<<

Lindblad Tomography of a Superconducting Quantum Processor

1, 2, 1 3, 4, 4 Gabriel O. Samach, ∗ Ami Greene, Johannes Borregaard, † Matthias Christandl, David K. Kim,2 Christopher M. McNally,1 Alexander Melville,2 Bethany M. Niedzielski,2 Youngkyu Sung,1 Danna Rosenberg,2 Mollie E. Schwartz,2 Jonilyn L. Yoder,2 Terry P. Orlando,1 Joel I-Jan Wang,1 Simon Gustavsson,1 Morten Kjaergaard,1, 5 and William D. Oliver1, 2, 6 1Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA 02139 2MIT Lincoln Laboratory, Lexington, MA 02421 3Qutech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands 4Department of Mathematical Sciences, University of Copenhagen, Copenhagen, Denmark 5Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, Copenhagen, Denmark 6Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA 02139 (Dated: June 15, 2021) As progress is made towards the first generation of error-corrected quantum computers, careful characterization of a processor’s noise environment will be crucial to designing tailored, low-overhead error correction protocols. While standard coherence metrics and characterization protocols such as T1 and T2, process tomography, and randomized benchmarking are now ubiquitous, these tech- niques provide only partial information about the dynamic multi-qubit loss channels responsible for processor errors, which can be described more fully by a Lindblad operator in the master equation formalism. Here, we introduce and experimentally demonstrate Lindblad Tomography, a hardware- agnostic characterization protocol for tomographically reconstructing the Hamiltonian and Lindblad operators of a from an ensemble of time-domain measurements. Performing Lind- blad Tomography on a small superconducting quantum processor, we show that this technique characterizes and accounts for state-preparation and measurement (SPAM) errors and allows one to place strong bounds on the degree of non-Markovianity in the channels of interest. Comparing the results of single- and two-qubit measurements on a superconducting quantum processor, we demon- strate that Lindblad Tomography can also be used to identify and quantify sources of crosstalk on quantum processors, such as the presence of always-on qubit-qubit interactions.

I. INTRODUCTION term, the need for redundancy can be substantially re- duced by tailoring the correction scheme to the specific Quantum computers offer computational power fun- noise environment and imperfections of the particular damentally distinct from that of their classical counter- quantum processor [7]. parts and are predicted to offer an advantage for certain In order to develop such tailored error-correction pro- problems in fields such as quantum chemistry and opti- tocols, we require robust measurement protocols for mization, which are often intractable on even the largest characterizing the performance and loss mechanisms of classical supercomputers [1, 2]. The promise of quan- real quantum devices. To this end, many techniques tum advantage over classical hardware has driven exten- have been proposed and utilized which focus on dif- sive efforts to build devices based ferent aspects of device performance, such as random- on a number of different hardware platforms—including ized benchmarking [8, 9], gate set tomography [10] and trapped ions [3], neutral atoms [4], and superconducting state/process tomography [11], each with their own circuits [5]—each of which is susceptible to characteris- strengths and weaknesses [12]. tic imperfections and noise mechanisms which can limit Randomized benchmarking, for example, provides an arXiv:2105.02338v2 [quant-ph] 13 Jun 2021 performance. approach for assessing the average fidelity of quantum To mitigate these sources of error, fault-tolerant quan- gate operations independently of state preparation and tum error correction protocols encode logical qubits measurement (SPAM) errors, and it has consequently be- across many physical qubits, provided the error rate of come a standard measure of performance for experimen- the physical qubits is below a threshold. This approach, tal quantum devices. However, the average fidelity alone however, comes with considerable overhead in terms of does not provide much information about the actual noise the additional qubits needed for the encoding [6]. While processes at play in the device, the details of which are the overhead required for generic device-agnostic error crucial to more fully modeling the device and developing correction schemes may prove prohibitive in the near- tailored error mitigation and correction techniques. State and process tomography, on the other hand, pro- vide more detailed information about discrete moments ∗ Corresponding author: [email protected] in a qubit’s evolution, such as the qubit state at a par- † Corresponding author: [email protected] ticular time or the quantum process corresponding to a 2 gate of a particular duration. However, caution must be (a) Initialize 6 Sweep Quantum Measure in 3 exercised in order to consistently interpret the results of Cardinal States Channel of Interest Pauli Bases tomography in the presence of SPAM errors [13]. Build- ing on traditional tomographic protocols, a number of theoretical and experimental works have demonstrated self-consistent characterization of SPAM errors in process tomography and gate characterization [13–16]. Common to many of these techniques is the use of maximum like- Idling Channel lihood estimation (MLE), which provides a robust and Measurement flexible estimation procedure capable of handling over- (Ramsey) Measurement complete data and constrained problems. While such techniques offer a promising step forward, the character- (b) Classical Optimizer ization of a discrete moment in a qubit’s evolution is not Maximum Likelihood Estimation (MLE) always sufficient, and one often requires detailed knowl- edge about how the noise environment and crosstalk be- Extract SPAM Errors tween qubits vary dynamically in time [17]. Imperfect Ground State POVM Here, we present a new technique for characterizing the dynamics of a multi-qubit system from an ensemble of time-domain measurements, which we call Lindblad Tomography (LT). As a characterization tool, LT can be Reconstruct Discrete Estimate Dynamic used to analyze any general noisy process governing the Kraus Operators Hamiltonian and Lindblad evolution of a quantum system in time. This could be the noise processes experienced by a qubit during free evolution—such as T1- and T2-processes, which can be Calculate Channel Markovianity formally described as amplitude damping and dephas- Independently check Markovianity ing channels respectively—but it could also refer to a by comparing operator fit to data sequence of intentional operations, such as a series of im- perfect gates or a noisy quantum algorithm. FIG. 1. Single-qubit protocol for Lindblad Tomography (LT) The goal of LT is to estimate the Hamiltonian, quan- of the Idling Channel 1˜(t). (a) The sequence of measure- tum jump operators, and corresponding decay rates that ments required for single-qubit LT. The qubit is prepared in describe the evolution of interest using maximum likeli- its imperfect ground state ρ0 and one of six single-qubit pre- hood estimation (MLE), a process we collectively refer pulses Rs is applied to rotate the qubit close to each cardinal to as extracting the of the channel. In doing state of the Bloch sphere; the quantum channel of interest is so, we assume the channel can be well-approximated by swept; and one of three post-pulses Rb is applied to rotate a time-independent master equation. Prior to extracting the measurement basis. (b) Analysis protocol for LT. Re- the Lindbladian of the channel, our protocol uses a sub- sults from all combinations of pre-/post-pulses and channel durations are passed to a classical optimizer based on maxi- set of measurement data to characterize the SPAM errors mum likelihood estimation (MLE). SPAM errors due to im- for the device, which we then include in our estimation of perfect ground state preparation and measurement infidelity the Lindbladian from the full set of measurement data. are extracted from data at t = 0, and the results are used In doing so, we assume that the SPAM errors are con- to separately estimate: (left path) the Kraus operators (ti) K stant across the full set of LT measurements, and thus are for each discrete channel duration ti and channel Markovian- time-independent during the duration of data collection. ity using the distance measure; (right path) the Hamiltonian The main requirements for LT are thus: Hˆ and Lindblad matrix L for continuous time t, where a poor fit serves as an independent indicator of channel non- 1. The evolution of the quantum system should Markovianity. be Markovian and well described by a time- independent master equation.

2. SPAM errors are assumed to be constant during tors of the channel at discrete times, which can then be the full duration of data acquisition. used to validate the assumption of Markovianity using In addition to these two fundamental assumptions, we the measure proposed in Ref. [18]. Performing LT on a note that further simplifications may be made based superconducting quantum processor, we show that this on specific hardware considerations—such as a third as- verification technique enables us to identify sources of sumption of perfect single-qubit gates for LT of high fi- non-Markovianity which arise due to always-on crosstalk delity superconducting qubits, discussed further below— between neighboring qubits. though such assumptions are not required for performing While estimation techniques for Lindblad noise oper- LT in general. ators have been proposed and demonstrated previously As we demonstrate, the set of measurements required for a single-qubit solid-state [19] and trapped-ion sys- for LT can also be used to extract the Kraus opera- tem [20], LT differs from past methods in its careful ac- 3 count of SPAM errors during the estimation, its ability measurement protocol as a hybrid of standard qubit char- to place bounds on the Markovianity of the channel, and acterization techniques, combining time-domain charac- its use of MLE. terization of T1/T2 with tomography. By iterating over The paper is organized as follows. In Section II, we the full set of pre- and post-rotations, LT of the Idling introduce the general technical framework behind Lind- Channel effectively pieces together all combinations of blad Tomography. Applying LT to a small superconduct- T1- and T2-like measurements to tomographically recon- ing quantum processor, we then continue with a charac- struct the full quantum loss channel. terization of SPAM errors for this device in Section III, estimation of Kraus operators and the degree of Marko- vianity in Sections IV and V, and finally Hamiltonian III. EXTRACTING SPAM ERRORS and Lindblad estimation in Section VI. Once we have collected the full set of data for the channel of interest, the first step of analysis is to char- II. SINGLE-QUBIT LINDBLAD TOMOGRAPHY MEASUREMENT PROTOCOL acterize the state-preparation errors of the device using MLE. We parameterize the imperfect initial ground state of the qubit as an arbitrary single qubit , We first introduce Lindblad Tomography in the con- ρ0, which is constrained by physicality conditions to be text of characterizing a single qubit. The generalization positive semi-definite and have unit trace. The condition to two or more qubits follows readily, as discussed below. of positivity is enforced in the optimization by expressing The structure of single-qubit LT is illustrated in Fig. 1. ρ0 as a Cholesky decomposition ρ0 = AA† and estimating To determine the Lindbladian describing a single-qubit the elements of the unconstrained matrix A from which channel, we perform the following over-complete set of ρ0 is computed. The unit trace condition is readily in- single-qubit rotations and basis measurements (Fig. 1a): cluded by normalization.

