<<

Gate error analysis in simulations of computers with

D. Willsch,1 M. Nocon,1 F. Jin,1 H. De Raedt,2 and K. Michielsen1, 3 1Institute for Advanced Simulation, J¨ulichSupercomputing Centre, Forschungszentrum J¨ulich,D-52425 J¨ulich,Germany 2Zernike Institute for Advanced Materials, University of Groningen, Nijenborgh 4, NL-9747AG Groningen, The Netherlands 3RWTH Aachen University, D-52056 Aachen, Germany (Dated: November 10, 2017) In the model of gate-based quantum computation, the qubits are controlled by a sequence of quantum gates. In superconducting systems, these gates can be implemented by voltage pulses. The success of implementing a particular gate can be expressed by various metrics such as the average gate fidelity, the diamond distance, and the . We analyze these metrics of gate pulses for a system of two superconducting transmon qubits coupled by a resonator, a system inspired by the architecture of the IBM Quantum Experience. The metrics are obtained by numerical solution of the time-dependent Schr¨odinger equation of the transmon system. We find that the metrics reflect systematic errors that are most pronounced for echoed cross-resonance gates, but that none of the studied metrics can reliably predict the performance of a gate when used repeatedly in a . Keywords: quantum computation; superconducting qubits; quantum circuits; ; ; average fidelity; fault-tolerance thresholds; transmon qubits

I. INTRODUCTION In this paper, we limit the analysis to two coupled by one resonator as fundamental errors can be Over the last decades, tremendous effort has gone into best understood for a small system containing only the building a universal quantum computer. In theory, such basic constituents. For the implemented gates, we eval- a device can solve certain problems, such as factoring, ex- uate the average gate fidelity [9], the diamond distance ponentially faster than classical digital computers. The [10], and the unitarity [11]. The obtained gate fidelities leading technological prototypes are based on supercon- agree with those reported in state-of-the-art experiments ducting transmon qubits containing on the order of 10 [12–15]. However, we find that the diamond error rates of qubits [1–4]. IBM provides public access to such a quan- all gates are larger than 2% (see also [7]). The precision of tum processor through the IBM Quantum Experience the gates is limited by the presence of non-computational (IBMQX) [5]. states in the transmons and the resonator. The corre- However, as reported in a recent independent bench- sponding errors occur naturally in the unitary evolution mark [6], IBM’s five-qubit quantum processor does not of the total system, but they have a detrimental effect yet meet the fundamental requirements for a computing on the computational subspace. For instance, we find device. For this reason, the underlying architecture and that they appear incoherent when looking at the compu- its operation call for a deeper analysis, one that goes be- tational subspace only, and they cannot be represented yond perturbation theory, rotating wave approximations, by Pauli channels [16]. and assumptions about Lindblad forms and Markovian In particular, for CNOT gates based on echoed cross- dynamics [7]. resonance pulses [17–19], we find a systematic error that We study the real-time dynamics of such quantum sys- can be reproduced in experiments on the IBMQX. We tems in detail by solving the time-dependent Schr¨odinger also propose a different, one-pulse CNOT gate that does equation (TDSE) for a generic model Hamiltonian. For not suffer from this error. this purpose, we have developed efficient product-formula The paper is structured as follows. In Section II, we algorithms that are tailored to key features of the model describe the simulation model and explain how quantum Hamiltonian [8]. This allows us to simulate each Gaus- gates are implemented by microwave pulses. This sec- sian control pulse that is used in experiments to realize a tion also sketches the optimization procedure that we use certain quantum gate, as dictated by the computational to find optimal pulses for the qubit system. Section III model of a quantum computer. We have implemented gives a summary of the gate metrics that frequently serve a parameter optimization scheme for obtaining the best as error rates in experimental and theoretical studies. pulse parameters for the gates on the transmon system. In Section IV, we present the gate metrics of the opti- This scheme makes use of the fact that in the simulation, mized pulses and analyze their behavior with respect to we have the advantage of getting the full information of repeated applications of the gates. Additionally, we per- the system dynamics at any time t. In brief, the simu- form identity operations and entanglement experiments lated system can be seen as a faithful model of an ideal as proposed in [6] to compare the performance of the gate quantum processor that works exactly as quantum theory sets with “real” quantum programs. Conclusions drawn dictates. from our analysis are given in Section V. 2

algorithm for the total unitary time-evolution operator TABLE I. Parameters for the model Hamiltonian given in Utotal(t) defined by Ψ(t) = Utotal(t) Ψ(0) (see Ap- Eqs. (1)–(3), inspired by the device parameters of the quan- | i | i tum processor of the IBMQX [5]. All energies are expressed pendix A for details on the algorithm). This solution is expanded in the product basis k m1 m2 where k is the in GHz (~ = 1). The CPB qubits are operated in the trans- | i| i| i mon regime with EJi/ECi ≈ 10, and their frequencies ωi number of in the resonator, and mi = 0, 1, 2,... and anharmonicities αi resulting from diagonalizing the CPB label the transmon eigenstates (i.e., the eigenstates of Hamiltonian are given for reference. The resonator operates HCPB given by Eq. (2) for ngi(t) = 0) of qubit i = 1, 2. at frequency ωr/2π = 7 GHz. Its coupling to the qubits is Thus, the result of the simulation is the set of coefficients weak as |gi|  |ωi − ωr|. akm1m2 (t) defined by

