ACS Publications. Most Trusted. Most Cited. Most Read
Benchmarking Quantum Gates and Circuits
My Activity
  • Open Access
Review

Benchmarking Quantum Gates and Circuits
Click to copy article linkArticle link copied!

  • Vinay Tripathi
    Vinay Tripathi
    Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United States
    Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
  • Daria Kowsari
    Daria Kowsari
    Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United States
    Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
  • Kumar Saurav
    Kumar Saurav
    Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
    Ming Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United States
    More by Kumar Saurav
  • Haimeng Zhang
    Haimeng Zhang
    Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
    Ming Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United States
  • Eli M. Levenson-Falk*
    Eli M. Levenson-Falk
    Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United States
    Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
    Ming Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United States
    *E-mail: [email protected]
  • Daniel A. Lidar*
    Daniel A. Lidar
    Department of Chemistry, University of Southern California, Los Angeles, California 90089, United States
    Ming Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United States
    Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United States
    Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
    *E-mail: [email protected]
Open PDF

Chemical Reviews

Cite this: Chem. Rev. 2025, XXXX, XXX, XXX-XXX
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.chemrev.4c00870
Published May 5, 2025

© 2025 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Abstract

Click to copy section linkSection link copied!

Accurate noise characterization in quantum gates and circuits is vital for the development of reliable quantum simulations for chemically relevant systems and fault-tolerant quantum computing. This paper reviews a variety of key benchmarking techniques, including Randomized Benchmarking, Quantum Process Tomography, Gate Set Tomography, Process Fidelity Estimation, Direct Fidelity Estimation, and Cross-Entropy Benchmarking. We evaluate each method’s complexities, the resources they require, and their effectiveness in addressing coherent, incoherent, and state preparation and measurement (SPAM) errors. Furthermore, we introduce Deterministic Benchmarking (DB), a novel protocol that minimizes the number of experimental runs, exhibits resilience to SPAM errors, and effectively characterizes both coherent and incoherent errors. The implementation of DB is experimentally validated using a superconducting transmon qubit, and the results are substantiated with a simple analytical model and master equation simulations. With the addition of DB to the toolkit of available benchmarking methods, this article serves as a practical guide for choosing and applying benchmarking protocols to advance quantum computing technologies.

This publication is licensed under

CC-BY-NC-ND 4.0 .
  • cc licence
  • by licence
  • nc licence
  • nd licence
© 2025 The Authors. Published by American Chemical Society

Special Issue

Published as part of Chemical Reviews special issue “Quantum Computing”.

1. Introduction

Click to copy section linkSection link copied!

