Pauli layout). We also present the logical failure rate as a func- errors are assigned the probability pz/ζ. We remark that tion of √n for both the XZZX code and the CSS surface 9

1 (a) (b) (c) 10− 100 100 XZZX, ζ = 100 CSS, ζ = 100 Biased CX, dX /dZ = 1/3, pz = 0.005 p 1.7% CX ∼ 1 1 10− − 10 2 p 2.1% pCX 1.1% 10− CX ∼ ∼ 2 CSS 10− 2 XZZX 10− Biased CX Standard CX 4 12 12 4 12 12 3 Biased CX × × × × 10− 3 5 15 15 5 15 15 3 9 9 10− Logical error rate 3 × × × × Logical error rate × × Logical error rate 10− 6 18 18 6 18 18 5 15 15 × × × × × × 7 21 21 7 21 21 4 7 21 21 × × × × 10− × × 8 24 24 8 24 24 9 27 27 × × × × × × 4 4 10− 10− 0.005 0.006 0.007 0.008 0.009 0.010 0.011 0.002 0.003 0.004 0.005 0.006 0.007 6 8 10 12 14 16 18 pz pz √n

FIG. 7. Error thresholds of surface code variants with a generic circuit noise model and bias ζ = 100. In (a), we compare the threshold of the XZZX code using bias-preserving CX gates (solid lines) to its threshold using standard non-bias-preserving CX gates (dotted lines). We find that the availability of bias-preserving CX gates leads to a 25% improvement in the threshold. This result highlights the fact that along with total gate error, structure of noise in the∼ gate plays an important role in determining the performance of the surface code variant. In (b), the threshold of the CSS code using bias-preserving CX gates is shown for comparison and is seen to be approximately half that of the XZZX code using the same bias-preserving gates. In this model, both types of CX gate have error rates approximately equal to 2pz (see text for details of the noise model). The crossing points are circled, and the total CX error probability pCX at these points is shown. In (c), we plot the logical error rate of both code variants against √n near the threshold of the CSS code, found in (b), where n is the number of data qubits. In all plots, error bars indicate one standard deviation. code at physical error rate p = 0.005 in Fig.7(c). a white thermal noise channel represented by the Lind- The procedure for choosing the aspect ratio for the bladian κ1(1 + nth) [a]ρ + κ1nth [a†]ρ. Thermal noise XZZX and CSS codes using bias-preserving CX gates was introduces leakage andD in order toD autonomously correct as described above, which yielded the same aspect ratio some of the leakage we add an additional two-photon dis- for both codes near threshold of around 3 : 1. To simplify sipation to the Lindbladian κ [a2]ρ, where the rate of 2D comparison, we used the same aspect ratio for the XZZX two-photon dissipation κ2 is fixed to κ2 = K/10. code with standard CX gates, however, this may not be When the two-photon dissipation is added care must optimal. be taken to adjust the phase of the two-photon drive to As can be seen, the XZZX code using a bias-preserving preserve the cat-qubit subspace [35]. In our simulations, gate achieves the highest threshold pz = 0.98 0.05%, the two-photon dissipation is added while the gates are corresponding to a two-qubit error rate of close± to 2%. being implemented for all but the CX gate. The CX gate This is approximately double the threshold of the CSS requires a time-dependent two-photon dissipation which surface code with the same noise model, which achieves becomes hard to numerically simulate. In practice, it a threshold of pz = 0.50 0.05%. The XZZX code with also increases the number of drives that must be carefully a standard non-bias-preserving± CX gate performs worse phase-tuned in an experiment. than the XZZX code with a bias-preserving gate, achiev- In order to avoid these complications, we simulate the ing a threshold of 0.80 0.05%. However this is clearly CX gate with the following steps. First the master equa- much higher than the CSS± surface code threshold, sug- tion with the Hamiltonian in Eq. (2) and dissipative P gesting that the XZZX code may be a good choice in terms [ρ] = κ (1 + n ) [a ]ρ + κ n [a†]ρ is L i=c,t 1 th D i 1 thD i biased-noise architectures, even when a bias-preserving simulated with φ(t) = πt/T for time T . Next, the in- CX gate is not available. teractions between the control and target Kerr-cats are removed and the Hamiltonian of the system is set to that of two uncoupled Kerr-cats. Finally, the master equation B. Kerr-cat noise thresholds with the two uncoupled Kerr-cats is simulated with time- independent two-photon dissipation, P κ [a2]ρ, in i=c,t 2D i To obtain the noise models used for threshold esti- addition to [ρ] for time T [35]. L mates of the XZZX Kerr-cat code, we first perform a Of course, more phase-flip errors are introduced in master equation simulation of the operations needed for the second step of the simulation. The resulting error the syndrome extraction circuit (Fig.2): the CX and CZ probabilities in this two-step CX gate, which takes time gates, as well as measurement and preparation schemes TCX = 2T , will be roughly twice as high as that in the described above. The simulations are performed for two CX gate simulated with time-dependent two-photon dis- 2 2 cat sizes, α = 1 and α = 6.25. sipation in one step, which takes time TCX = T . The noise channel of the larger cat is more biased than For estimating measurement errors, we performed a ( ) that of the smaller one. Based on the observations in the master equation simulation of the X( π/4) and S † recent experimental work [11], we subject the qubit to gates. Subsequent steps of CD and finite± efficiency ho- 10

2 modyne measurements are simulated by stochastic mas- α = 6.25, κ2 = K/10, nth = 8% ter equation simulations. We assume a homodyne mea- CX gate, TCX = 18.47/K (bias-preserving) TCX = 15.0/K (R CZ R†) surement efficiency of 60% [68]. We have found quan- Single-photon loss rate Infidelity Average bias titative agreement between the analytic expression for Biased-CX Biased-CX homodyne measurement fidelity in Ref. 69 and numeri- (R CZ R†)(R CZ R†) cal results from stochastic master equation simulations. 4 2 1.67 10− 0.047 (0.035) 319 (11.7) For α = 6.25 cat we use the beam-splitter Hamiltonian × 4 2 2.08 10− 0.054 (0.043) 337 (12.0) † × 4 Hreadout = Hcat + g(a1ar + ar†a1), while for α = 1 we 2.77 10− 0.071 (0.057) 357 (12.4) 4.16 × 10 4 0.104 (0.084) 375 (12.7) use Hreadout = Hcat + g(a1† + a1)(ar† + ar)/2. Note that, × − if the probability of bit-flip errors during the CD gate is CZ gate, J 0.08K, T 0.8/K small (a condition which is satisfied in the biased noise ∼ CZ ∼ Single-photon loss rate Infidelity Average bias cat-qubit) this step in the measurement protocol is equiv- 4 3 alent to the well studied longitudinal readout scheme [69]. 1.67 10− 0.0029 1.7 10 × 4 × 3 Consequently, known analytic expressions for longitudi- 2.08 10− 0.0036 1.9 10 × 4 × 3 nal readout error may be used. 2.77 10− 0.0039 2.3 10 4.16 × 10 4 0.0056 3.1 × 103 Observe from Fig.5 that all gate operations [other than × − × X(π/4)] become faster with increasing cat-size. For a Preparation, ε 0.15K, T = 2.6/K ∼ P given α we choose the gate speeds so that the proba- Single-photon loss rate Infidelity bility of leakage outside the cat-manifold is at least an 4 order of magnitude smaller than total infidelity. This 1.67 10− 0.003 × 4 2.08 10− 0.004 choice is justified as the goal, when designing pulses for × 4 2.77 10− 0.005 experimental implementation of gates, is to not limit the × 4 4.16 10− 0.007 gate fidelity by the amount of leakage. While no pulse × shaping is being used in our simulations, such techniques Measurement, κr = 1.0K, g 0.5K ∼ combined with DRAG can be used to further reduce leak- TM = TP + TCD,TCD = 2.5/K age without sacrificing the gate fidelity [70, 71]. In our Measurement efficiency= 60% Single-photon loss rate Infidelity simulations, the leakage probability per qubit, per sta- 4 3 4 1.67 10− 0.008 bilizer round is between 10− and 10− , which is con- × 4 2.08 10− 0.009 × 4 sistent with recent experiments [72]. Additionally, the 2.77 10− 0.011 × 4 two-photon dissipation channel applied during or after 4.16 10− 0.012 × the gates also checks the leakage growth. For the cat sizes considered here, the two-photon dissipation after TABLE I. Main parameters characterizing the operations every CX gate reduces the leakage by a factor of 10. used in syndrome extraction for 6.25-photon Kerr-cat qubit. ∼ The average gate bias is defined as the ratio of the sum of In this way, in our simulations of the qubit opera- the probabilities of I Z, Z I, and Z Z errors and tions, we suppress leakage to a reasonable amount at the sum of the remaining⊗ two-qubit⊗ Pauli errors.⊗ Leakage 4 5 the physical level but then neglect the residual leakage in all the operations is 10− 10− . The effective Pauli in the further surface code simulations. While, this is noise channel of preparation∼ and− measurement operations are the dominant practice in literature for studying the per- naturally biased. The total error in preparation is simply formance of a code, further investigation must be car- the sum of the probability of a Z and a Y error. The bias- ried out to model, quantify, and counteract the effect of preserving CX gate simulation is carried out in two steps, first leakage in the Kerr-cat-surface code architecture. Like by simulating the Hamiltonian in Eq. (2) with the Lindbladian P [ρ] = κ1(1+nth) [ai]ρ+κ1nth [a†]ρ is simulated for in other platforms, further suppression of leakage may L i=c,t D D i require leakage reduction units (LRUs). These units re- time T and in the second step the master equation of two un- coupled Kerr-cats with Lindbladian P κ [a2]ρ + [ρ] duce leakage errors to errors in the computational ba- i=c,t 2 i for time T , so that T = 2T . For fair-comparison,D a two-L sis [73, 74]. In the case of Kerr-cats, leakage in one qubit CX photon dissipation channel after the CXR = RCZR† gate is is unlikely to propagate to others during two-qubit gates. also applied, for time 9/K so that the residual leakage after ∼ Moreover, leakage is primarily confined to the lower en- the CX and CXR gates are the same. The time of R gate is ergy subspace so that the computational-basis errors in- 2.6/K, CZ = 0.8/K giving a total time of 15/K. duced after LRUs remain biased. Of course, additional circuit complexity of the LRUs may effect thresholds. As- sessing the performance of the surface code with LRUs optimized for Kerr-cats is beyond the scope of the current Figure8 shows the logical failure rate of the XZZX work, but will be the topic of future study. surface code as a function of the ratio of the single-photon We simulate the XZZX surface code using the noise loss rate and Kerr-nonlinearity κ/K. The total gate error scales as κ/K and so this is the important metric for channel obtained numerically as described above. Ta- ∝ blesI andII list the average fidelities, bias, and leakage in experimentalists aiming to design the system. different operations obtained from numerical simulations The threshold values obtained for κ/K are given by 2 2 4 2 for cat qubits with α = 6.25 and α = 1.0 respectively. (κ/K) 1.1 10− for α = 1.0 and (κ/K) thres ∼ × thres ∼ 11