Qubit i ECi/2π EJi/2π ωi/2π αi/2π gi/2π X Ψ(t) = akm m (t) k m1 m2 . (5) 1 1.204 13.349 5.350 −0.350 0.07 | i 1 2 | i| i| i km m 2 1.204 12.292 5.120 −0.353 0.07 1 2 The system is initialized in a computational basis state Ψ(0) = m1m2 , where the computational subspace | i | i II. SIMULATION MODEL AND METHOD is defined by m1m2 = k = 0 m1 m2 for m1, m2 0, 1 . Note that| thei simulation| i| explicitlyi| i includes non-∈ { } We consider a system of superconducting transmon computational states outside of this subspace. qubits [20]. The transmons are coupled by a transmission line resonator, which is essentially a quantum harmonic A. Quantum gates oscillator [21]. The publicly accessible five-qubit quan- tum processor of the IBMQX is of this type [5]. The model Hamiltonian of a system of N transmons For architectures of the transmon type, quantum gates coupled to a resonator reads [20] are implemented by applying microwave voltage pulses to the qubits. This is mathematically modeled through the H = HCPB + HRes, (1) control fields ngi(t) in Eq. (2). We consider a generic N sum of shaped microwave pulses applied on each qubit X  2  HCPB = ECi(ˆni ngi(t)) EJi cosϕ ˆi , (2) as described in [23], namely − − i=1 N X dr ngi(t) = Ωij(t) cos(ωij t γij), (6) † X † − HRes = ωraˆ aˆ + ginˆi(ˆa +a ˆ ). (3) j i=1 dr where Ωij(t) is the envelope of pulse j on qubit i, ωij Here, H describes the Cooper pair box (CPB) qubits CPB is the corresponding drive frequency, and γij is an offset whose capacitive energies ECi and Josephson energies phase. To model a situation close to experiments (cf. E are set in the transmon regime [20],n ˆ is the number Ji i [22, 24]), the envelopes Ωij(t) are Gaussians of the form operator of qubit i, and the bounded phase operatorϕ ˆi is its conjugate. The qubits are controlled by the exter-  (t−T/2)2   T 2  exp 2σ2 exp 8σ2 nal control field ngi(t), which is directly proportional to − − − ΩG(t) = Ω0   , (7) the voltage pulses applied to the qubit. Thus, quantum T 2 1 exp 2 gates are implemented by choosing a certain pulse form − − 8σ for ngi(t) (see [22, 23]). The resonator is described by † where Ω0 is the amplitude and T the time of the pulse, raising and lowering operatorsa ˆ anda ˆ, respectively. It and σ = T/4 defines the width of the Gaussian. operates at the microwave frequency ωr and its capacitive dr The drive frequencies ωij in Eq. (6) are usually set coupling strength to the qubits is given by gi. to one of the qubit frequencies ωi (see Table I). How- The values of the parameters in Eqs. (1)–(3) are given ever, as the presence of the resonator can slightly shift in Table I. In what follows, we consider the case N = 2 as these frequencies [21], we adjust ωi by measuring local the key results are most clearly demonstrated for a small rotations of the qubits in the lab frame. We do this isolated system of qubits. by initializing the system in the state Ψ(0) = ++ The dynamics of the joint system of the two transmons | i | i with + = ( 0 + 1 )/√2 and letting it evolve freely for and the resonator can be obtained by studying the time 4000 ns.| i On| thei respective| i Bloch spheres, both qubits evolution of the state Ψ(t) of the system. This state is | i then rotate about the z-axis. The frequencies of these the solution of the TDSE (~ = 1) rotations yield the shifted qubit frequenciesω ¯i. We ob- ∂ tainω ¯1/2π = 5.346 GHz andω ¯2/2π = 5.118 GHz. i Ψ(t) = H(t) Ψ(t) , (4) The computational states of the qubits at some time ∂t | i | i t > 0 are defined in a so-called locally rotating frame for the Hamiltonian given in Eq. (1). We obtain the so- R [16, 22]. This essentially removes the just mentioned lution numerically by implementing a product-formula rotation, so that the state ++ remains unchanged if | i 3 no quantum gate is applied. Mathematically, this is im- CR1 plemented by multiplying the coefficients of the solution Control given in Eq. (5) by time-dependent phase factors, yield- R ing a (t) := exp(it(¯ω1m1 +ω ¯2m2))akm m (t). km1m2 1 2 Target We have implemented the same quantum gate set as CR2 supported by the IBMQX [25]. Accordingly, a typical quantum gate sequence takes between 50 ns and 15 µs. Control In the following, we explain how the pulses are defined and modeled. Target CR4 Control

1. Single-qubit gates Target

Single-qubit rotations on the Bloch sphere can be re- alized by applying a Gaussian pulse with drive frequency FIG. 1. Pulse sequences for the three different realizations of a dr CNOT gate studied in this paper. Gaussians implement Xπ/2 ωi =ω ¯i on qubit i (see Eqs. (6) and (7)). In this case, and Xπ rotations, and flat-topped Gaussians represent cross- the amplitude Ω0 and the phase γ define the angle and resonance (CR) pulses (i.e., they oscillate at the frequency the axis of rotation, respectively [22] (e.g. γ = 0 for ro- ω¯T of the target qubit). The CR1 gate consists only of flat- tations about the x-axis, or γ = π/2 for rotations about topped Gaussian pulses at the target frequencyω ¯T . The CR2 the y-axis). gate is an echoed CR gate containing two additional Xπ pulses We utilize the virtual Z-gate (VZ gate) described in on the control qubit and one Xπ/2 pulse on the target qubit. [23] and used in the IBMQX [5] to implement rotations The CR4 gate is a four-pulse echoed CR gate that contains an additional X pulse on the target qubit. See Fig. 6 in about the z-axis. This means that instead of applying π Appendix B for the full pulse specifications. another pulse, we simply change the phase γ of all the following pulses (see Appendix B for details). Unfortunately, as transmons cannot be represented by 2. Two-qubit gates ideal two-level systems, such pulses may induce leakage out of the computational subspace, meaning that the so- We implement the CNOT gate by making use of the lution in Eq. (5) also has contributions from higher levels cross-resonance (CR) effect [17, 18]. The basic idea sim- such as mi = 2 . This effect can be mitigated by includ- | i ply amounts to applying another Gaussian pulse to the ing another pulse in Eq. (6) proportional to the deriva- control qubit C = 1, 2 but at the drive frequencyω ¯ ˙ T tive ΩG(t) with a phase shift of π/2. This technique of the target qubit T = C. Furthermore, the pulse is goes under the name of DRAG and has become standard stretched over a longer6 time period such that the Gaus- for transmon systems [24, 26]. Therefore we also adopt sian in Eq. (7) becomes a flat-topped Gaussian with 3σ DRAG in defining the pulses. rise time where σ = 5 ns (cf. [13]). For the single-qubit gates, we take T = 83 ns for the Various schemes have been used in experiments to con- gate duration of the Gaussian envelope ΩG(t) given by struct a CNOT gate based on the CR effect [13, 19, 27– Eq. (7), inspired by the choice made for the IBMQX [5]. 29]. We implement three particularly interesting repre- In summary, a single-qubit pulse on qubit i is defined sentatives to compare their performance with the perfor- by mance of the ideal system. The first is a simple one-pulse CR gate (CR1) that we found by including an additional driving of the target qubit, inspired by the observation π ngi(t) = ΩG(t) cos(¯ωit γ) + βΩ˙ G(t) cos(¯ωit γ ), in [13] (see Fig. 5 in Appendix B for details). The sec- − − − 2 ond is a two-pulse echoed CR gate (CR2) which is cur- (8) rently also used on the five-qubit quantum processor of the IBMQX [5]. The third is a four-pulse echoed CR gate with the parameters (Ω0, β, γ) being the amplitude, the (CR4) that has recently shown better performance (al- DRAG coefficient, and the phase, respectively (see Ap- beit worse fidelity) for an experiment on quantum error- pendix B for the theory behind these parameters). As detecting codes [29]. The pulse sequences of the three outlined below, we optimize the pulse parameters to im- CNOT gates are summarized in Fig. 1. plement ideal single-qubit rotations of the type Xπ/2 and The pulse parameters such as amplitudes, times, and Xπ. The former serves as a primitive to generate ar- phases for each sequence are scanned over ranges sug- bitrary single-qubit gates as in experiments [23]. The gested by the theory. As in the case of single-qubit gates, latter is used exclusively as a component in the echoed this provides initial values for the pulse optimization pro- two-qubit gates. cedure. 4