1. The qubit is initialized in a state ρ0 close to its Next, we characterize the measurement apparatus by ground state and one of six single-qubit gates Rs = extracting the positive operator-valued measure (POVM) 1, X , Y π , X π is applied, initializing the qubit describing the measurement. For most qubit modalities, π 2 2 as{ close as± possible∓ } to each of the six cardinal states measurements are natively performed in a fixed z-basis, of the Bloch sphere ( 0 , 1 , , i ) respectively. while measurement in other bases are performed by rotat- | i | i |±i |± i ing the state prior to measurement. Conveniently, these 2. A quantum channel of interest is swept as a func- rotations are a subset of the pulses needed to prepare the tion of time. In this work, we consider the Idling initial basis states of the Pauli matrices, and they thus do Channel 1˜(t), corresponding to a variable time de- not require separate characterization. In general, these lay between state preparation and measurement gates cannot be assumed error free and should therefore during which no experimental controls are per- be parameterized as arbitrary rotation matrices and es- formed on the qubit. In the absence of any noise, timated together with the POVM and initial state pa- the Idling Channel would correspond to the Iden- rameters in the LT sequence, similar to gate set tomog- tity Channel 1(t). raphy [16]. However, we note that randomized bench- marking, which is not influenced by SPAM errors, can 3. One of three single-qubit gates R = 1, Y π , X π b 2 2 is applied prior to measurement, corresponding{ − to} be performed prior to LT to obtain a good estimate of measurement in the Pauli z-, x-, and y-bases. the rotation pulse errors, which can simplify the problem and improve estimation results. These steps are repeated for all combinations of initial For high fidelity superconducting qubits, single-qubit state, channel duration, and measurement basis, and the operations are typically orders of magnitude less prone results are saved in classical memory for analysis. This to error (typical error rates < 0.05%) than measurements dataset is fed to an MLE routine to determine the el- (typical error rates 1%) [21]. In such systems, it is there- ements of ρ0 and the positive operator-valued measure fore reasonable to assume that errors in single-qubit ro- (POVM) representing the measurement (Fig. 1b blue tations have negligible impact on state-initialization and bubble, details in Sec. III). This initial characterization is POVM estimation in comparison to imperfect thermal- then used to determine the process map at each discrete ization and measurement error. In the device used for the time (Fig. 1b orange bubble and Sec. IV) or to extract present experiment, we find measurement fidelities on the the Hamiltonian and Lindblad operators for continuous order of 99%, in comparison to single-qubit gate fidelities time (Fig. 1b red bubble and Sec. VI). 99.99% measured using interleaved randomized bench- In this first implementation of Lindblad Tomography, marking∼ [22]. As such, for the following characterization we consider free evolution corresponding to applying the of this particular device, we will assume perfect rotation Idling Channel 1˜(t). For this channel, a subset of the pulses, requiring only estimation of the z-basis POVM LT measurement sequences are identical to conventional and initial state. T1 (pale green gates in Fig. 1a) and T2 (purple) mea- The single-qubit POVM corresponding to measure- surements. In this way, one can helpfully think of this ment in the z-basis has two operator elements (2 2 × 4

1 matrices) M0,M1 , where M0 + M1 = . The proba- (a) bility of measuring{ } a state ρ in the ground state is then p0 = Tr[ρM0] and the excited state is p1 = Tr[ρM1]. To estimate the POVM, we optimize over the matrix ele- ments of M0, subject to the constraint that both M0 and M = 1 M are positive semidefinite. 1 − 0 In order to find the initial state ρ0 and POVM M0 that best describes the measurements, we construct a maxi- (b) mum likelihood function SPAM for our SPAM errors, which allows us to optimizeL over the unknown elements of these matrices. To perform a log-likelihood estimation, we take the logarithm of this function

ln( ) = f(s, b) ln Tr [ρ M ] LSPAM s b Xb,s   + f¯(s, b) ln Tr [ρ (1 M )] (1) s − b   where b z, x, y runs over the measurement bases, (c) (d) s 0 ,∈1 {, + , } , +i , i runs over the imper- fect∈ input{| i | states,i | i and|−i f|(s,i b)(|−f¯i}(s, b)) is the total num- ber of ‘0’s (‘1’s) recorded during repeated measurements of state ρs in measurement basis b. In general, the matrices Mb = RbM0Rb† would require estimating the matrix elements of both M0 as well as the potentially faulty rotations (Rb = 1, Y π , X π ), and the matrices { − 2 2 } (e) (f) ρs = Rsρ0Rs† would require estimating the matrix ele- ments of both the initial state (ρ0) and the potentially faulty rotations (R = 1, X , Y π , X π ). However, as s π 2 2 detailed above, we assume{ perfect± rotations∓ } in the char- acterization of our superconducting device, so only the elements of M0 and ρ0 remain to be found. For LT of the Idling Channel, the experimental data required for estimating SPAM errors corresponds to the subset of data taken for the zero-duration channel 1˜(t= 0), where we prepare the qubit in each of the states ρs FIG. 2. Initial characterization of SPAM errors as part of and measure immediately in each basis b. Qubit measure- Lindblad Tomography. (a) Schematic of the two coupled ments are recorded as single-shots, and the outcomes are transmon qubits used in this experiment. (b) Skyscraper AB labeled as either ‘0’ or ‘1’ using standard superconducting plots of the imperfect two-qubit ground state ρ0 extracted qubit measurement techniques [23]. The unknown matrix during two-qubit LT (negative values shown in pale green, ideal ground state shown in wireframe, elements smaller than elements of the initial state and the measurement POVM 2 10− omitted for visual clarity). (c–f) Extracted two-qubit are then estimated by maximizing the log-likelihood in , corresponding to imperfect measurement of the Eq. (1) with respect to these elements. states 00 , 01 , 10 , and 11 respectively (negative values This technique extends naturally to multi-qubit sys- shown| in red,i | perfecti | i POVMs| showni in wireframe, imaginary 2 tems. For two qubits A and B, we characterize the initial parts and elements smaller than 10− omitted for visual clar- AB state as a general two-qubit density matrix ρ0 , and we ity). The full single- and two-qubit matrices for the extracted characterize measurement using four 4 4 POVM ma- initial states and POVMs are included in the supplement. × trices M00,M01,M10,M11 , corresponding to measure- ment of{ the states 00 , 01}, 10 , and 11 respectively. To determine the matrix| i | elementsi | i of the| initiali state and periment as qubit A and B respectively (Fig. 2a, full de- the POVMs, we maximize a log-likelihood function anal- vice characterization found in Ref. [22, 24]). In Fig. 2b, ogous to Eq. 1 containing four terms (corresponding to we plot the extracted matrix elements of the imper- AB measurement of each of the four two-qubit computational fect two-qubit ground states ρ0 , extracted during two- states) and summing over the full set of two-qubit pre- qubit LT. In Fig. 2c–f, we plot the elements of the two- and post-pulses (discussed further in Sec. IV). qubit POVM matrices, corresponding to measurement The experiment is run on a device consisting of of the four two-qubit computational states respectively. three capacitively-coupled flux-tunable transmon qubits, Bold wireframes in Fig. 2b–f highlight ideal ground state where we denote the two qubits characterized in this ex- preparation and perfect z-basis POVM matrices for com- 5 parison. IV. RECONSTRUCTING THE KRAUS We note that the initial single-qubit states are very OPERATORS AT DISCRETE TIMES similar to a thermal state of the form ρthermal = a 0 0 + (1 a) 1 1 , a [0, 1]. Minimizing the | ih | − | ih | ∈ Once we have performed an initial characterization D(ρthermal, ρ0) = ρthermal ρ0 /2 ( M = Tr √M †M ) of the SPAM errors for our device, we can proceed to | − | | | characterize the quantum channel of interest. For su- between the estimated initial states and a thermalh state,i with respect to the thermal parameter a, we find minimal perconducting devices, qubits are primarily subject to A B amplitude damping and dephasing noise over time [23], trace distances D(ρthermal, ρ0 ) = 0.01 (D(ρthermal, ρ0 ) = 0.02) for a = 0.86 (a = 0.88). This result is thus consis- which are traditionally characterized with simple T1 and tent with the observation that the imperfect ground state T2 measurements respectively. However, to accurately of a superconducting qubit at finite temperature can be model multi-qubit devices or develop tailored error cor- approximated as a thermal state of an anharmonic oscil- rection techniques [25], it is important to capture the lator [23]. interplay between these channels as well as how they de- Traditionally, the SPAM errors for a single qubit would pend on the state of neighboring qubits, details which are be found by characterizing the qubit while all nearby not readily obtainable from standard T1 and T2 measure- qubits are left in their ground states. This is also the ments. To obtain this information, we use LT to extract case for the above characterization in our device, which the Lindbladian of the channel, which requires process 1˜ would imply that the manipulation and readout of qubit tomography over varying Idling Channel durations (ti). B should have no effect on the POVM and initial state of Before estimating the Lindbladian of the channel for qubit A. The extent to which this holds can be tested by continuous time t, we can first separately extract the in- comparing the estimated single-qubit POVMs and initial stantaneous evolution maps of the channel at each dis- states for both qubits with the estimated joint two-qubit crete time ti, which can then be used to check the as- POVM and initial state. The probability of measuring sumption of Markovianity. Any quantum operation can be described by a set of Kraus operators such that the fi- qubit A in state x and qubit B in state y is Tr[ρMxy], nal state is related to the initial state as ρ = ρ †, where ρ is the two-qubit density matrix and x, y 0, 1 . j Kj 0Kj The likelihood function for the two-qubit system∈ is { then} 1 where the Kraus operators satisfy j j† jP= for a optimized under the constraint that the POVM elements trace-preserving process. Note that theK K Kraus opera- sum to identity, and Cholesky decompositions can be tors are only unique up to a unitaryP transformation: a used to ensure that they are positive semidefinite. All quantum channel can be described by two different but details of the estimated two-qubit POVM can be found equivalent sets of Kraus operators j and k0 , which in the supplementary material. will be related through a unitary{K matrix} U{Ksuch} that We can quantify the deviation between the estimated j = k Ujk k0 . In standard process tomography, one single- and two-qubit POVMs by viewing the measure- thereforeK oftenK estimates a process matrix χ which is ment procedure as a quantum channel, where we “for- uniqueP in a specified operator basis. Since the process get” the measurement outcome. Specifically, we define matrix can be readily calculated from the Kraus oper- the output state of the measurement channel to be ators and vice-versa, one can choose either description without loss of generality. As we explain below, we use Λ(ρ) = Tr [ρM ] (2) MLE to estimate the Kraus operators. We also used the Pxy xy x,y=0,1 same MLE approach to estimate the process matrix but X found a slower convergence of the optimization compared where ρ is the input state and xy is the projector on to the Kraus estimation. We believe this is likely due to state xy . We now compare theP measurement chan- | i the unitary freedom in fixing the elements of the Kraus nel Λ, corresponding to the estimated two-qubit POVM, matrices. We will therefore describe the estimation of with the measurement channel Λ0, corresponding to the the Kraus operators here. product of the estimated single-qubit POVMs Mxy0 = For LT of the Idling Channel, the task is to estimate Mx My. the Kraus operators describing the qubit evolution during To⊗ compare the single- and two-qubit POVMs, we discrete delay times. For time delays ti [t1, t2, . . . , tN ] maximize the trace distance D(Λ(ρ), Λ0(ρ)) between the we consider a maximum likelihood function∈ of a similar output state of the two-qubit and single-qubit measure- form to Eq. (1), ment channels, with respect to the input state ρ. Per- forming this calculation, we find a maximum trace dis- ln( (ti)) = f(s, b, i) ln Tr [ρs(ti)Mb] tance of 0.05. LK b,s The trace distance between the estimated initial two- X  AB A B + f¯(s, b, i) ln Tr [ρ (t )(1 M )] qubit state ρ0 and the product state ρ0 ρ0 con- s i b ⊗AB A − structed from the single-qubit estimates is D(ρ0 , ρ0 (3) B ⊗  ρ0 ) = 0.01. The full single- and two-qubit matrices for the extracted initial states and POVMs are included in with parameters defined as in Eq. (1) except that ρs(ti) the supplementary material for reference. now refers to the final state of the initial ρs at time ti, 6 under the evolution of the Kraus operators (a)