2 α = 1.0, κ2 = K/10, nth = 8% (a) 0 CX gate, TCX = 112.6/K 10 1-photon cat Single-photon loss rate Infidelity Average bias Biased-CX Biased-CX

4 0.67 10− 0.02 16.3 × 4 p 3.1% 0.83 10− 0.024 17.2 1 CX × 4 10− ∼ 1.11 10− 0.031 16.4 × 4 Biased CX 1.45 10− 0.041 16.5 × 4 3 9 9 1.67 10− 0.046 16.8 × × × 4 12 12 × × 5 15 15 CZ gate, J 0.013K, TCZ 30.8/K Logical error rate × × ∼ ∼ 2 6 18 18 Single-photon loss rate Infidelity Average bias 10− × × 7 21 21 × × 4 0.67 10− 0.006 9.6 8 24 24 × 4 × × 0.83 10− 0.008 10.6 × 4 1.11 10− 0.01 12.5 0.8 1.0 1.2 1.4 1.6 × 4 1.45 10− 0.013 13.2 κ/K 4 1.67 × 10 4 0.014 14.5 10− × − · Preparation, ε = 0.026K, T 16.5/K (b) 0 P ∼ 10 6.25-photon cat Single-photon loss rate Infidelity p 4 CXR 0.67 10− 0.0015 3.9% × 4 0.83 10− 0.0019 ∼ × 4 1.11 10− 0.0027 1 × 4 10− p 6.5% 1.45 10− 0.0032 CX ∼ 1.67 × 10 4 0.004 × − Biased CX R CZ R Measurement, κr = 0.1K, g 0.06K † ∼ T = T + T ,T = 42/K 2 3 15 15 3 9 9 P CD CD Logical error rate 10− × × × × Measurement efficiency= 60% 4 20 20 4 12 12 × × × × Single-photon loss rate Total error 5 25 25 5 15 15 4 × × × × 0.67 10− 0.017 6 30 30 6 18 18 × 4 × × × × 0.83 10− 0.017 × 4 1.11 10− 0.018 1.5 2.0 2.5 3.0 3.5 4.0 × 4 1.45 10− 0.019 κ/K 4 1.67 × 10 4 0.020 10− × − · TABLE II. Main parameters characterizing the operations FIG. 8. Error thresholds of the XZZX surface code with used in syndrome extraction for 1-photon Kerr-cat qubit. The the circuit noise model for Kerr-cat qubits with a cat size average gate bias is defined as the ratio of the sum of the prob- (a) α2 = 1 photon and (b) α2 = 6.25 photons. The bot- abilities of I Z, Z I, and Z Z errors and the sum of the tom axis shows the ratio of the single-photon loss rate κ or ⊗ ⊗ ⊗ remaining two-qubit Pauli errors. The effective Pauli noise equivalently the inverse T1 time of the SNAIL used to re- channel of preparation and measurement operations are nat- alize the cat and the Kerr-nonlinear constant. The simu- urally biased. The total error in preparation is simply the sum lations also take into account a constant 8% thermal noise 2 of the probability of a Z and a Y error. Like for α = 6.25, and two-photon dissipation κ2 = K/10 added for autonomous the CX is carried out in two steps so that TCX = 2T with leakage correction. The noise channel in (b) is significantly T the time of each step. Leakage in all the operations is more biased than in (a). We find the threshold κ/K of 4 5 4 4 10− 10− . 1.1 10− and 2.5 10− in (a) and (b) respectively which ∼ − corresponds× to a CX× error threshold of 3.1% and 6.5% respectively. For comparison, the threshold∼ with α2 ∼= 6.25 photon cat and non-bias-preserving gate CXR = R CZ R† 4 2 2.5 10− for α = 6.25. To illustrate the physical impli- × with R = exp(iπZt/4) exp(iπXt/4) is also shown in (b). cation of these thresholds, consider a SNAIL-based Kerr Even though the gate infidelity of CXR is lower than that cat with a Kerr-nonlinearity of K/2π = 10 MHz. Then of the bias-preserving CX in Eq. (2), the threshold is lower 4 2 4 (κ/K) 1.1 10 for α = 1.0 implies that the 1.84 10− because the bias is smaller. thres − ∼ × lifetime of the∼ SNAIL× (1/κ) should be longer than 144.7 µs for error correction with the XZZX surface code to 2 4 2 be successful. For α = 6.25, (κ/K)thres 2.5 10− the α = 6.25 cat the CX gate infidelity must be lower ∼ × implies that the lifetime of the SNAIL (1/κ) should be than 6.5%. The average CX gate bias at threshold for longer than 63.6 µs for error correction with the XZZX the α∼2 = 6.25 cat is 351 while that for α2 = 1 cat is surface code to be successful. 16. Clearly, as expected,∼ we find that the XZZX sur- ∼ The most noisy operation in the system is the CX gate face code is able to tolerate more errors in the gates when and we see that, with the α2 = 1.0 cat, the CX gate infi- the cat size is large because the noise bias increases. delity must be lower than 3.1%. On the other hand, with The SNAIL lifetime in the first experimental demon- 12 stration of the Kerr-cat was 15.5 µs, which shows α = 1, we also do not observe a threshold in the overall that an improvement in the lifetime∼ is required to op- error rate over the range of κ/K used in our XZZX code erate below threshold with Kerr-cat qubits. Lifetimes simulations. of 50 100 µs are routinely achieved in superconduct- We also consider the scenario of Ref. 30 where the − ing Josephson-junction circuits [54, 55, 60, 61], so that width of the CSS code is fixed dx = 3, and only the length reaching a lifetime above the T1 threshold appears very of the code dz is scaled up. The same above threshold be- reasonable using state of the art superconducting circuit havior is observed for α = 2.5, even when X logical errors devices and fabrication. Optimizing the gate implemen- are completely neglected. For α = 1, if we neglect logical tations and incorporating a one-step CX will further im- X logical errors, we observe a threshold in the Z logical prove the threshold requirements. Finally, we note that error rate comparable to the threshold in the overall er- 4 the first Kerr-cat experiment [11] showed a strong sup- ror rate of the XZZX code at κ/K = 1.1 10− . Given, pression of frequency fluctuations due to 1/f noise for the however, that X errors are non-negligible at× this bias, the pumped cat, compared to when operating the SNAILmon the overall failure rate still appears above threshold, and as a conventional transmon qubit without pumping. In so error correction with the CSS surface code does not practice, frequency fluctuations of the SNAIL were near appear to be functional in the parameter regions consid- negligible when pumped to the cat subspace. This sug- ered. gest that, albeit the Kerr-cat is a flux-tunable device, re- The high thresholds we obtain with the XZZX code are cent improvements in coherence times for fixed frequency due to the fact that it appears very well suited to Kerr-cat qubits may also carry over to Kerr-cat qubits [75]. qubit noise. One of the major advantages of this pairing is that a full round of syndrome measurements can be performed in a substantially shorter time than with the 1. Comparison with non-bias-preserving CX gate standard CSS surface code and therefore is far less prone to error. This is because, for large α, the gate that takes Note that with Kerr-cat qubits it is possible to im- the longest to execute is the CX gate, while the relative plement a non-bias-preserving CX gate by implement- time required to perform a CZ gate compared to CX ing the gate sequence CXR = R CZ R† with R = tends to zero as α increases. For the CSS surface code, exp(iπZt/4) exp(iπXt/4). The subscript t denotes that readout of an X check requires four data qubits to be the single qubit gate R acts on the target of the CZ. coupled to the ancilla via CX gates, which takes at least We have already discussed how exp(iπZt/4) = S†, time 4TCX. In contrast, every check in the XZZX code exp(iπXt/4) = X(π/4), and CZ gates are simulated. Un- requires two CX gates and two CZ gates, and given that like the bias-preserving CX gate implementation given in these can be parallelized, the entangling operations for a Eq. (2), CXR does not preserve the bias in noise. How- full round of syndrome measurements can be performed ever, the gate can be implemented faster resulting in in time 2(TCX + TCZ). In the limit of large α, the total higher total fidelity. TableI also lists the gate fidelity time taken to apply the gates for a round of syndrome and bias for the CX gate for the α2 = 6.25 photon measurements with the XZZX code is 2TCX, which is R ∼ cat. For fair-comparison, a two-photon dissipation chan- roughly equal to that of a 1D repetition code and half nel after the CXR gate is also applied so that the residual that of the CSS surface code. leakage after the CX and CXR gates are the same for a given κ1, nth, and κ2. To compare the two different gates, we repeat the sim- V. SMALL CODES ulation of the XZZX surface code but this time using the noise channel of the CXR gate. As shown in Fig.8(b), Here we perform exact simulations of different quan- the threshold in this case is lowered by 40% compared ∼ tum error-correcting codes for systems of small size un- to the threshold with the bias-preserving CX. Moreover, dergoing biased noise. In the previous section we were the below-threshold logical error rate for the XZZX code able to analyse the scalability of different quantum error- is significantly higher with the CXR gate then the bias- correcting codes by simulating systems of a large number preserving CX gate, despite the CXR gate having a lower of qubits. We obtained these results using an efficient de- total infidelity. coder and by approximating quantum noise channels as a Pauli noise channel to make simulations tractable. Ex- act simulations enable us to investigate other details of 2. Comparison with the CSS surface code various codes that are specialised to biased noise. We perform several analyses using small-scale simu- We find that the CSS surface code with Kerr-cat noise lations using the Kerr-cat qubit noise model described in this parameter region does not perform as well as the in Sec.IIIB and an optimal decoding algorithm. By fo- XZZX code. At α = 2.5, over the range of κ/K shown cusing on small codes we can perform optimal decoding in TableI, we observe a logical error rate of close to without concern of the efficiency of the decoder. We de- 50%, which does not decrease with system size, indicating scribe the optimal decoder and the methods we use to that the CSS code is far above threshold. At lower bias conduct exact simulations in AppendixA. 13