The realization of the promise of quantum computation (1,2) relies on the development of fault-tolerant, error-corrected quantum computers. Fault-tolerance is essential for the ultimate scalability of quantum computing, but requires high-fidelity operations that exceed an accuracy threshold. (3−7) When targeting chemical accuracy (1 kcal/mol) for electronic structure simulations, the impact of gate error becomes magnified because small inaccuracies in energies ─ especially for strongly correlated molecules ─ can alter reaction pathways and observables. Error probabilities per gate below 10–4 could be vital in large-scale fault-tolerant simulations of industrially relevant catalysts, as higher error rates dramatically inflate the number of logical qubits and error-correcting cycles required to achieve a final result with the desired precision. (8) Similarly, in quantum dynamics calculations, which often involve iterated time-step evolution, the accumulated error can become prohibitively large unless physical error rates are suppressed to the 10–5 – 10–6 regime. (9) In these scenarios, each additional order of magnitude in gate fidelity can translate into disproportionately lower qubit counts or reduced circuit depth, thereby expanding the range of molecular systems amenable to quantum simulation.
Characterizing quantum operations is thus crucial and plays a pivotal role in the development and implementation of universal quantum computers. Accurate characterization allows for the identification and subsequent improvement of errors in quantum gates. But even before the advent of universal, fault tolerant quantum computation, today’s noisy intermediate scale quantum (NISQ) computers (10) are being actively investigated in terms of quantum chemistry applications. (11,12) Early algorithms focused on exact diagonalization, (13,14) while much current attention is devoted to variational algorithms. (15,16) From a control-theoretic perspective, reducing gate errors ─ particularly multiqubit errors that can amplify coherent leakage or correlated noise channels ─ is indispensable for accurate electronic structure calculations. (17) Near-term algorithms such as the Variational Quantum Eigensolver (VQE) (18) and, eventually, more resource-intensive protocols like Quantum Phase Estimation (QPE), (19) hinge on efficient gate decompositions that preserve delicate coherence over increasingly large circuit depths. (20) Any deviation in gate fidelity can accumulate into significant numerical biases, thus compromising predictions of reaction energies or catalytic mechanisms. (21) Consequently, noise-reduction strategies have emerged as key tools in the quest to push current hardware toward the threshold of chemical accuracy. (22) For such techniques to function and improve performance, accurate characterization and benchmarking of quantum operations are essential.
Several methods have been developed over the years for benchmarking quantum operations, particularly focusing on quantum gates. These methods are part of the broader research field of Quantum Characterization, Verification and Validation (QCVV), (23) and include Quantum Process Tomography (QPT), (24−26) Randomized Benchmarking (RB) and its many variants, (27−30) Gate Set Tomography (GST), (31−33) Process Fidelity Estimation, (34) Direct Fidelity Estimation, (35,36) and Cross-Entropy Benchmarking. (37−39) QPT provides a full reconstruction of a quantum process, but is resource-intensive and scales poorly with system size. RB, on the other hand, offers a scalable alternative by estimating the average error rate of quantum gates, but it lacks the ability to identify specific error types. GST addresses some of these limitations by providing a self-consistent characterization of a complete set of quantum gates, albeit at the cost of increased experimental complexity.
Gate errors are fundamentally either coherent or incoherent. Coherent errors that arise from systematic deviations in quantum gate operations pose a significant challenge in the context of error correction. (40,41) While stochastic errors add up as probabilities, coherent errors add up as amplitudes, generally leading to a quadratically smaller accuracy threshold. (42) Traditional error characterization methods such as RB, which deliberately randomize noise maps, do not capture coherent error accumulation unless appropriately modified. (43) This can result in an underestimation of the gate performance needed to achieve accuracy thresholds, critical for achieving fault-tolerant quantum computing. (44,45) Various techniques exist that can mitigate the accumulation of coherent errors by converting them into stochastic noise, but only at the expense of significant additional cost in the number of required circuits. (46−49) It can be advantageous to, instead, characterize coherent gate errors, (50) and subsequently eliminate or at least suppress such errors at the control level, rather than placing the burden of doing so on fault-tolerant quantum error correction using large codes. (51)
In this Review, we provide a survey of quantum gate characterization methods, emphasizing the role of each strategy depending on the end goal. As the field is vast, we sacrifice generality somewhat by emphasizing the characterization of single-qubit and two-qubit gates, given that these are the fundamental building blocks of quantum circuits. (52) However, we also include methods designed to provide holistic benchmarks at the algorithm level.
After providing, in Section 2, a general introduction to the sources of gate errors and the formalism for characterizing them, we cover in Section 3, a variety of methods ranging from the most comprehensive and resource-intensive, such as quantum process tomography and gate set tomography, to the least detailed, such as randomized benchmarking. We discuss the trade-offs between these techniques, their scalability, and their applicability to various quantum computing architectures.
To keep this review focused, we refrain from providing a comprehensive analysis of all available benchmarking methods. Consequently, we do not discuss several significant methods, such as cycle benchmarking, (53) leakage and crosstalk characterization, (54) quantum volume benchmarking, (55) fidelity estimation with 2-designs, (56) mirror benchmarking, (57) variational quantum benchmarking, (58) average gate fidelity, (59) machine learning-based benchmarking, (60,61) channel spectrum benchmarking, (62) and error per layered gate (EPLG), (63) among others. Most of these methods are variants of the key methods we focus on here, so we hope that the reader will feel well-prepared to tackle these other methods after reading this review.
However, this work is more than a review, as we also introduce an original benchmarking protocol that we call deterministic benchmarking (DB). This new protocol, which is appropriate for characterizing single-qubit gates, requires fewer experimental runs than previous characterization methods, exhibits robustness to SPAM errors, and provides detailed information about coherent errors. We discuss DB in detail in Section 4, provide experimental evidence of its efficacy, and contrast it with RB.

2. Gate Errors

Click to copy section linkSection link copied!

2.1. Quantum Dynamical Maps

Assuming a factorized initial system-bath state, i.e., a system which is not entangled with its bath, the dynamics of an open quantum system with a D-dimensional Hilbert space H can be described by a completely positive trace-preserving (CPTP) linear map: (64)
E(ρ)=iAiρAi
(1)
where ρB(H) (the space of linear operators acting on H) is the initial state of the system and satisfies Tr(ρ) = 1 as well as ρ ≥ 0 (its eigenvalues are all non-negative). The Kraus operators (65) Ai satisfy ∑iAiAi = I (the identity operator), which guarantees that Tr[E(ρ)]=1. Let {Em}m=0D2–1 be a set of fixed Hermitian basis operators for B(H), which satisfy the orthogonality condition
Tr(EmEn)=Dδmn
(2)
For example, we may choose E0 = I and Em, m = 1, ..., D2 – 1, to be traceless Hermitian matrices. For a multiqubit system, the Em’s can be tensor products of the identity operator and the Pauli matrices, which we denote interchangeably by {σx, σy, σz} or {X, Y, Z}. We also write Pauli tensor products interchangeably as σx ⊗ σy or XY, etc.
The Ai operators can be decomposed as Ai = ∑maimEm, and therefore we have
E(ρ)=m,n=0D21χmnEmρEn
(3)
where χmn = ∑iain*aim = χnm*. The “process matrix” χ is not only Hermitian but also positive semidefinite (all its eigenvalues are non-negative). It captures all the information about the map E with respect to the {Ei} basis, i.e., a characterization of E is equivalent to a determination of all the matrix elements of χ, where the Ei play the role of observables. Since χ is Hermitian and the map E is trace-preserving, the corresponding process matrix χ has at most D4D2 independent real parameters [see eq 41 below]. We will focus most of our attention on the n-qubit case, i.e., D = 2n.
The measurement formalism of quantum mechanics is conveniently formulated using the language of CPTP maps as well. (66) A set of generalized measurement operators {Ai} is any set in B(H) that satisfies the constraint ∑iAiAi = I. When a system in the state ρ is measured, the postmeasurement state is ρi = AiρAi/pi with probability pi = Tr(AiρAi) = Tr(Miρ), where the set {Mi} = {AiAi} is known as a positive operator-valued measure (POVM). If the measurement outcome is not known, the resulting state becomes ∑iAiρAi.
The standard case of projective von-Neumann measurements is the special case when the generalized measurement operators are projectors, i.e., are Hermitian and satisfy AiAj = δijAi. When an observable (any Hermitian operator) Ω with spectral decomposition Ω = ∑kλkPkk and Pk are its eigenvalues and eigenprojectors, respectively) is measured, the postmeasurement state is ρk = PkρPk/pk with probability pk = Tr(Pkρ), and the measurement outcome is the eigenvalue λk. The state resulting from ignorance of the measurement result is then ∑kPkρPk, which generally corresponds to a mixed state, i.e., a state σ whose purity Tr(σ2) is less than 1.
The Choi-Jamiołkowski isomorphism (67,68) is a correspondence between maps acting on states in B(H) and states in HH. More concretely, the Choi state ρE of a quantum map E is
ρE=(IE)(|ϕϕ|)
(4)
where |ϕ=1Di=1D|i|i is a maximally entangled state. The Uhlmann fidelity (69) between two states ρ,σB(H) is
F(σ,ρ)=ρσ1=Tr((σ1/2ρσ1/2)1/2)
(5)
where the norm is the trace-norm: A1=Tr|A|=Tr(AA). The Uhlmann fidelity is bounded between 0 and 1 and is unitarily invariant. Note that some authors define it as the square of our quantity.
The process fidelity F between two quantum maps E and E, acting on the same quantum system, with respective Choi states ρE and ρE, is given by the Uhlmann fidelity between the Choi states:
Fproc(E,E)=F(ρE,ρE)=ρEρE1
(6)
Fproc quantifies the overlap between the two maps’ actions on all possible input states. A process fidelity value close to 1 indicates that the maps are very similar in their effect on quantum states. This metric is very useful for quantifying the similarity between a desired process, such as an ideal gate, and its actual, noisy implementation.