Raw Data Prediction from Kraus ρs(ti) = j(ti)ρs †(ti), (4) K Kj Prediction from Lindblad j X

The elements of the Kraus operators are then esti- -state Probability mated by maximizing the log-likelihood function with re- Markovian spect to the unknown matrix elements, using the SPAM error parameters ρs and Mb found during the estima- (b) tion in Section III. For N delay times, we obtain N sets of Kraus operators where, for a d-dimensional quantum system, the process at each time is described by at most d2 Kraus operators. Thus for a single qubit, we estimate 4 Kraus operators per time delay (16 for two qubits). -state Probability

We estimate the Kraus operators at each delay time Markovian by numerically maximizing the log-likelihood in Eq. (3) under the constraint of trace-preservation. As an initial (c) guess, we initialize the minimization with the Kraus op- erators for single-qubit amplitude damping and dephas- ing noise. In Fig. 3, we show the results of extracting the single-qubit Kraus operators for qubit A of our su- perconducting transmon device. Blue dots show the raw -state Probability measurement probability p0, averaged from 1000 single- Non-Markovian shot measurements of the final state ρs(ti), and orange ’s show the predicted outcome of an imperfect measure- × Idling Channel Duration (μs) ment Mb of the state ρs(ti), estimated by applying the extracted Kraus operators to the extracted imperfect ini- FIG. 3. Lindblad Tomography framework applied to the tial state ρs as in Eq. (4). As such, the orange ’s not only capture the channel noise, but also account× for the Idling Channel of a single qubit. (a) Data and analysis re- sults for the LT sequence corresponding to a T2 Ramsey mea- SPAM errors of our device. In Fig. 3a–c, we compare surement (purple in Fig. 1a), when the neighboring qubit is the results of one LT sequence corresponding to a T - 2 prepared in its ground state 0 . Blue points are p0 of the | i like measurement of qubit A when its nearest neighbor state ρs(ti), averaged from raw data (discussed in Sec. III, B is prepared close to either its 0 -, 1 -, or + -state re- fitted value of T recorded in the supplement). Orange ’s | i | i | i 2 spectively. Comparing the raw data (blue) to our Kraus are predicted measurement outcomes obtained from apply-× estimation (orange), we find that the extracted Kraus ing the Kraus operators estimated at each discrete time ti operators predict the measurement results for all three to the extracted initial state ρ0 given an imperfect measure- scenarios. However, comparing these three scenarios, it ment M0 (technique discussed in Sec. IV). Red line traces is also clear that the state of qubit B has a significant the predicted outcomes for continuous time t, based on the effect on the evolution of qubit A. To capture the full most likely Lindblad and Hamiltonian operators (technique discussed in Sec. VI). Results for the first 10µs are enlarged dynamics of this interaction, it is thus necessary to ex- for clarity in inset. We discuss the Lindblad fit for this partic- tract the Kraus operators describing the full two-qubit ular dataset and its dependence on temporal fluctations dur- channel. ing the protocol in the supplement. (b) The same measure- The estimation of the two-qubit Kraus operators fol- ment, taken when the neighboring qubit is in its excited state 1 . The always-on ZZ-coupling between the qubits causes lows from a straight-forward generalization of the single- | i qubit LT protocol, as shown in Fig. 4a. The correspond- a state-dependent frequency shift when the neighbor is ex- ing Kraus estimation is then performed by expanding the cited, which manifests here as a faster oscillation frequency. likelihood function in Eq. (3) with all elements of the two- (c) The same measurement, taken when the neighboring qubit is in a superposition state + . In this basis, the always-on qubit POVM. Some results of this extraction are shown ZZ-coupling is an entangling| operation,i and the data cannot in Fig. 4b,e. As in the single-qubit case, we find that be predicted from any single-qubit Lindblad operator (red), the estimated Kraus operators capture the dynamics re- a hallmark of non-Markovian error (see Sec. VI). markably well. Furthermore, now that we have access to both the single- and two-qubit Kraus operators for this channel, we can thoroughly investigate the Markovianity of the Idling Channel for our device. In particular, we V. QUANTIFYING MARKOVIANITY can directly investigate how Markovian two-qubit noise due to spurious interaction between qubits can manifest There exist a number of proposed measures for non- as non-Markovian single-qubit noise. Markovianity in the literature, and we refer the interested 7

(a) (c) (d)

Initialize Sweep Channel Measure in 36 Cardinal States 9 Pauli Bases

Idling Channel

(b)

Idling Channel Duration (μs) (e)

Raw Data Prediction from Kraus Prediction from Lindblad

Idling Channel Duration (μs) Idling Channel Duration (μs)

FIG. 4. Two-qubit Lindblad Tomography protocol and results. (a) Measurement protocol: the two qubits are initialized into AB their shared ground state ρ0 and prepared in each of 36 combinations of cardinal states; the channel of interest is swept; the qubits are rotated into each of nine combinations of Pauli bases and measured. The full set of measurement results are passed through the same classical optimizer as in the single-qubit protocol, SPAM errors are extracted, and the instantaneous process maps and dynamic operators are estimated using MLE. (b) Raw data (blue), predictions from extracted Kraus operators (orange), and predictions from estimated Hamiltonian and Lindblad operators (red) for several combinations of pre- and post- pulses. (c) Schematic of a large superconducting quantum processor, where the two qubits studied in this work are thought of as neighboring qubits (A and B) in a large patchwork. LT can be performed just as easily on distant qubits (i.e. A and C) to study non-local crosstalk. (d) Single-qubit LT on qubit A while B is in a superposition state (same dataset as Fig. 3c, enhanced for visual clarity). The poor Lindblad fit (red) indicates that no single-qubit Lindblad operator successfully predicts the measured data, a hallmark of non-Markovian evolution. (e) Two-qubit Lindblad Tomography, where qubits A and B are both initialized in superposition states. While the pulse sequence is identical to (d), the data is now well-predicted by a two-qubit Lindblad operator, indicating that the channel is Markovian in the two-qubit frame. Comparing (d) and (e), we conclude that the non-Markovian errors in the single-qubit data are due to spontaneous entanglement with qubit B, revealing the error source. reader to reviews such as Ref. [26] for reference. Notably, cess. a number of experimental works have implemented the Since non-Markovian processes cannot be captured by measure proposed in Ref. [18], which quantifies the back- a time-dependent master equation of the form in Eq. 5, flow of information from the environment characteristic an increasing trace distance between states under the of non-Markovian error [27, 28]. This measure is also evolution of a common channel signifies violation of Eq. 5 suitable for our purpose, because it considers the noise and thus the presence of non-Markovian errors. Based process over time, in contrast to instantaneous measures on this observation, the measure Nmarkov is suggested in such as in Ref. [29]. Ref. [18] as The measure exploits the following fact: for any quan- tum process which can be captured by a time-dependent N = max σ(t, ρ (0), ρ (0))dt, (6) master equation of the form markov ρ1(0),ρ2(0) 1 2 Zσ>0 i ρ˙(t) = [Hˆ (t), ρ(t)] where σ(t, ρ (0), ρ (0)) = d D(ρ (t), ρ (t)). In other − 1 2 dt 1 2 ~ words, we integrate the derivative of the trace distance 1 + γ (t) Lˆ (t)ρ(t)Lˆ†(t) Lˆ†(t)Lˆ (t), ρ(t) over all time intervals where it is positive (i.e. trace i i i − 2{ i i } i distance increasing). As such, the larger the value of X   (5) Nmarkov, the more non-Markovian the channel. Having already estimated the Kraus operators for both with positive decay rates γi(t) > 0, the trace distance the single- and two-qubit Idling Channels, we can per- D(ρ1(0), ρ2(0)) between two initial states ρ1(0), ρ2(0) can form the optimization in Eq. (6) over the initial states only decrease. Here Hˆ (t) and Lˆi(t) are the time- of the LT protocol to calculate the measure Nmarkov. In dependent Hamiltonian and jump{ operators} of the pro- Fig. 5, we use the results of single- and two-qubit LT to 8