The smallest realization of a surface code that can cor- in Sec.VD). The probabilities of X, Y and Z logical er- rect at least one X error and one Z error has n = 9 rors separately are shown as well as the total logical error data qubits arranged on a 3 3 lattice and uses the ‘ro- probability (which is the sum of the former three). The tated’ layout of Ref. 67. In× our small-scale simulations noise parameters were chosen to be slightly below the we compare the XZZX code on this layout to the tai- threshold estimated in Fig.8, which is a regime relevant lored surface code (TSC) [40] on the same layout and to to near-term experiments. the one-dimensional repetition code with n = 9 qubits. Certain properties of these codes are reflected in these The TSC is an alternative surface code variant where separate logical error probabilities. For instance, the Z we measure weight-four Pauli-X stabilizer checks and logical is the only one that would be affected in the pres- weight-four Pauli-Y stabilizer checks. This code has been ence of pure phase-flip noise (i.e. noise where the bit-flip demonstrated to have favourable properties to correct for probability is exactly zero). With the rotated layout on biased noise [40, 43]. We use open boundary conditions a 3 3 lattice, the weight of the smallest Z logical of the for all simulations of the n = 9 codes with two-qubit XZZX× code is 3, while for the TSC and repetition codes it checks on the boundaries of the surface code lattices. is 9 [40]. The Z logical is therefore much more protected While in Sec.IV we have shown that the XZZX code than X or Y logicals for the TSC and repetition codes has impressive fault-tolerant thresholds with finite noise compared to the XZZX code, as seen in Fig.9. We see biases, the TSC with open boundary conditions has the that the repetition code’s failures are almost exclusively advantage that it can tolerate up to (n 1)/2 dephasing due to bit-flip errors, rather than phase-flip errors since errors. In contrast, a logical failure can− be introduced in Fig.9 the Z logical errors are nearly undetectable. to the 3 3 XZZX code with as few as two Pauli-Z er- We point out that, like in the Pauli simulations used rors [39].× to determine the threshold in Sec.IV, we have approx- Like the TSC, the n = 9 repetition code can also cor- imated the cat qubits in the exact surface code simu- rect up to four dephasing errors with certainty. Moreover, lations as strict two-level systems, and have neglected it has the additional advantage that it requires only two additional levels into which the system may leak. While entangling gates, rather than four, in the circuits used the leakage suppression with engineered two-photon dis- for stabilizer parity measurements. As such, the stabi- sipation (and noise caused by this process) is still ac- lizer readout circuits for the repetition code are easier counted for in the qubit noise model, there remains a to implement and less error prone than surface code sta- small amount of residual leakage that is ignored. While bilizers. We might therefore expect that the repetition we believe this effect to be small (for the CX gate, the code is more accessible to realisation using near-term ex- probability of leakage is an order of magnitude smaller periments. than that of a non-dominant error), we leave a detailed This convenience of the repetition code comes at the study of this effect to future research. expense that it has no protection against bit-flip errors. Under the following subheadings we discuss our small- Indeed, in Ref. 30, numerical simulations showed that the scale analysis under several different criteria. We be- repetition code does not increase its suppression of logical gin by examining the break-even point of different codes. failures for a realistic biased noise model when increasing This is a pseudo threshold that tells us the noise param- the system size beyond n 10 qubits. Nevertheless, eters below which a small quantum error-correcting code for a small number of qubits∼ and a very large noise bias can outperform the constituent parts from which it is such that the bit-flip errors are very rare, repetition codes built. These numbers give us an indication of the quality may outperform surface codes. We can therefore expect of the qubits we need to demonstrate quantum error cor- variations in the performance of these codes in different rection using small-scale experiments. In addition to this, parameter regimes. we discuss the change in performance obtained by using Our exact simulation results highlight different fea- bias-preserving gates as compared with a more conven- tures of the TSC, XZZX and repetition code when us- tional gate noise model. We also use exact simulations ing using Kerr-cat qubits. The XZZX code in particu- to compare the true noise model extracted from master lar exhibits high performance in both high and low bias equation simulations (but still neglecting leakage) to the regimes, it is relatively insensitive to decoder miscalibra- Pauli-twirl approximation, which was used in Sec.IVB. tion, and it does not suffer from significant increases in Finally, we also use small-scale simulations to investigate logical error rate when standard CX gates are used in- the loss of performance due to a miscalibrated decoder. stead of bias-preserving CX. In contrast, the repetition code will perform better when the bias is high relative to the system size, and the TSC may have an advantage in A. Break-even error rates architectures where CX gates are not so noisy (compared to other operations). An important milestone for experimental demonstra- In Fig.9, we have plotted logical error probabilities tions of error correcting codes is reaching the psuedo- obtained using Kerr-cat qubits and bias-preserving oper- threshold or break-even point. We say that a code is ations alongside those obtained using non-bias preserv- below the pseudo-threshold if the probability of a logi- ing CX and a miscalibrated decoder (which we describe cal error on the encoded qubit after a fixed number of 14