2.2. Coherent and Incoherent Errors

A quantum gate aims to produce some unitary operation U which can be written as a superoperator U acting on the state ρ, i.e., U(ρ)=UρU. In practice, the actual operation is U~=EU, where E now represents all errors on the gate. These errors can be divided into coherent errors, where state purity is preserved but the unitary is altered, and incoherent (stochastic) errors, where state purity [Tr(ρ2) = 1] is lost due to some randomness in the evolution (i.e., decoherence). Put another way, coherent errors are deterministically repeatable, while incoherent errors give random evolution. Coherent errors are generally more damaging, as their amplitudes add in phase and can thus increase the gate failure probability quadratically faster than incoherent errors. However, coherent errors are always reversible in principle. Incoherent errors, in contrast, require quantum error correction to become reversible. (70) Next, we present an overview of common incoherent and coherent errors in single- and two-qubit gates.

2.3. Single-Qubit Gate Errors

In general the map E is very unconstrained. For a single qubit, i.e., for D = 2, this allows D4D2 = 12 real parameters in a CPTP map, and even more if we allow for leakage outside the qubit space (see below). On the other hand, the physical causes and impacts of error maps are more constrained. For a single qubit, the state ρ can be represented by the Bloch vector v ≡ (⟨σx⟩, ⟨σy⟩, ⟨σz⟩), with the length of the vector giving the state purity. A general single-qubit unitary can be parametrized as R(ϑ, θ, ϕ), a rotation by an angle ϑ about an axis with polar and azimuthal angles θ, ϕ. Coherent errors in the space of qubit states can then be categorized as
Rotation errors, where the value of ϑ is changed from the target. These may be the result of drifts in the amplitudes of control drives or the coupling of the qubit to these drives.
Phase errors, where the axis of rotation is altered─that is, θ and/or ϕ are changed from the target. These may result from drifts in the qubit frequency leading to unwanted z-axis rotations, or drifts in the control axis (often caused by a drift in the phase of an oscillatory control field).
Similarly we can characterize incoherent errors as due to
  • Relaxation from the qubit state to the equilibrium state. These bit-flip errors are typically modeled as Markovian processes leading to an exponential decay with characteristic time T1. Importantly, relaxation at finite temperature leads to a nonzero state purity as the system relaxes toward its ground state.

  • Dephasing due to noise in the qubit frequency relative to a reference (where the noise may originate from the qubit or the reference). These phase-flip errors may have both Markovian and non-Markovian components, and, in many condensed matter or solid-state systems, are typically modeled as a Markovian error plus a classical noise with a 1/f power spectrum. (71)