perform this optimization and calculate Nmarkov for sev- (a) eral situations. Interestingly, when qubit B is initialized in the state + (as in Fig. 3c), the Idling Channel of | i qubit A registers a significantly higher value of Nmarkov (Fig. 5c) than when qubit B is initialized in either the 0 - or 1 -state, or in the combined two-qubit channel Trace Distance Trace (Fig.| i 5a,b,d| i respectively). Furthermore, while the con- tributions to Nmarkov in Fig. 5a,b,d appear to arise from individual statistical fluctuations in the data due to finite sampling, the increases in trace distance in Fig. 5c corre- (b) spond to clear oscillations over the channel duration. We note that, were one to interpolate the data, the impact of statistical fluctations in the distance measure would be significantly reduced, resulting in a smaller value of Nmarkov for Fig. 5a,b,d, while the value for the case of Distance Trace Fig. 5c would be essentially unchanged. This indicates that, when qubit B is prepared in + , it induces dis- | i tinctly non-Markovian errors in the noise environment of (c) qubit A. The presence of single-qubit non-Markovian behavior in Fig. 3c is well understood from the physics of cou- pled transmon qubits. For two transmon qubits inter-

acting via a fixed capacitance, the resulting dispersive Distance Trace repulsion of the 20 - and 02 -states shifts the frequency of the 11 -state| andi gives| risei to a ubiquitous “always- on” ZZ|-interactioni in the computational subspace of the form [30]: (d)

ˆ ωzz Hzz/~ = ωzz 11 11 = (ZZ ZI IZ + II) (7) | ih | 4 − − Trace Distance Trace where ωzz = ω11 ω01 ω10 is the energy shift of the 11 -state due to the− qubit− coupling. | Consequently,i when the two qubits are far detuned from each other, this interaction results in an effective Idling Channel Duration (μs) two-qubit Hamiltonian of the form [31]: FIG. 5. Markovianity of single- and two-qubit Idling Chan- nels. (a–c) Comparing the measured Markovianity of qubit ˆ A’s Idling Channel when qubit B is prepared in 0 , 1 , or + H/~ = ωA 10 10 + ωB 01 01 | i | i | i | ih | | ih | respectively. We find the two initial states of qubit A with + (ωA + ωB + ωzz) 11 11 (8) the largest trace distance D at time t (blue points) and take | ih | i the difference in trace distance between sequential times (red where ω , ω are the 0 1 transition frequencies of triangles, values less than 0 omitted for visual clarity, since A B | i→| i qubit A and B respectively. When one of the qubits is they do not contribute to Nmarkov). Summing the red points prepared in either 0 or 1 , this interaction is manifest amounts to the discrete version of Eq. (6), and Nmarkov values | i | i are shown in each figure. When qubit B is prepared in + as as a state-dependent frequency shift (hence the difference | i in oscillation frequency between Fig. 3a and b). However, in (c), we observe clear periods of decreasing trace distance when the two qubits are prepared in an initial state ++ , and find the largest value of Nmarkov, indicating the greatest | i presence of non-Markovian errors. (d) The same protocol as they will evolve into an entangled state under the influ- above, taken using two-qubit LT and comparing the trace dis- ence of the Hamiltonian in Eq. 8. If we only consider the tance between two-qubit initial states (purple). Maximizing evolution of one of the qubits, as is the case for the single- over all two-qubit initial states, we find that none display a qubit Kraus estimation, the always-on coupling will swap non-Markovian signature as seen in (c), indicating the channel information between the qubit we are measuring and its errors are largely Markovian in the two-qubit frame. neighbor, which effectively functions as an environment with memory, resulting in non-Markovian error. How- ever, if the evolution of both qubits is considered, as in the two-qubit Kraus estimation, the always-on interac- tion is revealed to be unitary and this non-Markovianity disappears (Fig. 5d). 9

VI. EXTRACTING THE LINDBLADIAN results of our single- and two-qubit measurements (blue data in Fig. 3 and Fig. 4). To do this, we must account for While we have shown how the Markovianity of the all combinations of initial states and measurement axes noise environment can be assessed from the estimated (as in our Kraus extraction), for all channel durations ti. Kraus operators, it is often difficult to extract much We therefore seek to maximize the log-likelihood function physical insight from the Kraus operators alone. If the N noise is Markovian, one can therefore apply LT to esti- ln( ) = ln( (t )), (12) mate the time-independent Lindbladian which best fits LLT L i i=1 the measurement data for continuous times t (right path X in Fig. 1b). We note, however, that the Markovianity where ln( (ti)) is defined as in Eq. (3), except that we measure only tells if the process can be captured by a L no longer write ρs(ti) = iρs †. Instead, we numer- general master equation with a time-dependent Lindbla- i K Ki ically solve Eq. (10) with initial state ρs for each time ti dian, and it does not guarantee that the assumption P to obtain ρs(ti) for a given parametrization of the Hamil- of a time-independent Lindbladian is fulfilled. For a tonian and the Lindblad matrix. As with the SPAM and time-independent Lindbladian, the master equation from Kraus estimation, a Cholesky decomposition is used to Eq. (5) simplifies to ensure that the matrix is positive semidefinite. The evolution from the estimated Hamiltonian and d2 1 i − 1 Lindblad matrix for a single qubit is shown in Fig. 3. ρ˙ = [H,ˆ ρ] + γ (Lˆ ρLˆ† Lˆ†Lˆ , ρ ). (9) − i i i − 2{ i i } Here, the red solid line traces out the predicted measure- ~ i=1 X ment results for the continuous evolution of the qubit over all times t. When qubit B is in either its 0 - Choosing an operator basis σi consisting of a Hilbert- { } state (Fig. 3a) or 1 -state (Fig. 3b), the Lindblad evo-| i Schmidt orthogonal set of traceless Hermitian operators | i in dimension d (which can be constructed from tensor lution of qubit A well fits the results of both measure- products of single-qubit Pauli matrices and the identity), ment (blue dots) and the prediction of our extracted dis- the master equation can be rewritten as crete processes (orange ’s). However, when the neigh- boring qubit is prepared× in the superposition state + | i d2 1 (Fig. 3c), we find that our estimation fails to find a combi- i − 1 ρ˙ = [H,ˆ ρ] + L (σ ρσ† σ†σ , ρ ), (10) nation of Hamiltonian and Lindbladian which well fit the − ij i j − 2{ j i } ~ i,j=1 data. As discussed above, this is expected given the non- X Markovian signature we previously found from the Kraus where Lij is a Hermitian and positive semidefinite matrix estimation (Fig. 5c), since non-Markovian processes can- capturing the incoherent evolution, commonly referred to not be captured by a master equation of the form in Eq. as the Lindblad matrix. We note that, similar to the pro- 10. cess matrix χ, the Lindblad matrix is unique while the In Fig. 4, we plot a subset of the results of Hamilto- jump operators, like the Kraus operators, have a unitary nian and Lindblad extraction using two-qubit LT (red freedom: the Lindblad equation in Eq. (9) is invariant solid line), and we show that the estimated operators under a unitary transformation of the jump operators well fit both measurement (blue dots) and the predic- and decay rates. In particular, a new set of jump oper- tion of our extracted Kraus operators (orange ’s) for all × ators and decay rates γ Lˆ0 can be constructed from combinations of initial state and measurement axis. In { i0 i} the set γ Lˆ as γ Lˆ = U γ Lˆ where U is a particular, we note an excellent Lindblad fit for the se- √ j j ip0 i0 j ij√ j j unitary{ matrix.} Since the Lindblad matrix can readily be quence where both qubits are prepared in superposition obtained from the decayp ratesP and jump operators (and states ++ (Fig. 4e), indicating that the non-Markovian | i vice versa), one can choose either representation without errors observed in the corresponding single-qubit channel loss of generality. In the following analysis, we choose (Fig. 3c, 4d) largely disappear in the two-qubit frame. to directly estimate the Lindblad matrix, and we derive From the two-qubit data, we estimate the following the jump operators by diagonalizing this matrix. Specif- two-qubit Hamiltonian ically, by diagonalizing the Lindblad matrix, a (unique) 0.00 0.01i 0.00 0.00 set of jump operators can be obtained as ˆ 0.01i 1.03 0.02i 0.04 0.02i He = ~ −0.00 −0.02i 0.25− 0.−00 (13) 0.00 0−.04+0.02i −0.00 1.33 d2 1  −  − Lˆ = U σ (11) with angular frequencies in units of 2π MHz, relative to i ij j × j=1 the lab-frame of the driving pulses used for single-qubit X rotations. Were these pulses chosen resonant with the where U is a unitary matrix such that L = UDU †, where qubit frequencies, we would expect the diagonal elements D = diag(γ1, γ2, . . . , γd 1) is a diagonal matrix with the of the Hamiltonian to be zero, except for the 11 -state, decay rates. − where the always-on ZZ-coupling shifts the energy.| i Performing LT, our goal is to estimate the Hamiltonian From the Hamiltonian extracted in Eq. 13, we see and the Lindblad matrix which most likely describe the that there is a frequency detuning of ∆ωA/2π = 10