XZZX surface code Tailored surface code Repetition code (a) 0.03 0.08 0.15 α = 1 5 0.06 κ/K = 8.33 10− 0.02 0.10 × 0.04 0.01 0.05 0.02 Logical error probability Logical error probability 0.00 Logical error probability 0.00 0.00 Total X Y Z Total X Y Z Total X Y Z Logical error type Logical error type Logical error type

XZZX surface code Tailored surface code Repetition code (b) 0.10

0.08 α = 2.5 0.20 4 0.20 κ/K = 2.08 10− × 0.06 0.04 0.10 0.10 0.02 Logical error probability Logical error probability Logical error probability 0.00 0.00 0.00 Total X Y Z Total X Y Z Total X Y Z Logical error type Logical error type Logical error type

Bias-preserving CX Miscalibrated decoder Standard CX

FIG. 9. Logical error probabilities for low distance (n = 9) codes obtained by exact simulation. In (a) a moderate bias, 5 corresponding to α = 1, is used with κ/K = 8.33 10− . In (b) a high bias, corresponding to α = 2.5, is used with 4 × κ/K = 2.08 10− . In both cases, the overall noise strength κ is chosen to be slightly below the threshold estimated in Fig.8. Optimal× performance (blue) is compared to the performance obtained with a miscalibrated decoder (red), and non-bias preserving CX gate (orange) for the XZZX surface code, tailored surface code and repetition code. Each data point is evaluated over 12 000 samples and error bars represent the standard error of the mean. Aside for the red bars, which are obtained using a miscalibrated decoder, optimal decoding is used in all of these simulations. rounds of error correction is less than the error probabil- comes from the fact that it requires twice as many CX ity of the noisiest element in the syndrome circuit, which gates compared to the XZZX code (which are far noisier in this case is the CX gate. The number of rounds was than CZ gates). While we do not present the results here, set to three for all of the following simulations. we have found that for simple biased noise models where As shown in Fig.9(b), at high bias α = 2.5 and CZ and CX gates have roughly equal error probabilities, 4 the 3 3 TSC with optimal decoding can outperform the κ = 2.08 10− with a bias-preserving CX and opti- × mal decoding,× the logical error rate for the XZZX code is XZZX code in terms of logical error rate for a wide range 5.41 0.09% and TSC is 5.5 0.1%, which for both codes of biases. However, as efficient near-optimal decoding is equal,± within statistical error,± to the error rate of the appears more complicated for the TSC than the XZZX CX gate of 5.43%. This shows that the pseudo-threshold code for larger systems, more work is still required be- can be achieved with the XZZX and TSC at remarkably fore the TSC can be regarded as a practical or scalable high CX gate infidelity. At this level of bias, where the alternative to the XZZX code under this kind of noise. probability of bit-flip errors is extremely low, the repeti- tion code has an error rate of 0.0162 0.0002, which is ± lower than the both surface code variants. This is due to B. The necessity of bias-preserving CX gates its overall shorter circuit, having fewer qubits idle during CX gate execution and also its ability to correct up to One striking difference between the different codes four dephasing errors with certainty. is the relative importance of using bias-preserving CX In contrast, at a lower bias of α = 1 and with κ = gates. To test this we compared the performance of each 5 8.33 10− with a bias-preserving CX the XZZX code × code using the bias-preserving CX gate to a non-bias pre- is the only code of the three to be below the pseudo- serving preserving CX gate. See Sec.IVA for a descrip- threshold. The advantage of the XZZX code over the tion of the latter CX gate. We defined this gate with the repetition code here is that it can detect and correct bit- same infidelities as the bias-preserving CX gate. flip errors, which are non-negligible at this level of bias. The blue and orange bars in Fig.9 represent the logi- The drawback of the TSC in a Kerr cat architecture cal error probabilities obtained using bias-preserving CX 15 gates and standard (non-bias preserving) CX gates re- lower fidelity in the logical channel for α = 1 and 5 spectively. While both TSC and XZZX surface codes κ/K = 8.33 10− . Furthermore, off-diagonal terms have a higher logical error rate when not using a bias pre- in the logical ×χ matrix are larger than the dominant di- serving CX gate, the increase in error rate is much larger agonal terms. This is not particularly surprising, since for the TSC. At high bias α = 2.5, for the XZZX code, the X-component of the noise is not transformed from the error rate increases by around 70% when using non- the physical to the logical, e.g. a coherent X rotation on bias preserving CX, while the TSC sees a huge increase the first physical qubit of the repetition code results in of nearly 400% compared to when using a bias preserv- an undetectable coherent X rotation by an equal angle ing CX gate. This is likely due to fact that the TSC uses of the logical qubit. twice as many CX gates as the XZZX code, and so using non-bias preserving CX gates reduces the overall noise bias less in the XZZX code than the TSC. This effect is D. Miscalibrated decoding algorithms also observed, although is less pronounced, at lower bias α = 1. Like the threshold results of Sec.IV, these results An exact density matrix simulator can also function as suggest that high performance may be achieved with Kerr an optimal decoder. It can thus be used to probe the per- cats and the XZZX code even if bias-preserving gates formance gains possible with improved decoding, or, con- are not used (although the best performance is likely ob- versely performance loss if the decoder is not optimally tained with bias-preserving gates). calibrated to the noise. To demonstrate the importance Unlike the surface code variants, the repetition code is of calibrating the decoder to biased gate noise, we have entirely dependent on a bias-preserving CX gate. Since compared the performance of a miscalibrated decoder to it is only designed to correct phase-flip errors, the addi- the optimal decoder. With the physical CX gate chosen tional bit-flip errors introduced with a non-bias preserv- to be bias preserving, the miscalibrated decoder is tuned ing CX gate drastically reduce its performance. This can to the noise model of the standard CX (rather than the be seen in both α = 1 and α = 2.5 plots, where the logi- bias-preserving CX). Thus the miscalibrated decoder is cal error rate is roughly equal to the probability of an X not properly adapted to the noise on the CX gate, how- error occurring on a data qubit somewhere in the error ever it still correctly takes into account the bias on all correction circuit. other elements. The difference in performance observed between the miscalibrated decoder and optimal decoder tells us how important it is for the decoder to have an C. Interrogating the Pauli-twirl approximation accurate characterisation of bias on the CX gate. As shown in Fig.9 at high bias α = 2.5 the XZZX code The threshold simulations in Sec.IV are performed is much less sensitive to decoder miscalibration than the with a Pauli noise model, which allows efficient simu- TSC code at this bias. Using a miscalibrated decoder re- lation of large systems. This Pauli noise model is ob- sulted in an increase of less than 20% increase in logical tained by taking the Pauli twirl approximation of the error rate for the XZZX code, compared to over 100% for full noise model, which has non-Pauli or coherent terms. the TSC. The insensitivity of the XZZX code to miscali- The Pauli twirl approximation effectively sets non-Pauli brated decoding suggests that near optimal performance terms in the noise to zero. We do not need to apply the may be achieved with this code using heuristic decoding Pauli twirl to the noise in the exact simulations (albeit methods (like minimum weight perfect matching) where we still neglect leakage to other states). However, our a precise characterisation of the noise model does not exact simulation data suggests that non-Pauli terms in need to be used. the physical noise have a negligible contribution to the At lower bias α = 1 we see a smaller increase in logi- performance of the surface code variants. cal error rate when the decoder is miscalibrated for both Using the Pauli twirl of the noise maps on the physical XZZX and TSC surface codes. This suggests that the level in the exact simulation produces an average logical importance of tailored decoding increases as the bias in- χ matrix with diagonal terms (which can be interpreted creases. In contrast, the logical error rate of the repeti- as logical Pauli error probabilities) within statistical error tion code, due to its relatively simple decoding require- of those obtained using the full (non-Pauli) noise maps ments, appears almost completely unaffected by decoder on the physical level. Furthermore, off-diagonal terms in miscalibration at both low and high bias. the logical χ matrix are at least two orders of magnitude smaller than the diagonal terms for the averaged logical channel, suggesting that the averaged logical channel is VI. DISCUSSION close to a stochastic Pauli channel. This provides some justification for the use of the Pauli twirl approximation We have proposed a robust architecture for fault- in the threshold calculations in Sec.IV. tolerant quantum computing in which Kerr-cat qubits In contrast, the repetition code preserves coherence on are used as the elementary qubits in an XZZX surface the logical level. The full noise-model results in slightly code. The resulting system demonstrates exceptional worse performance than its Pauli twirl, with about 10% thresholds that can be met using a superconducting cir- 16 cuit architecture with parameters that appear fairly mod- ful in the production of the conceptual figure for the est compared to the current state of the art. We have Kerr-cat/SNAIL-surface code. A.S.D. was supported made conservative assumptions about the Kerr-cat pho- by JST, PRESTO Grant No. JPMJPR1917, Japan. ton number and imperfections such as thermal noise and B.J.B., A.L.G. and D.K.T. are supported by the Aus- measurement inefficiency. For the parameters used in our tralian Research Council via the Centre of Excellence in simulations, the Kerr-cat has an anharmonicity, given by Engineered Quantum Systems (EQUS) project number ωgap, comparable to that of a transmon, leading to fast CE170100009, and by the Army Research Office(ARO) gates. Note, however, that our simulations were per- under Grant Number: W911NF-21-1-0007. B.J.B. also formed without pulse optimization, which may give sig- received support from the University of Sydney Fellow- nificant improvements [76]. Further optimization is also ship Programme, and A.L.G. is also supported by the possible by improving the decoding algorithm, as our re- Australian Research Council via a Discovery Early Ca- sults were obtained using a standard minimum-weight reer Research Award, project number DE190100380. S.P. perfect-matching decoder [58]. is supported by the ARO under grant number W911NF- It remains to make a detailed analysis of the resource 18-1-0212. Most of the numerical computation in this cost to use our error-correcting system to perform a com- work was carried out at the Yukawa Institute Com- putation below some target logical failure rate. This will puter Facility. Access to high-performance computing be important to compare the performance of our pro- resources was also provided by the National Computa- posal with others [30, 36–47]. Resource estimates are tional Infrastructure (NCI), which is supported by the highly sensitive to the details of noise in laboratory sys- Australian Government. tems [77], and can be improved dramatically with op- timizations to the code (e.g. to the code layout and decoder) and hardware. Hence, we leave such calcula- Appendix A: Surface code simulation method tions to future work. Nevertheless, given the remarkable threshold that our system exhibits, we expect we will be We used two types of numerical simulation to evalu- able to demonstrate a scalable system by realizing our ate the performance of cat-surface codes. For large sys- proposal using modern technology. tems, in order to determine error thresholds, we used Theoretical research should progress hand-in-hand circuit-level Pauli simulations and for small codes more with experimental development. To this end, details in accessible to near term implementation we used exact our simulations have been motivated by the dominant state-vector simulations. Using two simulations methods sources of noise in Kerr-cat qubits that have been ob- provides complementary insights into surface code per- served in recent experiments [11]. We should expect that formance: the Pauli simulations allow system size scal- as the system is scaled up in terms of both the size of ing analysis to be performed, while the exact simulations the cats, and the number of qubits on a chip, that other allow general types of error and to see the potential per- sources of errors such as cross-talk, drive-induced heating formance achievable with optimal decoding. and leakage must be accounted for. It will be important to develop better methods to characterize the noise in physical realizations of our error-correcting systems [78], 1. Pauli simulation and to refine our quantum error-correcting codes accord- ingly. The Pauli simulations rely on the fact that the sur- The architecture motivating our present study is based face code is implemented with a Clifford circuit, which on the Kerr-cat qubit realized with on-chip supercon- transforms Pauli operators to Pauli operators. One can ducting SNAILs. This is a promising platform for realiz- therefore simulate a noisy circuit by sampling Pauli errors ing a large-scale surface-code architecture using systems in the circuit (according to an error model of each circuit presently available in the laboratory. The principles in element) then, by commuting the Pauli errors through to the design and fabrication of a large-scale SNAIL chip the end of the circuit, determine which ancilla measure- are very similar to other circuit QED based approaches, ments are flipped (i.e. the syndrome) and the residual for instance quantum processors based on tunable trans- error on the data qubits. mons [79]. The scalability of our proposal with existing The Pauli simulation used for threshold calculation as- technologies suggests a practical solution to achieve fault- sumes the initial state is a noiseless surface code state, tolerant quantum computing in the near term. to which a number of rounds of noisy syndrome mea- surements are applied, followed by a single final round of noiseless syndrome measurements. The assumption ACKNOWLEDGEMENTS that the final round of check measurements is noisy is justified in practice since the final readout of a surface- The authors would like to thank S. Bartlett, P. Bonilla code qubit is performed by single qubit measurements Ataides, and S. Flammia for conversations about de- on the data qubits for which there is no distinction be- coders and tailored codes. S.P would like to thank N. tween measurement errors and a data qubit error. After Frattini for numerous discussions which were also help- sampling an error, the decoding algorithm is run on the 17 corresponding syndrome to determine a correction. If the have applied idle noise to data qubits and ancilla qubits computed correction and the sampled error belong to the such that the total strength of the idle noise on each qubit same logical equivalence class, we regard the error cor- is the same as if the gates were interleaved to measure the rection as successful. If not, a logical failure is said to checks in parallel. This idle noise includes the noise on have occurred. By repeating the simulation many times qubits left idle during measurement and preparation of and counting the number of logical failures, the logical ancilla as well as noise on qubits left idle while CX gates failure rate can be determined. To determine the error are applied to other qubits. Noise on qubits left idle dur- threshold, the simulation is performed at varying error ing rounds of CZ gates is relatively small compared to rates as the code distance d is increased. For code dis- CX gates and was neglected in our exact simulations (al- tance d the number of rounds of noisy measurements was though it is included in our Pauli simulations). Thus the set to d. The threshold was determined to be the noise total amount of noise in the circuit is roughly equivalent strength below which increasing d led to an exponential in our exact simulations, to the circuit with parallelized decrease in logical failure rate. check readout which would be used in practice, however due to the different gate order there will be some differ- ences in how errors propagate between data and ancilla 2. Exact simulation qubits.