B. Pulse optimization and straightforwardness in the experimental implemen- tation (e.g. average gate fidelity [9]), while others such as The goal of applying the control pulses is to real- the diamond distance stem from mathematical consider- ize a certain transformation U (the unitary quantum ations [10]. As recently demonstrated by Sanders et al. gate) on the computational subspace. For two qubits, [33], the relation between fidelity and diamond distance is this subspace is spanned by the computational basis not direct in that the impressively high fidelities reported 00 , 01 , 10 , 11 , so U is essentially a complex 4 4 in experiments are not sufficient for fault-tolerant quan- matrix.{| i | Examplesi | i | fori}U are Xi (a π/2 rotation of qubit× tum computation, in contrast to claims made by other π/2 groups [14, 34] i about the x-axis) or CNOTij (a controlled NOT oper- ation where i, j with i = j denote the control and the In the following, we give an overview of the three met- target qubit, respectively)6 [16]. rics that we have selected to assess the quality of quantum As the simulation produces the full state Ψ(t) given gates. Evaluating these metrics requires the definition by Eq. (5), we can construct the actual transformation| i of appropriate quantum channels, which are completely matrix M implemented by a certain pulse. We do this positive (CP) linear maps on the space of density op- by initializing Ψ(0) in each of the four computational erators ρ [35]. For a two-qubit system, ρ is a Hermitian | i 4 4 matrix. Using the language from Section II B where basis states, evolving the system under the application × of the pulse according to Eq. (4), and extracting the four U denotes the ideal unitary gate matrix and M denotes the actual transformation implemented on the computa- complex coefficients a000 through a011 from the solution tional subspace, we define the ideal quantum channel id given by Eq. (5) (formally, M is a 4 4 block of the total G × and the actual quantum channel ac as time-evolution operator Utotal(t), see Eq. (A1)). Each G run for one of the four computational basis states pro- † duces one column of M, including the complex phases id(ρ) = UρU , (10) G that each basis state acquires in the time evolution. The † ac(ρ) = MρM . (11) four runs can be performed in parallel. G The aim is to find ideal pulse parameters so that the computational matrix M approaches the ideal gate ma- It can easily be seen that both maps are CP. However, trix U, up to a global phase. Note that the computa- note that in most cases M † = M −1 because of additional 6 tional space is a subspace of the whole non-computational states present in transmon systems. = span k m1 m2 , so it is by no means clear that This implies that ac is not trace-preserving (the alter- G −1 MH will be{| unitaryi| i| (andi} in almost all cases, it is not). native channel MρM does not preserve Hermiticity). We use a multidimensional optimization scheme in- For convenience, we define the discrepancy channel −1 troduced by Nelder and Mead [30, 31] to optimize the (ρ) = ac( id (ρ)) which approaches unity for a per- pulse parameters. Note that we only use the optimization Dfect controlG pulse.G procedure to refine the initial pulse parameters obtained from the theory [13, 23, 26] (see also Appendix B). The objective function to be minimized is given by A. Average gate fidelity ∆(M,U) = M zU 2 , (9) k − kF where is the Frobenius norm, and z = The average gate fidelity is defined as [9] p k·kF Tr(MU †)/Tr(MU †)∗ is a phase factor that minimizes ±the global phase difference between both matrices. We Z Favg = d ψ ψ ( ψ ψ ) ψ . (12) have tested other gate error rates as objective functions | i h | D | ih | | i (see Section III) and found that Eq. (9) produces the best results. In general, we have 0 Favg 1, and the maximum After optimizing the pulse parameters, we further im- ≤ ≤ fidelity Favg = 1 is attained in the ideal case where the prove the gates using the VZ phase correction to com- discrepancy channel is unity. pensate for off-resonant rotation errors etc. [23] (see Ap- D pendix B for details). In experiments, this number is estimated by a protocol called randomized benchmarking (RB) [36, 37]. However, as in our simulation we have access to the error chan- nel given by Eq. (11), we do not need to implement the III. GATE ERROR RATES RB protocol. Instead, we evaluate Eq. (12) directly by sampling the integrand and averaging it over states from Various quantities have been used in experiments and the computational subspace, as done in [38]. Specifically, P studied in the literature to measure the success of imple- we generate 100,000 random states ψ = cij ij by | i ij | i menting a quantum gate by a certain control pulse [32]. drawing real and imaginary parts of cij from a normal Some of these measures are motivated by their simplicity distribution and normalizing the state afterwards. 5

B. Diamond distance C. Unitarity

The error rate of a is defined in For transmon qubits, leakage into higher non- terms of the diamond distance [33] computational levels during the application of a pulse is a known problem [38, 42]. Mathematically, this leads 1 to the situation that the evolution of the computational η = 11 . (13) † −1 ♦ ♦ subspace is not unitary, resulting in M = M and thus 2 kD − k 6 Tr ac(ρ) < Tr ρ in terms of Eq. (11), so the process is This quantity is mathematically relevant for the thresh- notG trace-preserving. To quantify such effects, Wallman old theorem [39] that is often cited in the literature to et al. have proposed a quantity called unitarity [11] given argue that arbitrarily long, fault-tolerant quantum com- by putation is possible. The best known quantum error- Z correcting codes require η to be on the order of 1% or d  0 † 0  ♦ u = d ψ Tr ac( ψ ψ ) ac( ψ ψ ) , (17) less [33]. d 1 | i G | ih | G | ih | − Evaluating Eq. (13) is nontrivial as the diamond norm 0 where (ρ) = ac(ρ 11/d) Tr [ ac(ρ 11/d)] /√d. of a superoperator is defined by maximizing the trace Gac G − − G − A 0 Note that by construction, the errors we observe are norm Tr over all ancillary Hilbert spaces and all joint densityk·k operators on 0 [10, 33]. However,H one systematic, unitary, and coherent (in the sense of [11]) H ⊗ H on the total Hilbert space . Hence this quantity char- can show that this is equivalent to minimizing over all H generalized Choi-Kraus representations of [40]. As we acterizes how incoherent these errors appear on the com- have (ρ) = MU †ρUM † ρ, this amountsA to computing putational subspace. A − The integral in Eq. (17) is evaluated in the same way  ! 1/2 as the average gate fidelity given in Eq. (12). 1    † † −† −1 MU η♦ = inf UM , 11 S S 2 S  − 11 − 2 IV. RESULTS ! 1/2   MU †  UM †, 11 SS† . (14) ∗ 11 In this section, we analyze the performance of the op- 2  timized single-qubit and two-qubit quantum gate sets. First, we evaluate the gate metrics described in the pre- Here, denotes the matrix 2-norm (i.e. the maximum vious section. Then we study the repeated application k·k2 singular value [41]), and S is an invertible complex 2 2 of gates that mathematically constitute identity opera- × matrix. We solve this minimization problem by first sam- tions. Finally, we repeat a set of pling over 10,000 random matrices S and then running experiments to compare the simulated results with the the same optimization procedure that we already imple- corresponding experimental results obtained on the IB- mented for the pulse optimization in Section II B. This MQX [6]. scheme was found to produce reliable results, equal to the exact η♦ up to two significant digits for all tested cases for which we found closed expressions (see [40]). A. Gate metrics There are two asymptotically tight bounds for the error rate η♦ in terms of the average gate fidelity Favg given The gate metrics of the optimized pulses are given in by Eq. (12) [33], namely Table II. The overall performance of the pulses is close to optimal but still not perfect. Especially the error rate Pauli d + 1 η given by Eq. (13) is always bounded above 2%, even η = (1 Favg), (15) ♦ ♦ d − though our quantum-theoretical model of the transmon ub q qubit architecture does not account for decoherence or η = d(d + 1)(1 Favg), (16) ♦ − noise. The average gate fidelities are in the same ball- park as those reported for experiments based on the same for which we have ηPauli η ηub. In these expres- pulse schemes [1, 12–15]. In fact, the single-qubit gate ♦ ≤ ♦ ≤ ♦ sions, d = 2N = 4 is the dimension of the computational fidelities are slightly worse than the ones reported in ex- subspace. The upper bound leads to the estimate that periments. We shall see below, however, that the actual two-qubit gates need to reach a fidelity of 0.999995 in performance of the gates in quantum circuits is much order to qualify for fault-tolerant quantum computation better. We also observed similar gate metrics for shorter with known quantum error-correcting codes [33]. The single-qubit gates of about T = 10 ns, but then the per- lower bound is saturated if the error is a Pauli channel, formance of repeated applications of the pulses was worse and the difference η ηPauli represents the “badness” of (data not shown). ♦ ♦ the noise. We shall see− that all gates under investigation Note that we always find η ηPauli, so the dominant ♦ ♦ yield η ηPauli. errors are non-Pauli errors and belong to the “bad” class ♦  ♦ 6

TABLE II. Gate metrics for the set of optimized quantum gate pulses described in Section II. The distance objective ∆ from the optimization is given in Eq. (9). The average gate fidelity F , the error rates η , ηPauli, ηub, and the unitarity u are defined avg ♦ ♦ ♦ in Eqs. (12), (13), (15), (16), and (17), respectively.