0.25/2π MHz = 40kHz for qubit A and ∆ωB/2π = state of the unrestricted optimization has a finite tem- AB 1.03/2π MHz = 164kHz for qubit B, which give rise to perature. We also compare the steady state (ρss ) of AB the oscillations seen in Fig. 3a (qubit A) and Fig. 4b the free optimization to the initial two-qubit state (ρ0 ) bottom left figure (qubit B). The effect of the ZZ- from the SPAM estimation. We find a trace distance AB AB coupling is evident when qubit B is excited close to its of D(ρss , ρ0 ) = 0.06 showing that the two states are 1 -state (Fig. 3b), and we consequently observe an in- similar, consistent with the fact that the superconduct- crease| i in the frequency offset for qubit A. From the ex- ing qubits are initialized by waiting many multiples of tracted Hamiltonian, we can estimate the frequency shift T1, letting them relax to the steady state of the Idling from the ZZ-coupling to be ωzz/2π = (1.33 ( 0.25 Channel. 1.03))/2π MHz = 416kHz, which is consistent− with− inde-− pendent estimation from device parameters [22]. Having estimated the unitary part of the Idling Chan- VII. CONCLUSION AND DISCUSSION nel, we turn to the extracted two-qubit jump operators and decay rates corresponding to information loss during In this work, we have proposed a technique for extract- the channel. The full expressions of the extracted jump- ing the Hamiltonian, jump operators, and corresponding operators and decay rates are shown in the supplemen- decay rates from a general quantum channel, which we tary material. Motivated by our understanding of the refer to as Lindblad Tomography (LT). Combining tech- dominant error mechanisms for superconducting qubits, niques from state and process tomography with Hamil- we can compare our extracted operators to single-qubit tonian, Lindblad, and SPAM error estimation based on amplitude damping and dephasing noise (i.e. T1 and T2 maximum likelihood (MLE), Lindblad Tomography pro- processes respectively), which correspond to jump oper- vides detailed information about the errors and noise en- ators of the form vironment of quantum devices and can be used to identify ˆ crosstalk. Ld,1 σz I (14) ∝ ⊗ Applying LT to the characterization of a small super- ˆ Ld,2 I σz (15) ∝ ⊗ conducting device, we demonstrate how the framework ˆ quantifies crosstalk in qubit readout, estimates frequency La,1 σ I (16) ∝ − ⊗ offsets in driving pulses, and the strength of always-on ˆ La,1 I σ , (17) ∝ ⊗ − interactions between qubits, which can result in single- qubit non-Markovain noise which we quantified using the where the operators σz and σ = 0 1 correspond to − measure of Ref. [18]. Furthermore, our characterization dephasing and amplitude damping respectively.| ih | shows that the noise environment which best describes To investigate whether T and T accurately describe 1 2 the data is consistent with single-qubit amplitude damp- the evolution of our qubits, we run a separate maximum ing (T ) and dephasing (T ) channels to a large extent. likelihood optimization of the Lindbladian, this time fix- 1 2 ing the jump operators to be of the form in Eq. (14)–(17) While much of this proof-of-principal study focused on noise processes which arise due to the presence of and leaving only the rates (γ) and Hamiltonian (Hˆ ) as neighboring qubits, we note that LT can be naturally ap- free parameters. We will refer to this as the restricted op- plied to study a broad range of noise sources in the qubit timization, while the previous optimization over general environment—such as coupling to coherent two-level sys- jump operators is referred to as the free optimization. tems (TLSs) [32], dephasing due to photons in readout We compare the output of this restricted optimization resonators [33], and interaction with quasiparticles [34] in by looking at the diamond norm between the two Liou- superconducting systems—all of which may contribute to villian superoperators, the extracted Lindbladian and result in varying degrees of non-Markovian error. Additionally, we believe that δ(t) = Φ(Lfree, t) Φ(Lrestricted, t) (18) k − k further investigation of changes in the noise environment 3 where Lfree (Lrestricted) is the Liouvillian superoperator over time [17] could offer a promising direction for fu- corresponding to the free (restricted) optimization, and ture inquiry, and we are confident that LT will prove a Φ(L, t) is the Choi-matrix representation of Exp(Lt). For valuable tool in future work towards suppressing these t 80 µs, we find that δ(t) 0.20. We conclude that, errors using either quantum control or error correction while≤ the extracted jump-operators≤ from the unrestricted techniques. optimization do differ from single-qubit amplitude damp- We conclude by noting that the number of measure- ing and dephasing channels, the channels corresponding ments required for Lindblad Tomography scales exponen- to standard T1 and T2 measurements nonetheless de- tially with the number of qubits, as with standard pro- scribe the data reasonably well. For t , we find cess tomography. Thus, full characterization of a large a maximum deviation of δ(t) 0.44, indicating→ ∞ that quantum processor remains experimentally impractical. the two evolutions arrive at different→ steady states. This However, since crosstalk between qubits is almost en- can be understood from the fact that the steady state tirely 2-body, characterization of all combinations of two- of the amplitude damping channel in the restricted op- qubit patches on a large quantum processor using LT timization is the global ground state, while the steady will nevertheless provide valuable insights into the col- 11 lective noise environment of the full processor, and these permission of the US Government sponsors who funded measurements can be used to bootstrap higher-order er- the work. rors [35, 36]. As the number of two-qubit patches only scales as N 2 for an N-qubit device, characterization of direct two-qubit∼ crosstalk with Lindblad Tomography ACKNOWLEDGMENTS can therefore, in principle, be done efficiently. Further- more, for devices where it is reasonable to assume that The authors gratefully acknowledge Agustin Di Paolo, crosstalk is restricted to pairs of qubits within a certain Daniel Stilck Fran¸ca,and Joey Barreto for their feedback maximal separation (as may well be the case for devices on the manuscript; Chihiro Watanabe and Mirabella with equally spaced qubits), the number of pairs to be Pulido for their assistance during this project; and the characterized would only scale as (N), though investi- MIT facilities and custodial staff for all their work in O gating the validity of this approach remains the subject of maintenance of the laboratory space. AG acknowledges future work. As research scales to larger and more com- funding from the 2019 Google US/Canada PhD Fellow- plex systems and the possible sources of crosstalk and ship in Quantum Computing. JB acknowledges fund- unintentional qubit entanglement inevitably increases, ing from the NWO Gravitation Program Quantum Soft- we are confident that repeated Lindblad Tomography of ware Consortium. MC acknowledges financial support single- and two-qubit patches will provide an important from the European Research Council (ERC GrantA- step towards reduced-complexity modeling of large-scale greement No. 81876), VILLUM FONDEN via the quantum processors. QMATH Centre of Excellence (Grant No.10059) and the During the final preparation of this manuscript, the QuantERA ERA-NET Cofund in Quantum Technolo- authors became aware of a recent and separate the- gies implemented within the European Union’s Horizon ory work which proposes, implements, and numerically 2020 Programme (QuantAlgo project) via the Innovation benchmarks the fitting of tomography data to quantum Fund Denmark. MC also acknowledges the hospitality of noise models, and we direct the interested reader to that the Center for Theoretical Physics at MIT, where part manuscript for comparison [37]. of this work was carried out. MK was supported by Villum Fonden (grant 37467) through a Villum Young Investigator grant for part of this work. This research SOFTWARE AND DATA AVAILABILITY was funded in part by the U.S. Army Research Office Grant W911NF-18-1-0411 and the Assistant Secretary The code used for analyzing the results of Lindblad of Defense for Research & Engineering under Air Force Tomography throughout this work can be found in the Contract No. FA8721-05-C-0002. Opinions, interpreta- public GitHub repository [38]. The experimental data tions, conclusions, and recommendations are those of the used in this work may be made available upon reason- authors and are not necessarily endorsed by the United able request of the corresponding authors and with the States Government.