It is also important to consider leakage errors, which result from the state evolving outside the qubit space. This error is coherent as long as it preserves state purity (thus allowing for its reversal (72)), although it will always appear incoherent if we restrict ourselves to only considering the qubit space. This is the case in, e.g., the transmon superconducting qubit, (73) where the qubit is encoded in the lowest two states of a weakly anharmonic oscillator. Driving this transition can also drive higher transitions, leading to coherent rotations in these state manifolds. Importantly, repeated application of a gate with a coherent leakage error can cause a rapid buildup of errors that can be difficult to address.
Modeling leakage in the qubit space requires a non-trace-preserving map, as population leaves the space and is transferred to the leakage state. Such maps can be modeled by modifying the CPTP map formalism (Section 2.1) and replacing the equality in the ∑iAiAi = I constraint by ≤ I. The corresponding process matrix χ then has at most D4 independent real elements (16 for a single qubit).
A final error type, erasure, is a special type of leakage error where qubit information is lost with some characteristic signature that an error has occurred. This is the case in, e.g., dual-rail qubit encodings where single-excitation loss takes both qubit states to a single leakage state and thus erases all information, (74,75) or trapped atom qubits where an atom may escape the trap. (76) Because erasure scrambles qubit information it must be considered as an incoherent error. Some error correction architectures can correct erasure errors with far fewer resources than bit-flip or phase-flip errors. It can therefore be preferable to use encodings which transform other errors into erasures, (77) as is done with dual-rail qubits. On the other hand, in many common qubits such as the transmon, erasure errors are not present and leakage must be considered as a coherent effect. (78,79)
Incoherent errors are sometimes lumped together and modeled as a single depolarizing map
E(ρ)=(1p)ρ+13p(XρX+YρY+ZρZ)
(7)
where p is the probability of a qubit error. However, this both obscures their physical origin and does not account for the effects of biased noise (80−86) and finite temperature (see Section 4.8).
If we ignore detection of erasure errors, and limit ourselves to Markovian decoherence, we see 7 parameters for qubit gate error: errors in ϑ, θ, ϕ, relaxation rate, equilibrium population, dephasing rate, and leakage rotation. Some of these parameters (e.g., equilibrium population) may be independent of the gate itself, and so need not be characterized for every gate. Proper treatment of non-Markovian errors can be much more complicated (87,88) and depends on the circumstance, but can often be captured with relatively simple models. (71,89,90)

2.4. Two-Qubit Gate Errors

The set of two-qubit gates is much larger and more diverse than the simple rotation gate for a single qubit, though it is amenable to an elegant geometrical classification. (91) It is therefore more practical to consider the error maps for a particular two-qubit operation, and even for a particular physical implementation of that operation, rather than for general two-qubit operations. For instance, consider the parametrically activated iSWAP gate in transmon qubits: (92) an oscillatory flux drive causes oscillations of a tunable transmon’s frequency about some maximum, bringing the average frequency into resonance with another transmon so the two can exchange energy (possibly via an intermediate tunable transmon coupler). (93) Error maps include Markovian relaxation and dephasing of each qubit; non-Markovian dephasing of each qubit, typically modeled as a 1/f qubit frequency noise spectrum (note that the oscillatory nature of the drive acts as an “auto-echo” that eliminates much of this dephasing); two-qubit couplings to a Markovian or non-Markovian bath, e.g., a noisy tunable coupler; incorrect drive amplitude leading to transmon-transmon detuning and so causing incomplete swapping; two-qubit phase errors from unwanted ZZ interaction due to higher transmon levels; single-qubit phase errors due to improper calibration of level repulsion during transients at the beginning and end of the gate; and leakage into higher transmon levels or coupler states. Every other implementation of iSWAP can have its own error maps, to say nothing of other two-qubit operations.
The complexity and diversity of two-qubit gate error maps mean it is best to approach them holistically on a case-by-case basis, e.g, by developing an “error budget” for the different sources of error. (93,94) Another common approach is to characterize the gate completely once, using quantum process tomography or some other complete method. This identifies relevant error maps, which may then be measured more efficiently without needing to fully characterize the quantum process over and over. Typically the relevant coherent errors are characterized via some form of matrix element amplification, where repeated applications of the gate on a particular initial state build up a particular coherent error. There is a danger that this amplification will also amplify single-qubit coherent errors that may be difficult to distinguish. Adding dynamical decoupling sequences in between the two-qubit gate applications can help suppress these single-qubit errors that may confound the analysis. (50)
Incoherent errors in two-qubit gates are typically characterized by performing single-qubit coherence measurements, then using some error model (usually Markovian) to predict the impact of single-qubit decoherence. Unfortunately, two-qubit incoherent errors due to joint bath couplings are often ignored unless full process tomography is being performed. This simplification can be shown to give good predictions in many contexts, (95) but in some circumstances two-qubit incoherent errors may be quite significant. For instance, the two qubits are often coupled via a third system (such as a tunable coupler) that can act as a joint bath. Fortunately, it is straightforward to extend standard single-qubit coherence measurements─for instance, joint relaxation can be characterized by comparing the relaxation rates of the |10⟩, |01⟩, and |11⟩ states. (96)
Properly benchmarking two-qubit gates is a significant challenge, and the best approach depends on the exact goal: calibrating a gate, predicting algorithmic performance, characterizing the overall quality of a quantum processor, etc. We describe and compare the various single- and two-qubit gate benchmarking protocols below.

3. Gate and Circuit Benchmarking Protocols

Click to copy section linkSection link copied!