Type Gate T in ns ∆ F η ηPauli ηub u avg ♦ ♦ ♦ 1 Xπ/2 83 0.0022 0.9946 0.027 0.0068 0.33 0.990 2 X Xπ/2 83 0.0023 0.9942 0.028 0.0073 0.34 0.989 1 Xπ 83 0.0013 0.9949 0.020 0.0064 0.32 0.990 2 Xπ 83 0.0015 0.9943 0.023 0.0071 0.34 0.989 CNOT 71.865 0.0013 0.9842 0.029 0.0198 0.56 0.969 CR1 12 CNOT21 158.193 0.0023 0.9951 0.033 0.0062 0.31 0.991 CNOT 431.949 0.0061 0.9943 0.048 0.0071 0.34 0.991 CR2 12 CNOT21 369.116 0.0056 0.9947 0.048 0.0066 0.32 0.992 CNOT 652.954 0.0054 0.9934 0.049 0.0083 0.36 0.989 CR4 12 CNOT21 572.623 0.0045 0.9946 0.044 0.0068 0.33 0.991

TABLE III. Comparison of the error rate η for a single (a) X simulation (b) CR1 simulation ♦ 1 1 CNOT12 gate, twenty successive CNOT12 gates, and four suc- 1 X CNOT12 0.8 π/2 0.8 cessive QFT applications. A QFT contains five CNOT gates 2 CNOT21 Xπ/2 and two additional X pulses. The numbers reported are the 1 0.6 Xπ 0.6 error rates defined in Eq. (13), but the same relative trends are 2 0.4 Xπ 0.4 true for the average gate infidelity 1 − Favg given by Eq. (12) and the unitarity u given by Eq. (17). Error Rate 0.2 Error Rate 0.2

Pulse CNOT1 CNOT20 QFT4 5 10 15 20 5 10 15 20 Pulse Application n Pulse Application n CR2 0.048 0.73 0.27 (c) CR2 simulation (d) CR4 simulation CR4 0.049 0.33 0.32 1 1 CNOT CNOT 0.8 12 0.8 12 CNOT21 CNOT21 0.6 0.6 of errors [33]. Interestingly, there is almost one order 0.4 0.4 of magnitude difference between the actual error rate η♦ and the optimal bounds ηPauli and ηub calculated from Error Rate 0.2 Error Rate 0.2 ♦ ♦ the gate fidelity Favg according to Eqs. (15) and (16). 5 10 15 20 5 10 15 20 In Table II, it is also seen that a higher fidelity Favg cor- Pulse Application n Pulse Application n responds to a higher unitarity u. From this we conclude that leakage is still the dominant source of error limiting FIG. 2. (Color online) Evolution of the error rate given by the gate fidelity, even though the techniques DRAG [24] Eq. (13) after n applications of a certain gate. (a) Single-qubit and VZ phase correction [23] have been included in the Gaussian derivative pulses defined by Eq. (8); (b)-(d) two- construction of the pulses. It seems that the presence qubit CNOT gates based on the cross-resonance effect. While of the resonator and the entangling transverse coupling CR1 includes only one CR pulse on each of the transmons, cause this limitation (see also [43, 44]). CR2 and CR4 employ additional X gates to echo out certain errors (see Section II and Appendix B).

B. Repeated gate applications ter than the transformation M(T ) on the computational For each gate primitive of our universal gate set, we subspace suggests. However, this also means that the study n = 1,..., 20 repeated applications of the corre- non-computational levels are more significant in the time sponding pulses on each of the four computational ba- evolution than a simple two-state description of a quan- sis states. After each application of a pulse with to- tum computer can capture (see also [6]). tal time T , we construct the full transformation matrix In Fig. 2, we plot the error rate η♦ given by Eq. (13) n M(nT ) of the computational subspace as described in to compare M(nT ) with U . We choose η♦ because this Section II B and compare it with the ideal gate matrix quantity is central for fault-tolerant quantum computa- U n. Interestingly, we observed that the actual transfor- tion, and it also includes the statistical distance of the mation M(nT ) is always closer to U n than the product experimentally measurable output distribution [33]. We M(T )n, which means that the actual pulse performs bet- observed the same qualitative behavior for the distance 7 objective given by Eq. (9) and the average gate infidelity (a) IBMQX (using CR2) (b) CR1 simulation 1 Favg (data not shown). 1 1 − 00 00 The single-qubit pulses perform reasonably well. Al- 0.8 | i 0.8 | i 10 10 though the error rates are always above 2% (see Table II), 0.6 | i 0.6 | i they stay approximately constant even after successive 0.4 0.4 applications of the gates (see Fig. 2(a)). For the two- 0.2 0.2 qubit gates, the error rate already starts growing after Stat. Distance Stat. Distance 0 0 two applications of the CNOT gate (see Fig. 2(b),(c),(d)). 0 5 10 15 20 0 5 10 15 20 This is most clearly visible for the CR1 gate, for which Number of CNOT12 Number of CNOT12 the error rate makes a jump after every second applica- (c) CR2 simulation (d) CR4 simulation tion. Note that the echoed CNOT gates CR2 and CR4 1 1 00 00 are found to work equally well if the control and the tar- 0.8 | i 0.8 | i get qubit are exchanged. In experiments, usually only 10 10 0.6 | i 0.6 | i one type of CNOT is implemented because the CR in- teraction strength is weaker for the other type [5, 28, 29] 0.4 0.4 (see also Fig. 5 in Appendix B). The best performance is 0.2 0.2 Stat. Distance Stat. Distance seen for the four-pulse echoed gate CR4, even though the 0 0 0 5 10 15 20 0 5 10 15 20 gate metrics in Table II do not suggest that. Note that Number of CNOT12 Number of CNOT12 the same discrepancy between worse metrics and better actual performance was also observed in recent quantum FIG. 3. (Color online) Statistical distances between the ideal error-detection experiments on the IBMQX [29]. result and the measured distribution of states for the circuit n As an additional comparison between CR2 and CR4, CNOT12 |ψi with |ψi = |00i (stars) and |ψi = |10i (circles). we analyze both schemes in four applications of the (a) Experimental results on the IBMQX; (b)-(d) simulation Quantum Fourier Transform (QFT). The full circuit for results. Generically, the echoed CNOT versions show worse QFT4 also involves 20 CNOT gates (along with 8 ad- performance on state |00i than on state |10i, both in the ditional X pulses, see Appendix C). Based on the error experiment and the simulation. Interestingly, this systematic error is not present in the one-pulse version CR1. rates for 20 CNOT gates presented in Fig. 2, we might be led to believe that CR4 performs better in general. (a) (b) However, from the results presented in Table III, we see 1 0.2 that this is not true. Hence, the error rate does not pre- F 0 dict the behavior of a gate in actual applications. Note F1 0.5 F2 0.2 that the same is true for the average gate infidelity and E − E1/2 0.4 the unitarity (data not shown). 0 − 0.6 It is worth mentioning that the gate with the worst − 0.5 0.8 fidelity and the worst unitarity (CNOT12 from the group − − 1 CR1, see Table II) is in fact the fastest and performs − 1 1.2 relatively well after repeated use, as demonstrated in − 0 30 60 90 120 150 180 − 0 90 180 270 360 Fig. 2(b). Similarly, the gate with the best fidelity ϑ2 ϑ1(ϑ1 = 0) ϑ1 = ϑ2 (CNOT21 from the group CR1) performs worst. This − means that, although the analyzed gate metrics can be FIG. 4. (Color online) Results for a set of quantum circuits√ used to study errors in one application of a gate, they do creating and characterizing the singlet state (|01i − |10i)/ 2 not characterize the performance of the gates when used as a function of the angles ϑ1 and ϑ2. F1/2(ϑ1, ϑ2) are single- in a (see also [6]). qubit averages and F (ϑ1, ϑ2) is a two-qubit correlation func- tion. The corresponding theoretical expectations are given by E1/2(ϑ1, ϑ2) and E(ϑ1, ϑ2). (a) ϑ1 = 0 fixed and ϑ2 variable; (b) ϑ1 = ϑ2 variable. Apart from a small systematic deviation around ϑ1 = ϑ2 = 135, the agreement is almost perfect. C. Comparison with the IBM Quantum Experience