[1] Ashley Montanara, “Quantum algorithms: An ger, Matthew P. Harrigan, Michael J. Hartmann, Alan overview,” npj 2, 15023 (2016). Ho, Markus Hoffmann, Trent Huang, Travis S. Humble, [2] I. M. Georgescu, S. Ashhab, and Franco Nori, “Quantum Sergei V. Isakov, Evan Jeffrey, Zhang Jiang, Dvir Kafri, simulation,” Rev. Mod. Phys. 86, 153–185 (2014). Kostyantyn Kechedzhi, Julian Kelly, Paul V. Klimov, [3] C. Figgatt, A. Ostrander, N. M. Linke, K. A. Landsman, Sergey Knysh, Alexander Korotkov, Fedor Kostritsa, D. Zhu, D. Maslov, and C. Monroe, “Parallel entangling David Landhuis, Mike Lindmark, Erik Lucero, Dmitry operations on a universal ion-trap quantum computer,” Lyakh, Salvatore Mandr`a,Jarrod R. McClean, Matthew Nature 572, 368–372 (2019). McEwen, Anthony Megrant, Xiao Mi, Kristel Michielsen, [4] Alexander Keesling, Ahmed Omran, Harry Levine, Masoud Mohseni, Josh Mutus, Ofer Naaman, Matthew Hannes Bernien, Hannes Pichler, Soonwon Choi, Rhine Neeley, Charles Neill, Murphy Yuezhen Niu, Eric Os- Samajdar, Sylvain Schwartz, Pietro Silvi, Subir Sachdev, tby, Andre Petukhov, John C. Platt, Chris Quintana, Peter Zoller, Manuel Endres, Markus Greiner, Vladan Eleanor G. Rieffel, Pedram Roushan, Nicholas C. Ru- Vuleti´c, and Mikhail D. Lukin, “Quantum kibble–zurek bin, Daniel Sank, Kevin J. Satzinger, Vadim Smelyan- mechanism and critical dynamics on a programmable ry- skiy, Kevin J. Sung, Matthew D. Trevithick, Amit dberg simulator,” Nature 568, 207–211 (2019). Vainsencher, Benjamin Villalonga, Theodore White, [5] Frank Arute, Kunal Arya, Ryan Babbush, Dave Bacon, Z. Jamie Yao, Ping Yeh, Adam Zalcman, Hartmut Neven, Joseph C. Bardin, Rami Barends, Rupak Biswas, Sergio and John M. Martinis, “Quantum supremacy using a Boixo, Fernando G. S. L. Brandao, David A. Buell, Brian programmable superconducting processor,” Nature 574, Burkett, Yu Chen, Zijun Chen, Ben Chiaro, Roberto 505–510 (2019). Collins, William Courtney, Andrew Dunsworth, Ed- [6] Pavithran Iyer and David Poulin, “A small quantum ward Farhi, Brooks Foxen, Austin Fowler, Craig Gidney, computer is needed to optimize fault-tolerant protocols,” Marissa Giustina, Rob Graff, Keith Guerin, Steve Habeg- Quantum Science and Technology 3, 030504 (2018). 12

[7] David K. Tuckett, Stephen D. Bartlett, and Steven T. [23] P. Krantz, M. Kjaergaard, F. Yan, T. P. Orlando, S. Gus- Flammia, “Ultrahigh error threshold for surface codes tavsson, and W. D. Oliver, “A quantum engineer’s guide with biased noise,” Phys. Rev. Lett. 120, 050505 (2018). to superconducting qubits,” Applied Physics Reviews 6, [8] Joseph Emerson, Robert Alicki, and Karol Zyczkowski,˙ 021318 (2019), https://doi.org/10.1063/1.5089550. “Scalable noise estimation with random unitary opera- [24] Morten Kjaergaard, Mollie E. Schwartz, Ami Greene, tors,” Journal of Optics B: Quantum and Semiclassical Gabriel O. Samach, Andreas Bengtsson, Michael Optics 7, S347–S352 (2005). O’Keeffe, Christopher M. McNally, Jochen Braum¨uller, [9] Easwar Magesan, J. M. Gambetta, and Joseph Emerson, David K. Kim, Philip Krantz, Milad Marvian, Alexan- “Scalable and robust randomized benchmarking of quan- der Melville, Bethany M. Niedzielski, Youngkyu Sung, tum processes,” Phys. Rev. Lett. 106, 180504 (2011). Roni Winik, Jonilyn Yoder, Danna Rosenberg, Kevin [10] Robin Blume-Kohout, John King Gamble, Erik Nielsen, Obenland, Seth Lloyd, Terry P. Orlando, Iman Marvian, Kenneth Rudinger, Jonathan Mizrahi, Kevin Fortier, Simon Gustavsson, and William D. Oliver, “Program- and Peter Maunz, “Demonstration of qubit operations ming a quantum computer with quantum instructions,” below a rigorous fault tolerance threshold with gate set (2020), arXiv:2001.08838 [quant-ph]. tomography,” Nature Communications 8, 14485 (2017). [25] David K. Tuckett, Andrew S. Darmawan, Christo- [11] Matteo Paris and Jaroslav Rehacek, eds., pher T. Chubb, Sergey Bravyi, Stephen D. Bartlett, and Estimation (Springer-Verlag Berlin Heidelberg, 2004). Steven T. Flammia, “Tailoring surface codes for highly [12] Jens Eisert, Dominik Hangleiter, Nathan Walk, Ingo biased noise,” Phys. Rev. X 9, 041031 (2019). Roth, Damian Markham, Rhea Parekh, Ulysse Chabaud, [26] In´es de Vega and Daniel Alonso, “Dynamics of non- and Elham Kashefi, “Quantum certification and bench- Markovian open quantum systems,” Rev. Mod. Phys. 89, marking,” Nature Reviews Physics 2, 382–390 (2020). 015001 (2017). [13] D Mogilevtsev, J Reh´aˇcek,ˇ and Z Hradil, “Self- [27] Bi-Heng Liu, Li Li, Yun-Feng Huang, Chuan-Feng Li, calibration for self-consistent tomography,” New Journal Guang-Can Guo, Elsi-Mari Laine, Heinz-Peter Breuer, of Physics 14, 095001 (2012). and Jyrki Piilo, “Experimental control of the transi- [14] Seth T. Merkel, Jay M. Gambetta, John A. Smolin, Ste- tion from Markovian to non-Markovian dynamics of open fano Poletto, Antonio D. C´orcoles,Blake R. Johnson, quantum systems,” Nature Physics 7, 931–934 (2011). Colm A. Ryan, and Matthias Steffen, “Self-consistent [28] Marko Znidariˇc,ˇ Carlos Pineda, and Ignacio Garc´ıa- quantum process tomography,” Phys. Rev. A 87, 062119 Mata, “Non-Markovian behavior of small and large com- (2013). plex quantum systems,” Phys. Rev. Lett. 107, 080404 [15] J. Medford, J. Beil, J. M. Taylor, S. D. Bartlett, A. C. (2011). Doherty, E. I. Rashba, D. P. DiVincenzo, H. Lu, A. C. [29] M. M. Wolf, J. Eisert, T. S. Cubitt, and J. I. Cirac, “As- Gossard, and C. M. Marcus, “Self-consistent measure- sessing non-Markovian quantum dynamics,” Phys. Rev. ment and state tomography of an exchange-only spin Lett. 101, 150402 (2008). qubit,” Nature Nanotechnology 8, 654–659 (2013). [30] M. Ganzhorn, G. Salis, D. J. Egger, A. Fuhrer, M. Mer- [16] Dohun Kim, D. R. Ward, C. B. Simmons, John King genthaler, C. M¨uller,P. M¨uller,S. Paredes, M. Pechal, Gamble, Robin Blume-Kohout, Erik Nielsen, D. E. Sav- M. Werninghaus, and S. Filipp, “Benchmarking the noise age, M. G. Lagally, Mark Friesen, S. N. Coppersmith, sensitivity of different parametric two-qubit gates in a and M. A. Eriksson, “Microwave-driven coherent opera- single superconducting quantum computing platform,” tion of a semiconductor quantum dot charge qubit,” Na- Phys. Rev. Research 2, 033447 (2020). ture Nanotechnology 10, 243–247 (2015). [31] P. J. J. O’Malley, J. Kelly, R. Barends, B. Camp- [17] Jonathan J. Burnett, Andreas Bengtsson, Marco Scigli- bell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, uzzo, David Niepce, Marina Kudra, Per Delsing, and A. G. Fowler, I.-C. Hoi, E. Jeffrey, A. Megrant, J. Mu- Jonas Bylander, “Decoherence benchmarking of super- tus, C. Neill, C. Quintana, P. Roushan, D. Sank, conducting qubits,” npj Quantum Information 5, 54 A. Vainsencher, J. Wenner, T. C. White, A. N. Korotkov, (2019). A. N. Cleland, and John M. Martinis, “Qubit metrology [18] Heinz-Peter Breuer, Elsi-Mari Laine, and Jyrki Piilo, of ultralow phase noise using randomized benchmark- “Measure for the degree of non-Markovian behavior of ing,” Phys. Rev. Applied 3, 044009 (2015). quantum processes in open systems,” Phys. Rev. Lett. [32] P. V. Klimov, J. Kelly, Z. Chen, M. Neeley, A. Megrant, 103, 210401 (2009). B. Burkett, R. Barends, K. Arya, B. Chiaro, Yu Chen, [19] M Howard, J Twamley, C Wittmann, T Gaebel, A. Dunsworth, A. Fowler, B. Foxen, C. Gidney, F Jelezko, and J Wrachtrup, “Quantum process tomog- M. Giustina, R. Graff, T. Huang, E. Jeffrey, Erik raphy and linblad estimation of a solid-state qubit,” New Lucero, J. Y. Mutus, O. Naaman, C. Neill, C. Quintana, Journal of Physics 8, 33–33 (2006). P. Roushan, Daniel Sank, A. Vainsencher, J. Wenner, [20] Eitan Ben Av, Yotam Shapira, Nitzan Akerman, and T. C. White, S. Boixo, R. Babbush, V. N. Smelyan- Roee Ozeri, “Direct reconstruction of the quantum- skiy, H. Neven, and John M. Martinis, “Fluctuations master-equation dynamics of a trapped-ion qubit,” Phys. of energy-relaxation times in superconducting qubits,” Rev. A 101, 062305 (2020). Phys. Rev. Lett. 121, 090502 (2018). [21] Morten Kjaergaard, Mollie E. Schwartz, Jochen [33] D. I. Schuster, A. Wallraff, A. Blais, L. Frunzio, R.-S. Braum¨uller,Philip Krantz, Joel I.-J. Wang, Simon Gus- Huang, J. Majer, S. M. Girvin, and R. J. Schoelkopf, tavsson, and William D. Oliver, “Superconducting “ac stark shift and dephasing of a superconducting qubit qubits: Current state of play,” Annual Review of Con- strongly coupled to a cavity field,” Phys. Rev. Lett. 94, densed Matter Physics 11, 369–395 (2020). 123602 (2005). [22] See supplementary material. [34] K. Serniak, M. Hays, G. de Lange, S. Diamond, S. Shankar, L. D. Burkhart, L. Frunzio, M. Houzet, and 13