The exact simulation method involves storing the N qubit density operator of the noisy code state ρ as a 4N Appendix B: Kerr-cat and purely dissipative cat component vector in memory and updating it as gates, gate fidelities noise and measurements are applied with matrix-vector multiplication. Given the exponential scaling in required In this section, with the help of a simple numerical ex- memory, the method is only suitable for small system ample, we outline an important difference between Kerr- sizes. However, it has the advantage of allowing arbitrary cat and dissipative cats which leads to higher-fidelity gate noise models to be studied (rather than restricting to operations with Kerr-cat qubits. Consider the implemen- Pauli noise). Keeping track of the full density operator tation of the S-gate in the Kerr-cat qubit. The required also means that the simulation algorithm may be used for Hamiltonian is optimal decoding. This type of decoding is similar to the approach described in Ref. 80 and 81. By calculating how H = Hcat + ε(a† + a) 2 2 2 2 2 a linearly independent set of encoded states (represented Hcat = Ka† a + Kα (a† + a ) as 2 2 matrices) transforms under error correction for − (B1) a fixed× syndrome s, it is possible to calculate the logical channel s, represented as a 4 4 χ matrix, conditioned In order to understand the effect of the microwave drive on the syndrome.E The optimal× Pauli correction is then ε(a† + a) let us look at how the photon annihilation and the one that minimises the distance of the logical channel creation operators affect the computational cat states. to the identity i.e. Action of the photon annihilation operator causes transi- 1 tions between the cat states, a α± = αr± α∓ . Here argmin L s I . (A1) L I,X,Y,Z p 2α2 2α2 |C i ± |C i ∈ || ◦ E − || r = (1 e− )/(1 + e− ). The creation operator For Pauli noise, this is equivalent to maximum-likelihood on the other− hand, not only causes transitions between decoding. One simplification made for the exact simula- the cat states, but it can also cause transitions out- tion was that the check measurements were performed side the cat-subspace. In fact, in the limit of large α, sequentially, i.e. all the gates and the ancilla of a check a† α± α α∓ + ψ1∓ , where ψ1∓ are the first ex- were applied before measuring the next check. This is cited|C i states ∼ ± with|C i odd| andi even photon-number| i parity in contrast to the Pauli simulation, where all checks were respectively. These states are separated from the cat measured simultaneously with an interleaved 2 qubit gate subspace by the energy gap ωgap . While, the applied pattern (described in Sec.II). This simplification to the drive is resonant with the cat-qubit| | frequency, it is off- exact simulation was made to save computational time, resonant from the excited states by an amount ωgap . since performing check measurements in parallel neces- Consequently, this drive can only cause small amount| of| sitates keeping all ancilla qubits in memory (in addition population transfer to the leakage state. More impor- to the data qubits), rather than a single ancilla qubit at tantly, the amount of leakage on average (ε/ω )2, ∝ gap a time. We remark that there is a simulation method, decreases with ωgap [35]. described in Ref. 82, based on novel tensor-network con- An S-gate with| the| purely dissipative cat requires the 2 2 2 traction techniques which can obviate the large memory Hamiltonian H = i(κ2/2)α (a† a )+ε(a† +a) and the 2 − cost of the simple exact simulation method we have used, engineered dissipation κ2 [a ]ρ. The gate time is π/8αε, however the overall computational costs in this method which is the same as theDS gate time with Kerr-cats. In still remain daunting. this case, a† also causes transition to the first excited There is some choice in where to put idle noise in the manifold α± ψ1∓ . The excited state population simulation when checks are measured sequentially. We quickly decays|C i back → | toi the cat-subspace at rate 2κ α2. ∼ 2 18