As the simulation model is inspired by the quantum distance processor of the IBMQX, we perform two experiments 1 to compare the simulation model with the physical hard- 1 X D = pij p˜ij , (18) ware. In this way, the results of the simulated quantum 2 | − | i,j=0 processor can be directly compared to experimental re- sults for a device using the same pulse schemes to imple- between the ideal outcome distribution pij (i.e., the prob- ment quantum gates. ability to measure the state ij ) and the experimentally | i The first experiment again involves twenty successive measured relative frequenciesp ˜ij. The experiment on CNOT gates, but this time we measure the statistical the IBMQX was performed with 8192 shots on August 8

17, 2017 using Q3 as the control and Q4 as the target Gaussian microwave pulses to the system, with pulse pa- qubit. The results are presented in Fig. 3. rameters determined by an optimization routine. Hence, The simulated CR2 gate gives the best qualitative we are confident that they represent idealized versions of agreement between experiment and simulation. This the pulses used in recent experiments for this architec- makes sense because the CR2 pulse scheme shown in ture. This is confirmed by almost perfect results for the Fig. 1 is also used for the IBMQX [5]. Most remark- entanglement experiment [6]. Thus, our simulation can able is the fact that the performance for the initial state be seen as an ideal version of the experiment. Still, the 00 is much worse than for 10 (see also Table 3 in [6]). fact that all of our apparently ideal gates show diamond As| i this also shows up in the| ideali simulation, it points to error rates above 2%, suggests that the goal of building a a systematic error in the implementation of the CNOT universal, fault-tolerant quantum computer still remains gate. Note that this error is only present in the echoed a difficult, ongoing challenge. gates, and not in the proposed one-pulse gate CR1. The remaining difference between the CR2 simulation and the We have found that gate metrics such as the average experiment may be due to decoherence by the environ- gate fidelity [9], the diamond distance [10, 33], and the ment; we leave the study of including a heat bath in the unitarity [11] each provide insights into the errors of the simulation to future research. implemented gate pulses. Specifically, while the time evo- As a second experiment, we repeat the two-qubit en- lution of the total system is inherently unitary and the er- tanglement experiments proposed as part of a benchmark rors are systematic, they appear as incoherent non-Pauli for gate-based quantum computers [6]. The quantum cir- errors on the computational subspace. Conceptually, cuits first create the maximally entangled singlet state these errors originate from entanglement between the ( 01 10 )/√2 and then apply a set of rotations depen- computational states and the non-computational states | i − | i dent on the angles ϑ1 and ϑ2 to analyze the constructed in the transmons and the resonator. Such errors cause state (see Appendix C for the circuit). We select the two- most of the mismatch between the ideal gates and the pulse echoed CR2 gate for the CNOT operation, as also implemented pulses (see also [7, 11, 33, 47–49]). done for the IBMQX. We parse programs formulated in a quantum assembly language similar to the one used by However, the information obtained from the gate met- the IBMQX [25] to run the quantum circuits. The results rics is not enough to assess the error induced by repeat- are shown in Fig. 4. edly using the gate in quantum algorithms. To be precise, Although the gates used in the simulation, which are a gate showing close-to-ideal performance with respect to in some sense ideal versions of the gates used in experi- the studied gate metrics can still perform worse than an ments, do not reach perfect fidelities or error rates (see initially less ideal gate after multiple applications. In Table II), they still yield almost perfect results for the particular, the entangling two-qubit gates were found entanglement experiments. The results are much closer to lose performance over repeated applications. Espe- to those expected for a singlet state than the correspond- cially the two-pulse echoed CR2 gate exhibited system- ing experimental results on the IBMQX [6], even though atic errors that could also be observed in the IBMQX. the reported fidelities of the latter are the same or even In comparison, the longer CR4 gate seemed to perform better. This can have three reasons: (i) the procedure better in spite of worse fidelity, as also seen in recent of measuring the fidelities (i.e., randomized benchmark- experiments [28, 29]. When used in a QFT algorithm, ing) produces numbers that overestimate the gate per- however, this observation was reversed again. An ex- formance (cf. [33, 45, 46]), implying that the actual gate treme case was given by the one-pulse CNOT gate CR1, implementations are worse; (ii) the actual gates are good which for CNOT21 gave the best fidelity but the worst but the discrepancy is due to another process (such as the performance. In contrast, CNOT12 showed the worst measurement) that is not yet included in the simulation fidelity and the worst unitarity but a reasonably good model; or (iii) other unknown factors not included in the performance, without suffering from the systematic error quantum-theoretical description of the experiments play present in CR2 and CR4. Hence, the gate metrics un- a destructive role. der investigation do not provide reliable information of how well and how often a certain gate may be used in an algorithm (see also the conclusion in [50]). As this information is essential for potential users of gate-based V. DISCUSSION quantum computers, it should be included in the specifi- cation sheet of the physical device. We have implemented algorithms to solve the TDSE for a quantum-mechanical model of superconducting Future work will go into scaling up the simulation to transmon qubits coupled by transmission-line resonators. model experiments with more qubits and additional cou- The architecture of the publicly accessible quantum com- pling schemes, in accordance with the goal pursued in puter by IBM is of this type [5]. Great care has been experiments. This then enables a detailed simulation of taken to make no approximations to the Hamiltonian ob- quantum error-correcting codes under realistic conditions tained from the circuit quantization [20]. for various architectures. In addition to that, we plan to The tested quantum gates are realized by applying simulate the measurement process in detail. 9