M. H. Devoret, “Hot nonequilibrium quasiparticles in transmon qubits,” Phys. Rev. Lett. 121, 157701 (2018). [35] L. C. G. Govia, G. J. Ribeill, D. Rist`e,M. Ware, and H. Krovi, “Bootstrapping quantum process tomography via a perturbative ansatz,” Nature Communications 11, 1084 (2020). [36] Megan N. Lilly and Travis S. Humble, “Modeling noisy quantum circuits using experimental characterization,” arXiv:2001.08653 (2020). [37] Emilio Onorati, Tamara Kohler, and Toby Cubitt, “Fit- ting quantum noise models to tomography data,” (2021), arXiv:2103.17243 [quant-ph]. [38] Lindblad Tomography GitHub repository: https:// github.com/jborregaard/Lindblad_tomography.git. Supplementary Material for “Lindblad Tomography of a Superconducting Quantum Processor”

1, 2, 1 3, 4, 4 Gabriel O. Samach, ∗ Ami Greene, Johannes Borregaard, † Matthias Christandl, David K. Kim,2 Christopher M. McNally,1 Alexander Melville,2 Bethany M. Niedzielski,2 Youngkyu Sung,1 Danna Rosenberg,2 Mollie E. Schwartz,2 Jonilyn L. Yoder,2 Terry P. Orlando,1 Joel I-Jan Wang,1 Simon Gustavsson,1 Morten Kjaergaard,1, 5 and William D. Oliver1, 2, 6 1Research Laboratory of Electronics, Massachusetts Institute of Technology, Cambridge, MA, USA 2MIT Lincoln Laboratory, Lexington, MA, USA 3Qutech and Kavli Institute of Nanoscience, Delft University of Technology, Delft, The Netherlands 4Department of Mathematical Sciences, University of Copenhagen, Copenhagen, Denmark 5Center for Quantum Devices, Niels Bohr Institute, University of Copenhagen, Copenhagen, Denmark 6Department of Electrical Engineering and Computer Science, Massachusetts Institute of Technology, Cambridge, MA, USA (Dated: June 15, 2021)

CONTENTS

Appendix A: Device and Measurement Infrastructure 1

Appendix B: Single-Qubit Gate Characterization 1

Appendix C: POVM and Lindblad Estimation 2

Appendix D: Single-Qubit LT Results and Discussion 5

References 7

APPENDIX A: DEVICE AND MEASUREMENT INFRASTRUCTURE

The quantum processor characterized in this work consists of three capacitively coupled superconducting flux- tunable transmon qubits arranged in a linear chain (Fig. S1a). For this initial proof-of-principle demonstration of single- and two-qubit Lindblad Tomography, we chose to consider only the left and middle qubits of the chain, which we label qubit A and B respectively. The rightmost qubit is far detuned to its frequency minimum and left to idle in its ground state for the duration of the characterization protocol. Significant device parameters for qubits A and B are noted in Table I. In Fig. S1b, we outline the control and readout hardware used to perform gate operations and measure the state of the qubits inside a dilution refrigerator. Additional characterization of the device used in this

arXiv:2105.02338v2 [quant-ph] 13 Jun 2021 experiment can be found in Ref. [1].

APPENDIX B: SINGLE-QUBIT GATE CHARACTERIZATION

In its most general formulation, Lindblad Tomography characterizes not only errors which arise during the channel of interest, but also errors in qubit state preparation and measurement (SPAM). However, since these three categories of error do not necessarily have independent experimental signatures, it is not always possible to fully disambiguate the relative contribution of each source of error. To simplify our characterization protocol significantly, we add in the

∗ Corresponding author: [email protected] † Corresponding author: [email protected] 2

Readout (a) (b) Microwave Drive OUT Readout IN Flux AWG V Static Voltage V RT Amp 300 K

50 K

3 K RC RC HEMT

Still

M/C

Global Flux Bias

100 μm 50 Ω 50 Ω

TWPA VLF-7200 Bias-tee -20 dB

VLFX-300 Isolator -10 dB

RLC electronics 3 GHz HP Directional coupler RC RC filter 50 Ω

RLC electronics 12.4 GHz LP Amplifier

FIG. S1. Device and Wiring Diagram. (a) SEM image of an identically fabricated copy of the device characterized in this work. (b) Schematic of the control and readout hardware used to operate the quantum processor characterized in this experiment. Dashed horizontal lines indicate the thermal stages of the dilution refrigerator in which the processor is measured.

Parameter Qubit A Qubit B

Idling Frequency, ωi/2π 4.744GHz 4.222GHz Anharmonicity, η/2π -175MHz -190MHz Coupling Strength, g/2π 12MHz Junction Asymmetry 1:5 1:10 Single-qubit Gate Time 30ns 30ns

Readout Resonator Frequency, ωr/2π 7.252GHz 7.285GHz

Energy Relaxation Time, T1 30µs 40µs Ramsey Decay Time, T2 27µs 23µs

TABLE I. Device parameters for the two qubits characterized in this work. Reported values of T1 and T2 are found by fitting the raw data from the corresponding subsets of the single-qubit LT sequence (highlighted gates in Fig. 1a of the main text) and recording the decay time of the fit, consistent with standard experimental convention. LT generalizes this technique by extracting the decay channel from the full set of initial states, measurement axes, and channel durations, as emphasized in Fig. S4.

additional assumption that our state preparation and measurement axis rotations (Rs and Rb) have negligible error in comparison to thermalization, measurement, and channel errors. To verify that this assumption is valid for our experimental setup, we perform Interleaved Clifford Randomized Benchmarking (IRB) on the full set of single-qubit gates R 1, Xπ, Y π , X π required to run Lindblad Tomography ∈ { ±2 ∓2 } on each qubit. For each of these operations, we record fidelities in excess of 99.9%, over an order of magnitude greater than the fidelity observed for state initialization or measurement (Fig. S2).

APPENDIX C: POVM AND LINDBLAD ESTIMATION

We estimated the SPAM errors both for each qubit individually and for the combined two-qubit system. In both cases, the maximization of the log-likelihood function was performed using MATLAB’s built-in function fmincon, 3

(a) Qubit A (b) Qubit B -state Probability -state Probability

Number of Clifford Gates Number of Clifford Gates

FIG. S2. Single-qubit Reference and Interleaved Clifford Randomized Benchmarking. Characterization of nine single-qubit gates performed on qubit A (a) and qubit B (b). Clifford reference fidelity (Ref) and interleaved gate fidelities are recorded in the legend. This set includes all the gates required for state preparation (Rs) and measurement axis rotation (Rb). Note that each of these gates exceeds an interleaved RB fidelity of 99.9%, above the threshold for discounting rotation errors in our Lindblad Tomography protocol.

allowing us to enforce the constraint that the POVM matrices sum to identity. We note that, to avoid the optimizer getting stuck in local minima, multiple starting points were tried to find a good approximation of the global minimum. For the single-qubit estimation of qubit A, we used data where the neighboring qubit B was kept in the ground state. The POVMs found from the maximization are

A 1.00 0.01+0.01i B 1.00 0.00 M0 = 0.01 0.01i − 0.04 ,M0 = ( 0.00 0.04 ) (S1) − − indicating that there is a measurement error of 4% corresponding to measuring the state 1 incorrectly as 0 . The | i | i estimated initial states are:

A 0.86 0.01+0.01i B 0.88 0.01 0.02i ρ0 = 0.01 0.01i 0.14 , ρ0 = 0.01+0.02i 0−.12 . (S2) −  

For the two-qubit system, we found the following POVM elements

1.00 0.00 0.00 0.00 0.00 0.04 0.02 0.03i 0.00 M00 = 0.00 0.02+0.03i 0−.04 0.00 , (S3)  0.00 0.00 0.00 0.00  0.00 0.00 0.00 0.00 0.00 0.95 0.00 0.01+0.02i M01 = 0.00 0.00 0.00 0.00 , (S4) 0.00 0.01 0.02i 0.00 0.04  −  0.00 0.00 0.00 0.00 0.00 0.01 0.02+0.02i 0.00 M10 = 0.00 0.02 0.02i − 0.96 0.01 , (S5)  0.00− 0.−01 0.01 0.02  0.00 0.00 0.00 0.00 0.00 0.01 0.01i 0.02i M11 = 0.00 0.01i 0.01 − 0.01 (S6) 0.00− 0.02i 0.01− 0.94  −  and an initial two-qubit state of

0.76 0.01 0.01i 0.01+0.01i 0.00 AB 0.01+0.01i 0−.11 0.00 0.00 ρ0 = 0.01 0.01i 0.00 0.12 0.00 (S7)  0−.00 0.00 0.00 0.02  4 Diamond Norm

Idling Channel Duration (μs)

FIG. S3. Deviation between the estimated Liouvillian of the restricted and free optimization. The deviation is calculated as the diamond norm of the difference between the two superoperators [2].

The Lindblad operators estimated for the two-qubit system using the log-likelihood maximization are shown below.