A gate benchmarking protocol produces a set of numbers for an input gate or gate set, which can be used to compare the performance in different instances of gate implementation. Such a protocol can be judged on the basis of a few main criteria:
• Comprehensiveness, i.e., how much information about the error does the protocol provide. One simple way to quantify the amount of information provided is the number of parameters that end up describing the error.
• Specificity, i.e., what assumptions are made about the noise model in the protocol. Some common assumptions are that errors are independent of each other and weakly depend on the gate being applied, (97,98) or that they are of a particular type, such as Pauli errors. (99)
• Simplicity, i.e., how easily the scheme can be adapted to existing experimental setups. This includes how complicated the pulse control and measurement schemes are for the experiment as well as how sensitively the experiment must be calibrated.
• Resource costs, i.e., how expensive it is to run the protocol. Note that scalability, which is a critical consideration in quantum algorithms and processes, can become less relevant for a gate benchmarking protocol if we restrict the characterization effort to single one- or two-qubit gates in isolation. The number of such gates is at most n(n – 1)/2 for n fully connected qubits. Instead, the absolute resource cost of the protocol becomes important and needs to be optimized.
Apart from these criteria, a gate benchmarking protocol should also be robust against all errors, apart from the ones it is trying to quantify. Critically, for most protocols, this robustness should include state preparation and measurement (SPAM) errors, (100) which are among the most pervasive errors in experiments. (101) Exceptions to this rule include protocols that quantify the SPAM errors themselves.

3.1. Randomized Benchmarking

Randomized Benchmarking (RB) aims to give the most concise characterization of gate error by reporting a single number to quantify the average gate fidelities, while suppressing the effects of SPAM errors. (27−29,102) Over the past two decades, RB has become a standard tool used across different platforms to quantify the performance of single- and two-qubit gates. Many variants of RB have been introduced; here, we focus on the two most prominent variants, known as standard RB and interleaved RB. The main benefits of RB methods include reasonable assumptions, simplicity of implementation compared to process tomography, and relatively low resource cost. However, RB reduces the characterization to a single number. Indeed, this is the main reason for the method’s relative simplicity.

3.1.1. Standard RB

Standard RB characterizes a specific selection of gates, which are elements of the Clifford group (see below), by an average gate infidelity metric. The key idea behind standard RB is to transform errors into an effective depolarizing error [recall eq 7] via a process called “twirling”, which simplifies the analysis. The process is described in detail below.
The Pauli group Pn is the group of all possible n-fold tensor products of the three Pauli matrices and the identity matrix. The Clifford group consists of the set of gates that preserve the Pauli group under conjugation. I.e., C is in the Clifford group if CPCPn for all PPn.
The standard RB protocol involves the following steps:
  • Preparation: Initialize the n-qubit quantum system in a reference state |0⟩n, typically the ground state.

  • Random Gate Sequence: Apply a sequence of m randomly selected Clifford gates {Ci}i = 1m:

    Um=CmCm1···C1
    (8)

  • Inversion Gate: Apply the inverse of the accumulated operation to ideally return the system to the initial state:

    Cinv=Cm+1=Um=C1···Cm
    (9)
    The inverse can be calculated efficiently classically. (103)

  • Measurement: Measure the system in the computational basis (the eigenbasis of the Pauli σz matrix) and record whether it has returned to the initial state.

  • Averaging: Average the measurement outcomes over K different random sequences for a given m to obtain the survival probability P(m). Repeat the process for various sequence lengths m to get a sequence of survival probabilities {P(m)}.

  • Fitting to exponential: The final step is to fit the measured survival probability to an exponential function Apm + B in order to extract p, which can be used to quantify the average gate fidelity of the Clifford gate sequences. Here, A and B capture the SPAM errors.

3.1.2. RB Theory

A randomly selected Clifford gate C can be represented by the ideal superoperator G(·)=C(·)C, and the associated error process is modeled as G~=EG, where the error E is assumed to be deterministic. The total process map I~m for the sequence of m + 1 Clifford gates (including the last one which inverts the previous m gates) is given by
I~m=G~m+1G~m···G~1
(10)
Ideally, I~m should have been the identity operation. Assuming gate-independent noise i.e., EiEi, we obtain
I~m=(EGm+1)(EGm)···(EG1)
(11)
which can be written as
I~m=Ei=1m(Di1EDi)
(12)
where Di=k=1iGk and we have used Πk=1m+1Gk=I. This is the process generated by one of the realizations; repeating the same depth-m sequence K times results in an average process defined as
I¯m1KjI~m(j)=E1Kji=1m(Di(j)1EDi(j))
(13)
Here the error map E is “twirled” over the Clifford group, and it can be shown (104) that this reduces it to an m-fold depolarizing map Edepm which replaces a state ρ by the maximally mixed state I/D with probability 1 – pm, and leaves the state unchanged with probability pm. We note that this process of applying random unitaries that leave only a Markovian contribution is closely related to the earlier idea of randomized dynamical decoupling. (105−107)
The average survival probability of the initial state ρ0 = (|0⟩⟨0|)n is then given by
Pavg(m)=Tr[ρ0I¯m(ρ0)]=Tr[ρ0EEdepm(ρ0)]=Tr[ρ0E(pmρ0+(1pm)ID)]=Apm+B
(14)
where
A=Tr[ρ0E(ρ0ID)]
(15)
and
B=Tr[ρ0E(ID)]
(16)
The key assumptions include gate-independent and time-independent errors in the above derivation. Pavg(m) from eq 14 is used to fit the experimental decay curves to extract p. Since twirling generates a depolarizing noise map of the form Edep(ρ)=pρ+(1p)ID, its average gate fidelity with respect to an identity can be calculated to be F¯Edep,I=p+(1p)/D, where the depolarization parameter p is identified with the experimentally measured exponential decay parameter. Therefore, we can relate the average Clifford gate infidelity rC=1F¯Edep,I to the experimental decay parameter p obtained from the RB experiments as
rC=D1D(1p)
(17)
Since rC depends only on the depolarization parameter p of the effective process matrix generated by the gate, RB becomes insensitive to SPAM errors, which are captured by A and B.
If we consider the case when the errors are gate-dependent as well as time-dependent, a perturbation theory calculation can be performed to extract the higher order corrections. In particular, the first order correction yields (29,102)
Pavg(1)(m)=A1pm+B1+C1(m1)(qp2)pm2
(18)
where A1, B1 and q are functions of m. The term (qp2) is a measure of the degree of gate-dependence in the errors. It can be shown that the second order term can be neglected when γ ≪ 2/m, where γ is a measure of the noise variation. (29) The zeroth order approximation [see eq 14] provides a good estimate for small errors and the first order correction may be needed in cases where the errors are large. Note that even if the gate dependence term qp2 is large, it would still be possible to fit the experimental data with eq 18 as long as m is large enough. However, this can conflict with the γ ≪ 2/m condition. In addition, it is also necessary for m to be large enough to have enough data points for a reasonable estimate of p in the fitting model. These considerations highlight some of the restrictions of using RB for gate and time-dependent noise.