ACKNOWLEDGMENTS stress that no further approximation needs to be made to obtain the solution of the TDSE. We are grateful to the IBM Quantum Experience The software is written in C++ and the implemen- project team for sharing technical details with us. This tation of the algorithms has been validated by compari- work does not reflect the views or opinions of IBM or any son with exact diagonalization for smaller Hilbert spaces. of its employees. Furthermore, we have checked that the results are quali- tatively independent of small variations in the time step τ, the number of charge and states included in Appendix A: Description of the algorithm the basis, and the particular device parameters given in Table I. The algorithm that we employ to solve the TDSE given by Eq. (4) is a Suzuki-Trotter product-formula algorithm constructed from the Hamiltonian by using the general Appendix B: Details about the gate pulses framework presented in [8]. The algorithm is explicit, inherently unitary, and unconditionally stable by con- The quantum gate set that we physically implement struction. Among others, the framework has been used reads to devise algorithms for NMR systems for quantum com- 1 2 1 2 putation [51], and it also forms the basis of the massively = Xπ/2,Xπ/2,Xπ,Xπ, CNOT12, CNOT21 , (B1) parallel quantum computer simulator [52] that can nowa- S { } days simulate systems with up to 45 qubits [53]. where Xj = exp( iϕσx/2) is a rotation of qubit j about The model Hamiltonian H given by Eq. (1) needs to be ϕ − j the x-axis by an angle of ϕ, and CNOTij is defined by expressed in an appropriate basis to derive the algorithm. negating the target qubit j if the control qubit i is in the In this work, we choose the charge basis n1n2 : ni Z {| i ∈ } state 1 and doing nothing if it is in the state 0 . We (i.e. the joint eigenbasis of the number operatorsn ˆ1 and | i j | i z additionally support the VZ gates Zϕ = exp( iϕσj /2) nˆ2) for the qubits and the Fock basis k : k N0 for − {| i ∈ } for arbitrary angles ϕ to make the gate set universal the resonator. This basis has the nice property that H for quantum computation [16]. By analogy withS experi- can be generically expressed as a sum of tensor products ments, a VZ gate does not correspond to a separate pulse, of tridiagonal matrices. but it changes the phases of all the following pulses (see At the heart of the algorithm lies a decomposition of [23]). For this purpose, we keep track of two offset phases the total unitary time-evolution operator φ1 and φ2 during the evolution, and the phase γ in Eq. (6) Z t0 ! of every subsequent pulse oscillating atω ¯1 (¯ω2) is shifted 0 U (t , t) = exp i dτH(τ) , (A1) by φ1 ( φ2). total − − T − t We also employ these zero-duration VZ gates to cor- rect phase errors in the gate sequence resulting from where is the time-ordering symbol. This expression T phase shifts due to other non-computational levels or is first discretized in time steps τ, i.e., we consider the off-resonant driving [23]. In particular, this means that propagator Ut+τ,t = exp( iτH(t + τ/2)). Note that τ each gate is followed by a local Z rotation of the form needs to be chosen small enough− with respect to the en- Zϕ Zϕ which essentially only results in an update of ergy scales and the other relevant time scales of H(t) 1 ⊗ 2 the tracked phases. The phases ϕ1 and ϕ2 are found by such that the exact mathematical solution of the TDSE an additional optimization step using the objective func- is obtained up to some fixed numerical precision. Sub- tion given by Eq. (9), but this time replacing the result sequently, the exponential of H(t + τ/2) is decomposed M of the first optimization by (Zϕ1 Zϕ2 )M. While this using the Lie-Trotter-Suzuki product-formula [54]. This does not change a single gate before⊗ the measurement, is done by partitioning the tridiagonal matrices into even and the optimized correction phases ϕ1 and ϕ2 are close and odd sums of 2 2 block-diagonal matrices such × to 0, we have observed that it considerably improves the that each matrix exponential can be evaluated analyt- performance of the gates after repeated applications be- ically (cf. [8, 51]). With these, we iteratively update P cause it mitigates the accumulation of phase errors. the state vector Ψ(t) = akn1n2 (t) k n1 n2 us- ing the second-order| expressioni of the framework.| i| i| i Fi- nally, the solution is transformed to the transmon ba- 1. Single-qubit gates sis m1m2 : mi N0 (see Eq. (5)) by computing {| i P ∈ 1 } ∗ 2 ∗ akm1m2 (t) = n1n2 (Bn1m1 ) (Bn2m2 ) akn1n2 (t), where i P i The general pulse for a single qubit gate is given the Bnimi are defined by mi = ni Bnimi ni and ob- | i | i by Eq. (8) and depends on the parameters (Ω0, β, γ). tained from the eigenvectors of HCPB given by Eq. (2) The amplitude Ω0 is directly proportional to the im- for ngi(t) = 0. In practice, we set the time step to solve the TDSE plemented angle of rotation ϑ π/2, π and can be ∈ { R T } to τ = 0.1 ps and the number of states included in the obtained from the relation ϑ = bi 0 ΩG(t) dt where 1/4 product basis to ni = 8,..., 8 and k = 0,..., 3. We bi = 2ECi(EJi/8ECi) [22]. The DRAG coefficient β − 10

(a) CNOT12 TABLE IV. Parameters defining the single-qubit pulses as ob- 0 GHz 0.8 GHz 1.6 GHz -0.16 GHz 0 GHz 0.16 GHz tained by the pulse optimization with the initial values taken 40 80 30 IX 60 IX from the theory of transmon qubit control [22]. ZX 40 ZX 20 IX (theo) ZX (theo) 20 10 0 Pulse Ω0 β in ns ϕ1 ϕ2 0 20 − GD1 0.00222 0.231 −0.00202 0.00328 10 40 π/2 − −60 2 20 − GD 0.00227 0.289 −0.00013 −0.00159 − 80 π/2 Int. Strength [MHz] 30 Int. Strength [MHz] −100 1 − 0 0.05 0.1 − 0.01 0 0.01 GDπ 0.00444 0.219 −0.00354 0.00283 − Amplitude Amplitude GD2 0.00454 0.224 −0.00026 −0.00339 ΩCR ΩCancel π (b) CNOT21 0 GHz 0.8 GHz 1.6 GHz -0.16 GHz 0 GHz 0.16 GHz 0 80 1 60 IX 40 ZX TABLE V. Parameters defining the two-qubit pulses, result- −2 − 20 ing from the pulse optimization procedure. The parameters 3 0 − ϕCR,ΩCancel, and ϕCancel are only needed for the CR1 scheme. 4 20 − IX −40 5 ZX − − IX (theo) 60 Pulse TCR in ns ΩCR ϕCR ΩCancel ϕCancel ϕ1 ϕ2 6 − − ZX (theo) 80 Int. Strength [MHz] 1 7 Int. Strength [MHz] −100 GFCR1 41.86 0.079 0.54 0.0062 0.00 −2.10 0.04 − 0 0.05 0.1 − 0.01 0 0.01 2 − GFCR1 128.19 0.094 −2.89 −0.0016 1.72 3.25 1.40 Amplitude ΩCR Amplitude ΩCancel 1 GFCR2 102.97 0.011 − − − 0.00 0.00 2 FIG. 5. (Color online) Scan of the CR drive amplitudes ΩCR GFCR2 71.56 0.071 − − − 0.00 0.00 on the control qubit and Ω on the target qubit. The GF1 50.24 0.010 − − − 0.00 −0.01 Cancel CR4 dimensionless amplitudes can be converted to the strength of GF2 30.16 0.069 − − − −0.01 0.00 1/4 CR4 the drive by multiplying them with bi = 2ECi(EJi/8ECi) (shown on top of the plots). The IX and ZX interaction strengths are inferred by measuring the oscillations of the target qubit conditional on the control qubit being in state is initially set to 1/2αi (see [26]), where the anhar- − |0i and state |1i [13]. The linear theory predictions can be de- monicity αi is given in Table I. These two parameters rived perturbatively [22] and are only valid for weak drivings. are refined in the optimization procedure as described in Note that the additional drive on the target qubit (the two Section II B. figures on the right, shown for ΩCR = 0.1 fixed) linearly dis- In the following, we denote the resulting single-qubit places IX only. Thus, it can either be tuned to single out ZX i i pulses on qubit i by GDπ/2(γ) and GDπ(γ). The or to generate the CNOT gate directly up to local Z rotations. corresponding pulse parameters along with the above- mentioned VZ phase corrections ϕ1 and ϕ2 are given in Table IV. The only parameter left in these pulses is the phase γ. This phase is used to implement VZ gates ac- cording to the scheme [23]