0.48 0.12+0.11i 0.00 0.00 ˆ 0.08 0.05i − 0.48 0.00 0.00 L1 = − 0.−00− 0.00 0.48 0.12+0.11i , (S8) 0.00 0.00 0.08 0.05i − 0.48  − − −  0.10+0.05i 0.63+0.02i 0.01 0.00 ˆ 0.28 0.02i 0.10 0.06i −0.00 0.01 L2 = 0−.00− 0.−00 0.10+0.06i 0.63+0− .02i , (S9) 0.00 0.00 0.28 0.02i 0.10 0.06i  − − −  0.50 0.00 0.02i 0.00 ˆ 0.00 0.50− 0.00 0.02i L3 = 0.00 0.00 0.50− 0.00 , (S10) 0.00 0.00− 0.00 0.50  −  0.01 0.01i 0.66i 0.00 ˆ −0.00 − 0.01− 0.00 0.66i L4 = 0.26i −0.00 0.01 −0.01i . (S11) 0.00 0.01+0.26i 0.00− 0.01  − 

The corresponding decay rates were found to be γ1 = 0.10 MHz, γ2 = 0.05 MHz, γ3 = 0.07 MHz, and γ4 = 0.06 MHz. We fixed the unitary freedom of the jump operators by deriving them from the diagonalization of the Lindblad matrix, as discussed in the main text, and we normalize them such that Tr LˆLˆ† = 1. As noted in the main text, the { } above jump operators are similar to single-qubit amplitude damping and dephasing channels, which are of the form ˆ ˆ ˆ ˆ Ld,1 σz I, Ld,2 I σz, La,1 σ I, and La,1 I σ . Using the same normalization as above: ∝ ⊗ ∝ ⊗ ∝ − ⊗ ∝ ⊗ − 1/2 0 0 0 0 1/2 0 0 ˆ − Ld,1 = 0 0 1/2 0 (S12) 0 0 0 1/2 ! − 1/2 0 0 0 ˆ 0 1/2 0 0 Ld,2 = 0 0 1/2 0 , (S13) − 0 0 0 1/2 ! − 0 1/√2 0 0 Lˆ = 0 0 0 0 , (S14) a,1 0 0 0 1/√2 0 0 0 0 ! 0 0 1/√2 0 √ Lˆa,2 = 0 0 0 1/ 2 (S15) 0 0 0 0 0 0 0 0 !

Note that Lˆ1 and Lˆ3 are very similar to single-qubit dephasing channels (Lˆd,1 and Lˆd,2, respectively). Meanwhile, the jump operators Lˆ2 and Lˆ4 are similar to single-qubit amplitude damping channels (Lˆa,1 and Lˆa,2, respectively) with the notable difference that there are a number of significant non-zero elements which correspond to elements σ+ I ˆ ˆ ∝ ⊗ (L2) and I σ+ (L4), consistent with the observation that the initial state of the qubits correspond to the ground ∝ ⊗ state of an anharmonic oscillator at finite temperature (see Eq. (S2) and (S7)). We also ran a restricted optimization with the jump operators fixed to correspond to single qubit dephasing and 5 amplitude damping noise (Eq. (S12)–(S15)). For this restricted optimization, we extracted a Hamiltonian

0.02 0.01+0.01i 0.01i 0.02 0.10i ˆ 0.01 0.01i − 1.00 0.01+0.09i 0.01− 0.10i H = ~ − 0.−01i 0.−01 0.09i − 0.24− 0.02+0−.07i (S16) 0.02+0− .10i −0.01+0− .10i 0.02− 0.07i 1.33  − −  with angular frequencies in units of 2π MHz and decay rates of γ = 0.03 MHz,γ = 0.04 MHz,γ = 0.16 MHz, × a,1 a,2 1,d and γ2,d = 0.11 MHz. The deviation between the restricted and free optimization, δ(t), as defined in the main text, is seen in Fig. S3. We note that the maximum likelihood was found for the unrestricted optimization. Extracting the unrestricted Hamiltonian for the two-qubit system, we find an estimated state-dependent frequency shift ωzz/2π = 416kHz due to the always-on ZZ-coupling [2]. As a check, the frequency shift ωzz can be independently estimated from device parameters using the relation [3]:

2g2 2g2 ω = + (S17) zz ∆ η ∆ η − B − − A where ηA, ηB are the anharmonicity of qubit A and B respectively, g is the coupling strength between the two qubits, and ∆ = ωA ωB is the frequency detuning between them. Substituting in the parameters for our device from i − i Table I, we estimate a state-dependent frequency shift ωzz/2π = 425kHz, consistent with the value found from the Hamiltonian extraction using LT.

APPENDIX D: SINGLE-QUBIT LT RESULTS AND DISCUSSION

The estimated jump operators for the single-qubit noise channel of qubit A with the neighboring qubit in the ground state are

ˆ 0.70+0.05i 0.01 0.14i ˆ 0.11 0.01i 0.22 0.89i L1 = 0.01+0.07i 0.70− 0.05i , L2= −0.22+0−.30i 0.11+0− .01i , (S18) − − with decay rates γ1 = 0.09 MHz and γ2 = 0.06 MHz. We note that these operators fit the data well, as shown in Fig. S4, and are similar to single-qubit dephasing (Lˆ1) and amplitude damping (Lˆ2) channels. For comparison, the jump operators of the single qubit dephasing (Lˆd) and amplitude damping (Lˆa) channels are of the form:

ˆ 1 0 ˆ 0 1 Ld 0 1 , La ( 0 0 ) , (S19) ∝ − ∝  Running LT on the full set of pre- and post-pulses for this particular dataset, we note that there is some small dis- agreement between the Lindblad fits (red) and the data (blue) for sequences corresponding to T2-type measurements. For example, we observe that the Lindblad fit slightly underestimates the decay time for the dataset when qubit A is prepared in + and measured in the x-basis (Fig. 3a in the main text, purple highlighted plot in Fig. S4). However, | i we note that there appears to have been a sudden change in the decay rate during the measurements corresponding to preparation in i and measurement in the x- and y-bases (which happened to be the final data taken during our |− i measurement loop and were measured sequentially), and we see that the Lindblad fit overestimates the decay time of these abnormally noisy datasets. Since LT finds the single time-independent Lindblad operator which best describes all combinations of pre- and post-pulses, this spurious temporal fluctuation in the channel during a small set of measurements constitutes a partial violation of Assumption 1 of LT [2], and this fluctuation will affect the fit of the other datasets. To confirm this, we rerun our Lindblad estimation on this dataset, first excluding the single dataset corresponding to preparation of the i -state and measurement in the y-basis (Fig. S5), and then excluding all datasets where qubit A is prepared |− i in i (Fig. S6). Excluding these datasets from the fit, we see that we find a Lindblad operator which even better |− i matches the remaining datasets. 6

1.0

0.5

0.0 1.0

0.5 -state Probability 0.0 1.0

0.5

0.0 0 40 80 0 40 80 0 40 80 0 40 80 0 40 80 0 40 80 Idling Channel Duration (μs) Raw Data Measurement Prediction from Kraus (Ramsey) Measurement Prediction from Lindblad

FIG. S4. Single-qubit LT results for the full set of initial states and measurement bases, performed on qubit A while qubit B is its ground state. The purple T2 plot is the same used in Fig. 3a in the main text. We emphasize that, while standard T1 and T2 metrics are determined by simply fitting the measurement results of a single set of pre- and post-pulses (pale green or purple, respectively), the Kraus operators (orange) and Lindbladian (red) here are determined from the full set of 18 measurement sequences (36 initial states 9 measurement bases = 324 sequences for two-qubit LT). ×

1.0

0.5

0.0 1.0

0.5 -state Probability 0.0 1.0

0.5

0.0 0 40 80 0 40 80 0 40 80 0 40 80 0 40 80 0 40 80 Idling Channel Duration (μs) Measurement Raw Data (Ramsey) Measurement Prediction from Lindblad Data Excluded from Lindblad Estimation

FIG. S5. Single-qubit LT, performed on qubit A while qubit B is its ground state. Here, we exclude the dataset corresponding to preparation of the i -state and measurement in the y-basis from our Lindblad estimation, and we observe an improved Lindblad fit for the remaining|− i datasets. 7

1.0

0.5

0.0 1.0

0.5 -state Probability 0.0 1.0

0.5

0.0 0 40 80 0 40 80 0 40 80 0 40 80 0 40 80 0 40 80 Idling Channel Duration (μs) Measurement Raw Data (Ramsey) Measurement Prediction from Lindblad Data Excluded from Lindblad Estimation

FIG. S6. Single-qubit LT, performed on qubit A while qubit B is its ground state. Here, we exclude the three datasets corresponding to preparation of the i -state from our Lindblad estimation, and we observe a significantly improved Lindblad fit for the remaining datasets. |− i

[1] Morten Kjaergaard, Mollie E. Schwartz, Ami Greene, Gabriel O. Samach, Andreas Bengtsson, Michael O’Keeffe, Christo- pher M. McNally, Jochen Braum¨uller,David K. Kim, Philip Krantz, Milad Marvian, Alexander Melville, Bethany M. Niedzielski, Youngkyu Sung, Roni Winik, Jonilyn Yoder, Danna Rosenberg, Kevin Obenland, Seth Lloyd, Terry P. Or- lando, Iman Marvian, Simon Gustavsson, and William D. Oliver, “Programming a quantum computer with quantum instructions,” (2020), arXiv:2001.08838 [quant-ph]. [2] See main text. [3] P. J. J. O’Malley, J. Kelly, R. Barends, B. Campbell, Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, A. G. Fowler, I.-C. Hoi, E. Jeffrey, A. Megrant, J. Mutus, C. Neill, C. Quintana, P. Roushan, D. Sank, A. Vainsencher, J. Wenner, T. C. White, A. N. Korotkov, A. N. Cleland, and John M. Martinis, “Qubit metrology of ultralow phase noise using randomized benchmarking,” Phys. Rev. Applied 3, 044009 (2015).