3.1.3. Interleaved RB

Whereas RB provides an averaged fidelity for a random set of Clifford gates, Interleaved RB (IRB) (108) allows the estimation of the fidelity of a specific Clifford gate C. The key idea is to analyze the effect of inserting the gate of interest G (represented below by a superoperator G) into an RB sequence. The IRB protocol consists of two different experiments: the first is the standard RB as discussed in Section 3.1.1, while the second is an RB sequence with G interleaved between every pair of consecutive random Clifford gate. More concretely, the sequence of gates in the second experiment after interleaving is
Um,inter=GmGGm1···GG2GG1
(19)
The above sequence is then followed by Gm+1=Um,inter1. Similar to standard RB, this also generates an exponential decay characterized by a parameter pinter. It can then be shown (108) that using the two decay parameters p and pinter from the two experiments, we can estimate the gate infidelity contribution of the particular gate G using
rG=D1D(1pinterp)
(20)
We note that IRB is not always a reliable method for estimating a gate’s infidelity because unitary errors in the interleaved gate G can coherently add or cancel with errors in the random Clifford gates, sometimes even resulting in a negative error rate. (23) Another way to see the problem with IRB is via eq 20 with the interleaved decay slower than the decay of standard RB (pinter > p), which results in a negative infidelity rG. This, in turn, implies an unphysical average gate fidelity larger than 1. Such issues lead to potentially large discrepancies between the estimated and true fidelity.
In fact, it can be shown (108) that the original error rate for the Clifford gate of interest G must lie in the range [rGE,rG+E] where
E=min{(D1)[|ppinter/p|+(1p)]D2(D21)(1p)pD2+41pD21p
(21)
which arises from imperfect random gates. When the noise channel is fully depolarizing (p = 1) then E = 0, but in general E > 0. This means that rG+E can exceed 1, which is unphysical. When this happens, IRB results must be treated with caution. Possible remedies include increasing the number of random sequences and the number of circuit repetitions, and checking for systematics in how errors are introduced (e.g., calibration or readout errors might be overshadowing gate error in unexpected ways).
Despite its shortcomings, randomized benchmarking is a fundamental tool in quantum computing for quantifying average gate fidelities with robustness to SPAM errors. By leveraging the mathematical properties of the Clifford group and statistical averaging, RB provides a scalable method for characterizing quantum processors. Ongoing research continues to refine RB protocols and extend their applicability to more general settings. IRB is just one possible extension of RB; there is a plethora of other work focusing on other generalizations of randomized benchmarking. These include simultaneous RB (109) (which detects the effects of simultaneous gate operations), leakage RB, (110,111) non-Markovian RB, (112,113) mirror RB, (114) character RB (115,116) (which goes beyond the Clifford group), and more.

3.2. Quantum Process Tomography

If randomized benchmarking sacrifices some comprehensiveness and specificity in service of simplicity and low resource cost, quantum process tomography (QPT) does the opposite. This method seeks to fully characterize a quantum process E, instead of providing a single number as in RB. Since D4D2 real numbers are required to fully specify a CPTP map E, QPT is, of course, much more resource intensive than RB.

3.2.1. Standard Quantum Process Tomography

In “standard” quantum process tomography (SQPT), illustrated in Figure 1(a), we prepare D2 linearly independent inputs {ρk}k=0D2–1, where each ρkB(H), and then measure the output states E(ρk) using quantum state tomography. (24,25) Since the map E [eq 1] is linear, it can be in principle reconstructed from the measured data by proper inversion. Let {ρk}k=0D2–1 be a linearly independent basis set of operators for B(H). A convenient choice is ρk = |m⟩⟨n|, where {|m⟩}m=0D–1 is an orthonormal basis for H and k = (m, n). The coherence |m⟩⟨n| can be constructed from four populations:
|mn|=|+mn+mn|+|mnmn|1+i2(|mm|+|nn|)
where |±mn=(|m±|n)/2. The linearity of E then implies that a measurement of E(|+mn+mn|), E(|mnmn|), E(|mm|), and E(|nn|) suffices for the determination of E(|mn|). In addition, the channel output E(ρk) itself can be expressed in terms of a linear combination of basis states, as E(ρk)=lλklρl, where the parameters
λkl=Tr(ρkE(ρl))
(22)
contain the measurement results. Similarly, each summand in the RHS of eq 3 can be expanded in the basis states as EmρkEn = ∑lBmn,lkρl. The resulting equation then reads
E(ρk)=mnlχmnBmn,lkρl=lλklρlmnBmn,lkχmn=λkl
(23)
where the second line follows from the linear independence of {ρk}. This relation can in turn be written in the following matrix form:
Bχ=λ
(24)
where the D4 × D4-dimensional matrix B is determined by the choice of operator bases {ρk} and {Em}, and the D4-dimensional vector λ is determined from state tomography experiments. The process matrix χ can thus be determined by inversion of eq 24.