i i tial goal of CR gates was to single out a ZX interac- GD (γ) Zϕ ψ = Zϕ GD (γ ϕ) ψ . (B2) ϑ | i ϑ − | i tion [19], the CR2 and CR4 gates use an echo scheme to echo out the IX interaction. The one-pulse gate CR1, 2. Two-qubit gates in contrast, uses the additional drive on the target qubit to shift IX such that the implemented transformation is exp( iπ(3σx +σz σx )/4), which is equal to a CNOT gate The central pulse in two-qubit gates for the present − T C T up to local Z rotations. The correct time TCR for each architecture is the cross-resonance (CR) pulse, depicted pulse is obtained from a separate scan. as a flat-topped Gaussian in Fig. 1. It always oscillates at the frequency of the target qubitω ¯T and it is defined by its amplitude ΩCR and the time TCR of the flat top The final pulse parameters are then found in the pulse (thus the time of the CR drive including rise and fall is optimization procedure (see Section II B). By analogy TCR + 30 ns). The CR1 scheme additionally includes the with the single-qubit pulses, we denote the flat-topped i amplitude of the target drive ΩCancel and two phases ϕCR CR drivings on qubit i by GFCR∗(γ). The correspond- and ϕCancel, inspired by the observations in [13]. ing parameters and the VZ phase corrections are given Although there are theoretical predictions based on in Table V. Again, γ is the only variable parameter, and perturbation theory for the specific choice of parameters it can be used to implement VZ gates in the same way [13, 22], they need to be fine-tuned to the specific set as in Eq. (B2). Note that, as the CNOT gate commutes of qubits. We do this by scanning the amplitudes for with Z gates on the control qubit, only phase shifts of a CR drive and obtaining the CR interaction strengths the target qubit affect γ [23]. The full specifications in- from the conditional rotation of the target qubit, as done cluding the scheme to implement VZ gates are given in in [13]. Such a scan is shown in Fig. 5. As the ini- Fig. 6. 11

(a) CNOT (b) CR1 (c) CR2

GDC GFC GDC GFC C ZϑC • C ZϑC π CR2 π CR2 Z C (0) (−ϑ ) (π/2) (π−ϑ ) ϑC +ξ−π/2 C T T GFCR1(−ϑT ) GDT Z Z π/2 T ϑT T ϑT T ZϑT (ξ−ϑT )

(d) CR4

C C C C C C GFCR4 GDπ GFCR4 GFCR4 GDπ GFCR4 C ZϑC +ξ−π/2 (−ϑT ) (π/2) (π−ϑT ) (π−ϑT ) (3π/2) (−ϑT )

GDT T π/2 GDπ T ZϑT (ξ−ϑT ) (π+ξ−ϑT )

FIG. 6. Specifications of the pulse sequences in Fig. 1 to implement a generic CNOT gate with VZ phases. The elementary Gaussian pulse GD is defined in Table IV, and GF is defined in Table V. C (T) is the control (target) qubit. (a) Generic CNOT gate with the preceding VZ phases that all pulses need to be capable of shifting through; (b) one-pulse CR1 gate which includes a flat-topped Gaussian pulse on the control and the target qubit simultaneously; (c) two-pulse echoed CR2 gate; (d) four-pulse echoed CR4 gate. In the CR2 and the CR4 scheme, the additional phase shift ξ is 0 ifω ¯C > ω¯T and π otherwise to handle the case when ZX is negative (see Fig. 5).

Appendix C: Circuits for the quantum programs † † |m1i H T T S • •

|m2i • • T H • In the following, we show the quantum circuits for the QFT algorithm and the entanglement experiments from Section IV. The gates contained in the circuits map to the FIG. 7. Circuit for the two-qubit QFT. pulses defined in Appendix B in the same way as for the IBMQX [25]. In particular, we have H = Zπ/2Xπ/2Zπ/2, † S = Zπ/2, T = Zπ/4, T = Z−π/4, and U1(ϑ) = Zϑ (up to global phases). The two-qubit QFT in principle contains two |0i X H • H U1(ϑ1) H Hadamard gates H, one controlled-S gate, and one

|0i X H U1(ϑ2) H SWAP gate [16]. Rewriting this in terms of the gates supported by our system leads to the circuit given in Fig. 7. As only the H gate and the CNOT gate result in FIG. 8. Circuit for experiments on the singlet state. actual hardware pulses, this circuit involves two X pulses and five CNOT pulses in total. The circuit to analyze the singlet state as a function of the angles ϑ1 and ϑ2 is taken directly from [6] and is given in Fig. 8.

[1] J. M. Gambetta, J. M. Chow, and M. Steffen, npj Quan- wick, M. Block, B. Bloom, A. Bradley, C. Bui, S. Cald- tum Information 3, 2 (2017). well, L. Capelluto, R. Chilcott, J. Cordova, G. Crossman, [2] J. Kelly, R. Barends, A. G. Fowler, A. Megrant, E. Jef- M. Curtis, S. Deshpande, T. E. Bouayadi, D. Girshovich, frey, T. C. White, D. Sank, J. Y. Mutus, B. Campbell, S. Hong, A. Hudson, P. Karalekas, K. Kuang, M. Leni- Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, I.-C. Hoi, han, R. Manenti, T. Manning, J. Marshall, Y. Mohan, C. Neill, P. J. J. O’Malley, C. Quintana, P. Roushan, W. O’Brien, J. Otterbach, A. Papageorge, J. P. Paque- A. Vainsencher, J. Wenner, A. N. Cleland, and J. M. tte, M. Pelstring, A. Polloreno, V. Rawat, C. A. Ryan, Martinis, Nature 519, 66 (2015). R. Renzas, N. Rubin, D. Russell, M. Rust, D. Scarabelli, [3] C. Song, K. Xu, W. Liu, C. Yang, S.-B. Zheng, H. Deng, M. Selvanayagam, R. Sinclair, R. Smith, M. Suska, T. W. Q. Xie, K. Huang, Q. Guo, L. Zhang, P. Zhang, D. Xu, To, M. Vahidpour, N. Vodrahalli, T. Whyland, K. Yadav, D. Zheng, X. Zhu, H. Wang, Y. A. Chen, C. Y. Lu, W. Zeng, and C. T. Rigetti, (2017), arXiv:1706.06570. S. Han, and J. W. Pan, (2017), arXiv:1703.10302. [5] IBM, “Quantum experience,” https://www.research. [4] M. Reagor, C. B. Osborn, N. Tezak, A. Staley, ibm.com/ibm-q/ (2016). G. Prawiroatmodjo, M. Scheer, N. Alidoust, E. A. Sete, [6] K. Michielsen, M. Nocon, D. Willsch, F. Jin, Th. Lippert, N. Didier, M. P. da Silva, E. Acala, J. Angeles, A. Best- and H. De Raedt, Comp. Phys. Comm. 220, 44 (2017). 12