2 2 (a) 10− The quantity 2κ2α is also called the Lindbladian gap. · ∼ 2 2 The residual excited state population is (ε/2κ2α ) . Be- Kerr cat 2 1.0 cause of the transition matrix elements α± a ψ1± 2α Purely dissipative cat 2 hC | | i ∼ and α± a ψ1∓ = 0, the dissipation causes transitions Kerr + two-photon dissipation hC | | i 0.8 within the even parity and odd parity subspaces respec- tively. On the other hand, the leakage caused transi- 0.6 tions between the even and odd parity subspace. Hence, cooling down of the excited state population leads to 0.4 additional phase-flip errors. These additional phase-flip errors in the purely-dissipative cats are also termed as non-adiabatic errors in literature [28, 83]. The amount 0.2 3 Average gate infidelity of phase flips, ε/κ2α , increases with ε [83]. Fortu- nately, the additional∝ phase-flip errors are noticeably ab- 0.0 sent in Kerr-cats. Consequently, gates with Kerr-cats have higher fidelity than those with purely-dissipative 3 (b) 10− cats. 1.5 ·

1.0

Leakage 0.5

We demonstrate the effect of additional phase-flip er- 0.0 rors in purely dissipative cats by simulating the S gate 5 10 15 20 and comparing its fidelity with the fidelity of S gate with Kerr-cats. Figure 10 shows this comparison. To keep the K/, κ /2 2 analysis simple we neglect other sources of decoherence in either of the two cat qubits. In the simulations we fix ε, FIG. 10. (a) S gate infidelity and (b) leakage with Kerr- α = 2 and hence the gate time π/8αε is the same for the cats (blue), purely dissipative cats (red), and Kerr-cats with two cats. We choose κ2 = 2K so as to keep the energy two-photon dissipation (orange), κ2 = K/10. In (b) the grey and Lindbladian gaps (or equivalently the residual leak- dashed line indicates the approximate analytic expression for 2 2 2 2 age) approximately the same in the two cats. Figure 10 average residual leakage (ε/4Kα ) , (ε/2κ2α ) . In these sim- ulations we take α = 2 and all other sources of decoherence, clearly shows that the S gate infidelity can be signifi- such as single photon loss and thermal excitations, are ne- cantly higher for Kerr-cats than purely dissipative cats glected. Note that the oscillations in the leakage for Kerr- for the same amount of leakage. We also show the the cats results from off-resonant population transfer back and infidelity of the S gate in the hybrid approach with small forth between the cat and leakage space due to the Hamilto- two-photon dissipation κ2 = K/10 included in addition nian dynamics. These oscillations disappear when a smoother to the Hamiltonian Eq. (B1). The result shows that ad- time-dependent pulse ε(t) is used for implementing the S- dition of small amount of two-photon dissipation during gate. Nonetheless, the average behaviour agrees well with the the Kerr-cat S gate does not degrade the gate fidelity in simple analytic estimate given by the grey line. the parameter regime used in this paper.

[1] M. D. Reed, L. DiCarlo, S. E. Nigg, L. Sun, L. Frunzio, J. Wenner, A. N. Korotkov, A. N. Cleland, and J. M. S. M. Girvin, and R. J. Schoelkopf, Realization of three- Martinis, Superconducting quantum circuits at the sur- qubit quantum error correction with superconducting cir- face code threshold for fault tolerance, Nature 508, 500 cuits, Nature 482, 382 (2012), arXiv:1109.4948 [quant- (2014), arXiv:1402.4848 [quant-ph]. ph]. [3] D. Nigg, M. M. E. A. Martinez, P. Schindler, M. Hen- [2] R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, nrich, T. Monz, M. A. Martin-Delagado, and R. Blatt, E. Jeffry, T. C. White, J. Mutus, A. G. Fowler, B. Camp- Experimental quantum computations on a topologically bell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, encoded qubit, Science 345, 302 (2014), arXiv:1403.5426 C. Neill, P. O’Malley, P. Roushan, A. Vainsencher, [quant-ph]. 19