Figure 1

Figure 1. Three different QPT methods. (a) Schematic of SQPT. An ensemble of states {ρj} is prepared and each state is subjected to the map E followed by measurements {Em}. (b) Schematic of separable AAPT. A “faithful” input state ρ is subjected to the map EI. The operators {Ej} are measured on the joint system, which results in the required joint probability distributions or expectation values. (c) Schematic of DCQD. The system and the ancilla are prepared in one of the input states as in Table 1 and are subjected to the map EI. The joint system is measured in the Bell-state basis (Bell State Measurement).

In general, SQPT involves the preparation of D2 linearly independent inputs {ρl}, each of which is subjected to the quantum process E, followed by quantum state tomography on the corresponding outputs. For each ρl, we must measure the expectation values of the D2 fixed-basis operators {Ek} in the state E(ρl). Thus, the total number of measurements required is D4. Since measurement of an expectation value cannot be done on a single copy of a system, whenever we use the term “measurement” we implicitly mean measurement on an ensemble of identically prepared quantum systems corresponding to a given experimental setting.

3.2.2. Ancilla-Assisted Process Tomography

There is a duality between quantum state tomography schemes and QPT. This duality is based upon the Choi-Jamiolkowski isomorphism (67,68) we discussed in Section 2.1, which establishes a correspondence between CP maps and states. This one-to-one correspondence enables all theorems about quantum maps to be directly derived from those of quantum states. (117) In particular, it allows one to represent a quantum process as a quantum state in a larger Hilbert space. For example, choosing {Em} = {|i⟩⟨j|} results in χ=DρE, where ρE is the Choi state of the CPTP map E. Therefore, QPT has a natural reduction to quantum state tomography, and state identification methods can be applied to the characterization of quantum processes as well. The “ancilla-assisted process tomography” (AAPT) (118−120) scheme is based on this observation and is schematically illustrated in Figure 1(b).
The AAPT scheme involves attaching an ancilla B to the principal system A and preparing the combined system in a particular joint state so that complete information about E can be imprinted on the final state. Then, by performing quantum state tomography in the extended Hilbert space HAB=HAHB, one can extract complete information about the unknown map E acting on A. In principle, the input state of the system and ancilla can be prepared in either an entangled mixed state (entanglement-assisted) or a separable mixed state. The only requirement for the input state in AAPT is that it must be “faithful” to the map E, i.e., performing quantum state tomography on the output identifies E completely and unambiguously. (121) The faithfulness condition implies that a state ρ can be used as input for AAPT iff it has maximal Schmidt number D2, which is essentially an invertibility condition. By performing state tomography on the output state (EI)(ρ), we obtain a set of linear relations among possible measurement outcomes and the elements of E and ρ; the choice of ρ must be such that an inversion is possible.
Maximal entanglement is not a necessary property of the input state ρ in AAPT. However, there is experimental evidence that maximally entangled pure states offer optimal performance in the sense of reducing experimental error. (120) Indeed, experimental error amplification is related to inversion of the matrix involving E and ρ, which, in turn, depends on the magnitude of the inverse of the eigenvalues of ρ. Thus, a smaller eigenvalue leads to an amplified experimental error. As a consequence, the optimal (in the sense of minimal experimental errors) faithful input states are maximally entangled pure states.
One can realize the required quantum state tomography via separable measurements (separable AAPT), i.e., joint measurement of tensor product operators, or via collective measurements on both the system and the ancilla (nonseparable AAPT). Both types of measurements are performed on the same Hilbert space HAB. Furthermore, by adding another ancilla it is possible to perform a generalized measurement on the joint system AB. It turns out that the nonseparable measurement schemes (whether in the same Hilbert space or in a larger one) hardly offer any practical advantage because they require many-body interactions which are experimentally unrealizable. (26) We thus focus on joint separable measurements.
Let us assume that the initial state of the system and the ancilla is ρAB = ∑ijρijEiAEjB, where {EmA} ({EnB}) is the operator basis for the linear operators acting on HA (HB). The output state, after the action of the unknown map E on the principal system, is
ρAB=(EAIB)(ρAB)=kjα~kjEkAEjB
(25)
where α̃kj ≡∑mniχmnρijαkm,i,n, with αkm,i,n defined via EmAEiAEnA = ∑kαkm,i,nEkA. Note that αkm,i,n depends only on the choice of operator basis. The result of measuring the observables EkAEjB is
α~kj=Tr(ρAB(EkAEjB))
(26)
Now, let
χ~ki=mnαkm,i,nχmn
(27)
where the process matrix χ is the primary objective of QPT [recall eq 3] and the α parameters are known from the choice of operator basis. Thus, the procedure to obtain χ simplifies to two matrix inversions: after measuring {α̃kj}, we obtain χ̃ by inverting the following matrix equation:
α~=χ~ϱ
(28)
Using χ̃, χ is then found by inverting eq 27.