[7] D. Puzzuoli, C. Granade, H. Haas, B. Criger, E. Mage- [31] W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P. san, and D. G. Cory, Phys. Rev. A 89, 022306 (2014). Flannery, Numerical Recipes 3rd Edition: The Art of [8] H. De Raedt, Comp. Phys. Rep. 7, 1 (1987). Scientific Computing (Cambridge University Press, New [9] M. A. Nielsen, Physics Letters A 303, 249 (2002). York, USA, 2007). [10] A. Y. Kitaev, Russian Mathematical Surveys 52, 1191 [32] J. J. Wallman, (2015), arXiv:1511.00727. (1997). [33] Y. R. Sanders, J. J. Wallman, and B. C. Sanders, New [11] J. Wallman, C. Granade, R. Harper, and S. T. Flammia, J. Phys. 18, 012002 (2016). New J. Phys. 17, 113020 (2015). [34] J. M. Chow, J. M. Gambetta, A. D. C´orcoles, S. T. [12] S. Sheldon, L. S. Bishop, E. Magesan, S. Filipp, J. M. Merkel, J. A. Smolin, C. Rigetti, S. Poletto, G. A. Keefe, Chow, and J. M. Gambetta, Phys. Rev. A 93, 012301 M. B. Rothwell, J. R. Rozen, M. B. Ketchen, and (2016). M. Steffen, Phys. Rev. Lett. 109, 060501 (2012). [13] S. Sheldon, E. Magesan, J. M. Chow, and J. M. Gam- [35] H. Breuer and F. Petruccione, The Theory of Open Quan- betta, Phys. Rev. A 93, 060302 (2016). tum Systems (OUP Oxford, 2007). [14] R. Barends, J. Kelly, A. Megrant, A. Veitia, [36] J. Emerson, R. Alicki, and K. yczkowski, J. Opt. B: D. Sank, E. Jeffrey, T. C. White, J. Mutus, A. G. Quantum Semiclassical Opt. 7, S347 (2005). Fowler, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, [37] E. Magesan, J. M. Gambetta, and J. Emerson, Phys. A. Dunsworth, C. Neill, P. O’Malley, P. Roushan, Rev. A 85, 042311 (2012). A. Vainsencher, J. Wenner, A. N. Korotkov, A. N. Cle- [38] C. J. Wood and J. M. Gambetta, (2017), land, and J. M. Martinis, Nature 508, 500 (2014). arXiv:1704.03081. [15] J. Kelly, R. Barends, B. Campbell, Y. Chen, Z. Chen, [39] D. Aharonov and M. Ben-Or, SIAM Journal on Comput- B. Chiaro, A. Dunsworth, A. G. Fowler, I.-C. Hoi, E. Jef- ing 38, 1207 (2008). frey, A. Megrant, J. Mutus, C. Neill, P. J. J. O’Malley, [40] N. Johnston, D. W. Kribs, and V. I. Paulsen, Quantum C. Quintana, P. Roushan, D. Sank, A. Vainsencher, Info. Comput. 9, 16 (2009). J. Wenner, T. C. White, A. N. Cleland, and J. M. Mar- [41] G. Golub and C. Van Loan, Matrix Computations, Johns tinis, Phys. Rev. Lett. 112, 240504 (2014). Hopkins Studies in the Mathematical Sciences (Johns [16] M. A. Nielsen and I. L. Chuang, Quantum Computation Hopkins University Press, 1996). and Quantum Information: 10th Anniversary Edition [42] Z. Chen, J. Kelly, C. Quintana, R. Barends, B. Camp- (Cambridge University Press, New York, USA, 2011). bell, Y. Chen, B. Chiaro, A. Dunsworth, A. G. Fowler, [17] J. M. Chow, A. D. C´orcoles,J. M. Gambetta, C. Rigetti, E. Lucero, E. Jeffrey, A. Megrant, J. Mutus, M. Nee- B. R. Johnson, J. A. Smolin, J. R. Rozen, G. A. Keefe, ley, C. Neill, P. J. J. O’Malley, P. Roushan, D. Sank, M. B. Rothwell, M. B. Ketchen, and M. Steffen, Phys. A. Vainsencher, J. Wenner, T. C. White, A. N. Korotkov, Rev. Lett. 107, 080502 (2011). and J. M. Martinis, Phys. Rev. Lett. 116, 020501 (2016). [18] C. Rigetti and M. Devoret, Phys. Rev. B 81, 134507 [43] P.-M. Billangeon, J. S. Tsai, and Y. Nakamura, Phys. (2010). Rev. B 91, 094517 (2015). [19] A. D. C´orcoles, J. M. Gambetta, J. M. Chow, J. A. [44] S. Richer and D. DiVincenzo, Phys. Rev. B 93, 134501 Smolin, M. Ware, J. Strand, B. L. T. Plourde, and (2016). M. Steffen, Phys. Rev. A 87, 030301 (2013). [45] T. Proctor, K. Rudinger, K. Young, M. Sarovar, and [20] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I. R. Blume-Kohout, (2017), arXiv:1702.01853. Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, [46] J. J. Wallman, (2017), arXiv:1703.09835. and R. J. Schoelkopf, Phys. Rev. A 76, 042319 (2007). [47] R. Kueng, D. M. Long, A. C. Doherty, and S. T. Flam- [21] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and mia, Phys. Rev. Lett. 117, 170502 (2016). R. J. Schoelkopf, Phys. Rev. A 69, 062320 (2004). [48] J. Ghosh, A. G. Fowler, J. M. Martinis, and M. R. Geller, [22] J. M. Gambetta, “Control of superconducting qubits,” Phys. Rev. A 88, 062329 (2013). (2013), Lecture Notes of the 44th IFF Spring School. Ed. [49] E. Magesan, D. Puzzuoli, C. E. Granade, and D. G. by David DiVincenzo. Schriften des Forschungszentrums Cory, Phys. Rev. A 87, 012324 (2013). Juelich. [50] A. Darmawan, P. Iyer, and D. Poulin, “Critical noise [23] D. C. McKay, C. J. Wood, S. Sheldon, J. M. Chow, and parameters for fault-tolerant quantum computation,” J. M. Gambetta, (2016), arXiv:1612.00858. https://www.physique.usherbrooke.ca/~dpoulin/ [24] F. Motzoi, J. M. Gambetta, P. Rebentrost, and F. K. utilisateur/files/seminaires/2016_Yale.pdf Wilhelm, Phys. Rev. Lett. 103, 110501 (2009). (2016). [25] A. W. Cross, L. S. Bishop, J. A. Smolin, and J. M. [51] H. De Raedt and K. Michielsen, Handbook of Theoretical Gambetta, (2017), arXiv:1707.03429. and Computational Nanotechnology, Vol. 3: Quantum [26] J. M. Gambetta, F. Motzoi, S. T. Merkel, and F. K. and molecular computing, quantum simulations, Chapter Wilhelm, Phys. Rev. A 83, 012308 (2011). 1, pp. 1-48, M. Rieth and W. Schommers eds., American [27] A. C´orcoles,E. Magesan, S. J. Srinivasan, A. W. Cross, Scientific Publisher, Los Angeles (2006). M. Steffen, J. M. Gambetta, and J. M. Chow, Nat. Com- [52] K. De Raedt, K. Michielsen, H. De Raedt, B. Trieu, mun. 6, 6979 (2015). G. Arnold, M. Richter, Th. Lippert, H. Watanabe, and [28] M. Takita, A. D. C´orcoles, E. Magesan, B. Abdo, N. Ito, Comp. Phys. Comm. 176, 121 (2007). M. Brink, A. Cross, J. M. Chow, and J. M. Gambetta, [53] N. Yoshioka, H. Watanabe, N. Ito, F. Jin, K. Michielsen, Phys. Rev. Lett. 117, 210505 (2016). and H. De Raedt, “Massively parallel quantum computer [29] M. Takita, A. W. Cross, A. D. Crcoles, J. M. Chow, and simulator, ten years later,” In preparation. J. M. Gambetta, (2017), arXiv:1705.09259. [54] M. Suzuki, S. Miyashita, and A. Kuroda, Prog. Theor. [30] J. A. Nelder and R. Mead, The Computer Journal 7, 308 Phys. 58, 1377 (1977). (1965).