[4] J. Kelly, R. Barrends, A. G. Fowler, A. Megrant, E. Jef- tum memories, Rev. Mod. Phys. 87, 307 (2015), frey, T. C. White, D. Sank, J. Y. Mutus, B. Campbell, arXiv:1302.3428 [quant-ph]. Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, I.-C. Hoi, [18] B. J. Brown, D. Loss, J. K. Pachos, C. N. Self, and J. R. C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan, Wootton, Quantum memories at finite temperature, Rev. A. V. J. Wenner, A. N. Cleland, and J. M. Martinis, Mod. Phys. 88, 045005 (2016), arXiv:1411.6643 [quant- State preservation by repetitive error detection in a su- ph]. perconducting , Nature 519, 66 (2015), [19] E. T. Campbell, B. M. Terhal, and C. Vuillot, Roads arXiv:1411.7403 [quant-ph]. towards fault-tolerant universal quantum computation, [5] A. D. C´orcoles, E. Magesan, S. J. Srinivasan, A. W. Nature 549, 172 (2017), arXiv:1612.07330 [quant-ph]. Cross, M. Steffen, J. M. Gambetta, and J. M. Chow, [20] D. Aharonov and M. Ben-Or, Fault-tolerant quantum Demonstration of a quantum error detection code using a computation with constant error, in STOC ’97 Pro- square lattice of four superconducting qubits, Nat. Com- ceedings of the Twenty-Ninth Annual ACM Symposium mun. 6, 6979 (2015), arXiv:1410.6419 [quant-ph]. on Theory of Computing (1997) p. 176, arXiv:quant- [6] M. Takita, A. D. C´orcoles, E. Magesan, B. Abdo, ph/9611025. M. Brink, A. W. Cross, J. M. Chow, and J. M. Gambetta, [21] I. L. Chuang, D. W. Leung, and Y. Yamamoto, Bosonic Demonstration of weight-four parity measurements in the quantum codes for amplitude damping, Phys. Rev. A 56, surface code architecture, Phys. Rev. Lett. 117, 210505 1114 (1997). (2016), arXiv:1605.01351 [quant-ph]. [22] P. T. Cochrane, G. J. Milburn, and W. J. Munro, Macro- [7] N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas, scopically distinct quantum-superposition states as a B. Vlastakis, Y. Liu, L. Frunzio, S. M. Girvin, L. Jiang, bosonic code for amplitude damping, Phys. Rev. A 59, M. Mirrahimi, M. H. Devoret, and R. J. Schoelkopf, Ex- 2631 (1999). tending the lifetime of a quantum bit with error correc- [23] D. Gottesman, A. Kitaev, and J. Preskill, Encoding a tion in superconducting circuits, Nature 536, 441 (2016), qubit in an oscillator, Phys. Rev. A 64, 012310 (2001). arXiv:1602.04768 [quant-ph]. [24] M. Mirrahimi, Z. Leghtas, V. V. Albert, S. Touzard, [8] L. Hu, Y. Ma, W. Cai, X. Mu, Y. Xu, W. Wang, Y. Wu, R. J. Schoelkopf, L. Jiang, and M. H. Devoret, Dynami- H. Wang, Y. P. Song, C. L. Zou, S. M. Girvin, L.-M. cally protected cat-qubits: a new paradigm for universal Duan, and L. Sun, Quantum error correction and univer- quantum computation, New J. Phys. 16, 045014 (2014), sal gate set operation on a binomial bosonic logical qubit, arXiv:1312.2017 [quant-ph]. Nature Physics 10.1038/s41567-018-0414-3 (2019). [25] J. M. Gertler, B. Baker, J. Li, S. Shirol, J. Koch, and [9] C. Fl¨uhmann,T. L. Nguyen, M. Marinelli, V. Negnevit- C. Wang, Protecting a bosonic qubit with autonomous sky, K. Mehta, and J. P. Home, Encoding a qubit in quantum error correction, Nature 590, 243 (2021). a trapped-ion mechanical oscillator, Nature 566, 513 [26] C. Vuillot, H. Asasi, Y. Wang, L. P. Pryadko, and (2019). B. M. Terhal, Quantum error correction with the toric [10] P. Campagne-Ibarcq, A. Eickbusch, S. Touzard, E. Zalys- gottesman-kitaev-preskill code, Phys. Rev. A 99, 032344 Geller, N. E. Frattini, V. V. Sivak, P. Reinhold, S. Puri, (2019). S. Shankar, R. J. Schoelkopf, et al., Quantum error cor- [27] K. Noh and C. Chamberland, Fault-tolerant bosonic rection of a qubit encoded in grid states of an oscillator, quantum error correction with the surface-gkp code, Nature 584, 368 (2020). arXiv:1908.03579 (2019). [11] A. Grimm, N. E. Frattini, S. Puri, S. O. Mundhada, [28] J. Guillaud and M. Mirrahimi, Repetition cat qubits for S. Touzard, M. Mirrahimi, S. M. Girvin, S. Shankar, and fault-tolerant quantum computation, Phys. Rev. X 9, M. H. Devoret, Stabilization and operation of a Kerr-cat 041053 (2019), arXiv:1904.09474 [quant-ph]. qubit, Nature 584, 205 (2020), arXiv:1907.12131 [quant- [29] A. L. Grimsmo, J. Combes, and B. Q. Baragiola, Quan- ph]. tum computing with rotation-symmetric bosonic codes, [12] R. Lescanne, M. Villiers, T. Peronnin, A. Sarlette, Physical Review X 10, 011058 (2020). M. Delbecq, B. Huard, T. Kontos, M. Mirrahimi, and [30] C. Chamberland, K. Noh, P. Arrangoiz-Arriola, E. T. Z. Leghtas, Exponential suppression of bit-flips in a qubit Campbell, C. T. Hann, J. Iverson, H. Putterman, T. C. encoded in an oscillator, Nat. Phys. 16, 509 (2020), Bohdanowicz, S. T. Flammia, A. Keller, G. Refael, arXiv:1907.11729 [quant-ph]. J. Preskill, L. Jiang, A. H. Safavi-Naeini, O. Painter, [13] Y. Ma, Y. Xu, X. Mu, W. Cai, L. Hu, W. Wang, X. Pan, and F. G. S. L. Brand˜ao,Building a fault-tolerant quan- H. Wang, Y. P. Song, C. L. Zou, and L. Sun, Error- tum computer using concatenated cat codes (2020), transparent operations on a logical qubit protected by arXiv:2012.04108 [quant-ph]. quantum error correction, Nat. Phys. 16, 827 (2020), [31] K. Noh, C. Chamberland, and F. G. Brand˜ao, Low arXiv:1909.06803 [quant-ph]. overhead fault-tolerant quantum error correction with [14] P. W. Shor, Fault-tolerant quantum computation, in Pro- the surface-gkp code, arXiv preprint arXiv:2103.06994 ceedings of 37th Conference on Foundations of Computer (2021). Science, FOCS ’96 (IEEE Computer Society, USA, 1996) [32] B. M. Terhal, J. Conrad, and C. Vuillot, Towards scalable p. 56, arXiv:quant-ph/9605011. bosonic quantum error correction, Quantum Science and [15] A. Kitaev, Fault-tolerant quantum computation by Technology 5, 043001 (2020). anyons, Ann. Phys. 303, 2 (2003), arXiv:quant- [33] A. Joshi, K. Noh, and Y. Y. Gao, ph/9707021. processing with bosonic qubits in circuit qed, Quantum [16] E. Dennis, A. Kitaev, A. Landahl, and J. Preskill, Topo- Science and Technology (2021). logical , J. Math. Phys. (N.Y.) 43, 4452 [34] S. Puri, S. Boutin, and A. Blais, Engineering the quan- (2002), arXiv:quant-ph/0110143. tum states of light in a kerr-nonlinear resonator by [17] B. M. Terhal, Quantum error correction for quan- two-photon driving, npj Quantum Inf. 3, 18 (2017), 20

arXiv:1605.09408 [quant-ph]. Comput. 1, 43 (2009). [35] S. Puri, L. St-Jean, J. A. Gross, A. Grimm, N. E. Frat- [53] B. J. Brown and D. J. Williamson, Parallelized quan- tini, P. S. Iyer, A. Krishna, S. Touzard, L. Jiang, A. Blais, tum error correction with fracton topological codes, S. T. Flammia, and S. M. Girvin, Bias-preserving gates Phys. Rev. Research 2, 013303 (2020), arXiv:1901.08061 with stabilized cat qubits, Sci. Adv. 6, eaay5901 (2020), [quant-ph]. arXiv:1905.00450 [quant-ph]. [54] P. Jurcevic, A. Javadi-Abhari, L. S. Bishop, I. Lauer, [36] P. Aliferis and J. Preskill, Fault-tolerant quantum com- D. Borgorin, M. Brink, L. Capelluto, O. Gunluk, putation against biased noise, Phys. Rev. A 78, 052331 T. Itoko, N. Kanazawa, et al., Demonstration of quan- (2008), arXiv:0710.1301 [quant-ph]. tum volume 64 on a superconducting quantum comput- [37] P. Aliferis, F. Brito, D. P. DiVincenzo, J. Preskill, ing system, Quantum Sci. Technol. 6, 025020 (2021), M. Steffen, and B. M. Terhal, Fault-tolerant computing arXiv:2008.08571 [quant-ph]. with biased-noise superconducting qubits: a case study, [55] M. Kjaergaard, M. E. Schwartz, J. Braum¨uller, New J. Phys 11, 013061 (2009), arXiv:0806.0383 [quant- P. Krantz, J. I.-J. Wang, S. Gustavsson, and W. D. ph]. Oliver, Superconducting qubits: Current state of play, [38] A. M. Stephens, W. J. Munro, and K. Nemoto, Annu. Rev. Condens. Matter Phys. 11, 369 (2020), High-threshold topological quantum error correction arXiv:1905.13641 [quant-ph]. against biased noise, Phys. Rev. A 88, 060301 (2013), [56] Z. Nussinov and G. Ortiz, A symmetry principle for arXiv:1308.4776. topological quantum order, Ann. Phys. 324, 977 (2009), [39] D. K. Tuckett, S. D. Bartlett, and S. T. Flammia, Ultra- arXiv:cond-mat/0702377 [cond-mat.str-el]. high error threshold for surface codes with biased noise, [57] B. J. Brown, W. Son, C. V. Kraus, R. Fazio, and Phys. Rev. Lett. 120, 050505 (2018), arXiv:1708.08474 V. Vedral, Generating topological order from a two- [quant-ph]. dimensional using a duality mapping, New [40] D. K. Tuckett, A. S. Darmawan, C. T. Chubb, S. Bravyi, J. Phys. 13, 065010 (2011), arXiv:1105.2111 [quant-ph]. S. D. Bartlett, and S. T. Flammia, Tailoring surface codes [58] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N. for highly biased noise, Phys. Rev. X 9, 041031 (2019), Cleland, Surface codes: Towards practical large-scale arXiv:1812.08186 [quant-ph]. quantum computation, Phys. Rev. A 86, 032324 (2012), [41] X. Xu, Q. Zhao, X. Yuan, and S. C. Benjamin, arXiv:1208.0928 [quant-ph]. High-threshold code for modular hardware with asym- [59] Y. Tomita and K. M. Svore, Low-distance surface codes metric noise, Phys. Rev. Applied 12, 064006 (2019), under realistic quantum noise, Phys. Rev. A 90, 062320 arXiv:1812.01505. (2014), arXiv:1404.3747 [quant-ph]. [42] M. Li, D. Miller, M. Newman, Y. Wu, and K. R. Brown, [60] E. A. Sete, N. Didier, A. Q. Chen, S. Kul- 2D compass codes, Phys. Rev. X 9, 021041 (2019), shreshtha, R. Manenti, and S. Poletto, Parametric- arXiv:1809.01193 [quant-ph]. resonance entangling gates with a tunable coupler (2021), [43] D. K. Tuckett, S. D. Bartlett, S. T. Flammia, and B. J. arXiv:2104.03511 [quant-ph]. Brown, Fault-tolerant thresholds for the surface code in [61] J. Kelly, R. Barends, A. G. Fowler, A. Megrant, E. Jef- excess of 5% under biased noise, Phys. Rev. Lett. 124, frey, T. C. White, D. Sank, J. Y. Mutus, B. Campbell, 130501 (2020), arXiv:1907.02554 [quant-ph]. Y. Chen, et al., State preservation by repetitive error [44] S. Huang, M. Newman, and K. R. Brown, Fault-tolerant detection in a superconducting quantum circuit, Nature weighted union-find decoding on the , Physi- 519, 66 (2015), arXiv:1411.7403 [quant-ph]. cal Review A 102, 10.1103/physreva.102.012419 (2020), [62] Z. Leghtas, S. Touzard, I. M. Pop, A. Kou, B. Vlastakis, arXiv:2004.04693 [quant-ph]. A. Petrenko, K. M. Sliwa, A. Narla, S. Shankar, M. J. [45] J. Guillaud and M. Mirrahimi, Error rates and re- Hatridge, et al., Confining the state of light to a quantum source overheads of repetition cat qubits (2021), manifold by engineered two-photon loss, Science 347, 853 arXiv:2009.10756 [quant-ph]. (2015), arXiv:1412.4633 [quant-ph]. [46] L. H¨anggli,M. Heinze, and R. K¨onig,Enhanced noise [63] S. Touzard, A. Grimm, Z. Leghtas, S. O. Mundhada, resilience of the surface–Gottesman-Kitaev-Preskill code P. Reinhold, C. Axline, M. Reagor, K. Chou, J. Blumoff, via designed bias, Phys. Rev. A 102, 052408 (2020), K. M. Sliwa, et al., Coherent oscillations inside a quan- arXiv:2004.00541 [quant-ph]. tum manifold stabilized by dissipation, Phys. Rev. X 8, [47] O. Higgott and N. P. Breuckmann, Subsystem codes with 021005 (2018), arXiv:1705.02401 [quant-ph]. high thresholds by gauge fixing and reduced qubit over- [64] G. J. Woeginger, Open problems around exact algo- head (2020), arXiv:2010.09626 [quant-ph]. rithms, Discrete Applied Mathematics 156, 397 (2008). [48] X.-G. Wen, Quantum orders in an exact soluble model, [65] J. Gambetta, A. Blais, D. I. Schuster, A. Wallraff, Phys. Rev. Lett. 90, 016803 (2003), arXiv:quant- L. Frunzio, J. Majer, M. H. Devoret, S. M. Girvin, ph/0205004. and R. J. Schoelkopf, Qubit-photon interactions in a [49] J. P. Bonilla Ataides, D. K. Tuckett, S. D. Bartlett, S. T. cavity: Measurement-induced dephasing and number Flammia, and B. J. Brown, The XZZX surface code, Nat. splitting, Phys. Rev. A 74, 042318 (2006), arXiv:cond- Commun. 11, 2172 (2021), arXiv:2009.07851 [quant-ph]. mat/0602322 [cond-mat.mes-hall]. [50] S. B. Bravyi and A. Y. Kitaev, Quantum codes on a [66] S. Puri, A. Grimm, P. Campagne-Ibarcq, A. Eickbusch, lattice with boundary (1998), arXiv:quant-ph/9811052 K. Noh, G. Roberts, L. Jiang, M. Mirrahimi, M. H. De- [quant-ph]. voret, and S. M. Girvin, Stabilized cat in a driven non- [51] J. Edmonds, Paths, trees and flowers, Can. J. Math. 17, linear cavity: A fault-tolerant error syndrome detector, 449 (1965). Phys. Rev. X 9, 041009 (2019), arXiv:1807.09334 [quant- [52] V. Kolmogorov, Blossom V: A new implementation of a ph]. minimum cost perfect matching algorithm, Math. Prog. [67] H. Bombin and M. A. Martin-Delgado, Optimal re- 21

sources for topological two-dimensional stabilizer codes: form for superconducting transmon qubits with coher- Comparative study, Phys. Rev. A 76, 012305 (2007), ence times exceeding 0.3 milliseconds, Nature Communi- arXiv:quant-ph/0703272. cations 12, 1 (2021). [68] S. Touzard, A. Kou, N. Frattini, V. Sivak, S. Puri, [76] F. Motzoi, J. M. Gambetta, P. Rebentrost, and F. K. A. Grimm, L. Frunzio, S. Shankar, and M. Devoret, Wilhelm, Simple pulses for elimination of leakage in Gated conditional displacement readout of supercon- weakly nonlinear qubits, Phys. Rev. Lett. 103, 110501 ducting qubits, Phys. Rev. Lett. 122, 080502 (2019), (2009), arXiv:0901.0534 [cond-mat.mes-hall]. arXiv:1809.06964 [quant-ph]. [77] M. E. Beverland, B. J. Brown, M. J. Kastoryano, and [69] N. Didier, J. Bourassa, and A. Blais, Fast quantum non- Q. Marolleau, The role of entropy in topological quan- demolition readout by parametric modulation of longitu- tum error correction, J. Stat. Mech. 2019, 073404 (2019), dinal qubit-oscillator interaction, Phys. Rev. Lett. 115, arXiv:1812.05117 [quant-ph]. 203601 (2015), arXiv:1504.04002 [quant-ph]. [78] R. Harper, S. T. Flammia, and J. J. Wallman, Efficient [70] M. Werninghaus, D. J. Egger, F. Roy, S. Machnes, F. K. learning of quantum noise, Nat. Phys. 16, 1184– (2020), Wilhelm, and S. Filipp, Leakage reduction in fast super- arXiv:1907.13022 [quant-ph]. conducting qubit gates via optimal control, npj Quantum [79] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, Inf. 7, 14 (2021), arXiv:2003.05952 [quant-ph]. R. Barends, R. Biswas, S. Boixo, F. G. Brandao, D. A. [71] Z. Chen, J. Kelly, C. Quintana, R. Barends, B. Campbell, Buell, et al., Quantum supremacy using a programmable Y. Chen, B. Chiaro, A. Dunsworth, A. Fowler, E. Lucero, superconducting processor, Nature 574, 505 (2019). et al., Measuring and suppressing quantum state leakage [80] A. S. Darmawan and D. Poulin, Tensor-Network Simu- in a superconducting qubit, Physical review letters 116, lations of the Surface Code under Realistic Noise, Phys. 020501 (2016), arXiv:1509.05470 [quant-ph]. Rev. Lett. 119, 040502 (2017), arXiv:1607.06460 [quant- [72] M. McEwen, D. Kafri, Z. Chen, J. Atalaya, K. Satzinger, ph]. C. Quintana, P. Klimov, D. Sank, C. Gidney, A. Fowler, [81] A. S. Darmawan and D. Poulin, Linear-time general de- et al., Removing leakage-induced correlated errors coding algorithm for the surface code, Phys. Rev. E 97, in superconducting quantum error correction, Nature 051302 (2018), arXiv:1801.01879 [quant-ph]. Communications 12, 1 (2021), arXiv:arXiv:2102.06131 [82] C. Huang, X. Ni, F. Zhang, M. Newman, D. Ding, [quant-ph]. X. Gao, T. Wang, H.-H. Zhao, F. Wu, G. Zhang, C. Deng, [73] P. Aliferis and B. M. Terhal, Fault-tolerant quantum H.-S. Ku, J. Chen, and Y. Shi, Alibaba cloud quantum computation for local leakage faults, Quant. Inf. Comp. development platform: Surface code simulations with 7, 139 (2007), arXiv:quant-ph/0511065. crosstalk (2020), arXiv:2002.08918 [quant-ph]. [74] M. Suchara, A. W. Cross, and J. M. Gambetta, Leakage [83] C. Chamberland, A. Kubica, T. J. Yoder, and suppression in the toric code, Quant. Inf. Comp. 15, 997 G. Zhu, Triangular color codes on trivalent graphs (2015), arXiv:1410.8562 [quant-ph]. with flag qubits, New J. Phys. 22, 023019 (2020), [75] A. P. Place, L. V. Rodgers, P. Mundada, B. M. Smitham, arXiv:1911.00355 [quant-ph]. M. Fitzpatrick, Z. Leng, A. Premkumar, J. Bryon, A. Vrajitoarea, S. Sussman, et al., New material plat-