3.2.3. Direct Characterization of Quantum Dynamics

An alternative to SQPT and AAPT is “direct characterization of quantum dynamics” (DCQD). (26,122−126) In DCQD, similarly to AAPT, an ancilla system is utilized, but in stark contrast, it does not require inversion of a D2 × D2 matrix; hence, “direct”. The main idea in DCQD is to use certain entangled states as inputs and to perform a simple error-detecting measurement on the joint system-ancilla Hilbert space. The choice of input states and measurements is such that the measurement results directly encode elements of the quantum map, which removes the need for state tomography. In the case of a single qubit, the measurement scheme turns out to be equivalent to a Bell state measurement (BSM). In DCQD the choice of input states is dictated by whether diagonal or off-diagonal elements (quantum dynamical population and coherence, respectively) of the process matrix are to be determined. Population characterization requires maximally entangled input states, while coherence characterization requires nonmaximally entangled input states. We review the scheme for the case of a single qubit below.
Let us first demonstrate how to determine all diagonal elements of the process matrix, {χmm}, in a single (ensemble) measurement. We choose {I, XA, Y A, ZA} as our operator basis acting on the principal qubit A. We first prepare the maximally entangled Bell state |Φ+=(|00+|11)AB/2, and then (as in AAPT) subject only qubit A to the map E.
A stabilizer code is a subspace HC of the Hilbert space of n qubits that is an eigenspace of a given Abelian subgroup S that excludes −I (the stabilizer group) of the n-qubit Pauli group Pn, with eigenvalue +1. (127) In other words, for every |ΨCHC and all SiSPn, we have SiC⟩ = |ΨC⟩, where [Si, Sj] = 0 ∀i,j. Consider the action of an arbitrary (error) operator EPn on a stabilizer code state |ΨC⟩. Such an error is detectable if it anticommutes with (at least one of) the stabilizer generators: {Si, E} = 0. To see this note that
Si(E|ΨC)=E(Si|ΨC)=(E|ΨC)
(29)
i.e., EC⟩ is a –1 eigenstate of Si. Hence, measuring Si detects the occurrence of an error or no error (the –1 or +1 outcomes respectively). Measuring all the stabilizer generators then yields a list of errors (the “syndrome”), which allows one to detect the errors that took place. More precisely, each stabilizer code has a distance parameter d. The code detects all errors whose weight (number of nonidentity operators) is at most d – 1. E.g., a distance d = 3 code can detect all one- and two-body errors, i.e., errors of the form XII, IYZ, etc., but not weight-3 errors such as XYZ.
The Bell state |Φ+⟩ is a +1 eigenstate of the commuting stabilizer generators ZAZB and XAXB, i.e., it is stabilized under their action. The nonidentity error operators in the operator basis set {EmA}m=03 = {I, XA, YA, ZA} anticommute with at least one of the stabilizer generators, and therefore by measuring these generators we can detect the error. Note that, as mentioned above, measuring ZAZB and XAXB is equivalent to a BSM: it can be represented by the four projection operators PΨ± = |Ψ±⟩⟨Ψ±| and PΦ± = |Φ±⟩⟨Φ±|, where |Φ±=(|00±|11)/2, and |Ψ±=(|01±|10)/2 are the Bell states. Expressing the map on qubit A as E(ρ)=m,nχmnEmAρEnA, it is easily verified that the probabilities of obtaining the no-error outcome I and the errors XA, ZA, and YA can be written as
pm=Tr[Pm(EI)(|Φ+Φ+|)]=χmm
(30)
where m = 0, 1, 2, 3, and the projectors Pm, for m = 0, 1, 2, 3, correspond to the states Φ+, Ψ+, Ψ, and Φ, respectively. For example:
Tr[P0(EI)(|Φ+Φ+|)]=m,n=03χmnTr[PΦ+(EmAI)PΦ+(EnAI)]=m,n=03χmnδm0δn0=χ00
(31)
Eq 30 shows that the diagonal elements of the process matrix are directly obtainable from an ensemble BSM; this is the core observation that leads to the DCQD scheme. In particular, one can determine the quantum dynamical populations, χmm, by simultaneously measuring the operators ZAZB and XAXB on multiple copies of the state |Φ+⟩.
To determine the quantum dynamical coherence elements, χmn, a different strategy is needed. As the input state we take a nonmaximally entangled state: |Φα+⟩ = α|00⟩ + β|11⟩, with |α|,|β|{0,1/2} and Im(αβ*) ≠ 0. The sole stabilizer of this state is ZAZB. It is simple to verify that by measuring ZAZB on the output state E(ρ), with ρ = |Φα+⟩⟨Φα+|, we obtain the following probabilities for ±1 outcomes:
Tr[P+1E(ρ)]=χ00+χ33+2Re(χ03)ZATr[P1E(ρ)]=χ11+χ22+2Im(χ12)ZA
(32)
where P+1 = PΦ+ + PΦ, P–1 = PΨ+ + PΨ and ⟨ZA⟩ = Tr(ρZA) ≠ 0 because of our choice of a nonmaximally entangled input state. The experimental data, Tr[P+1E(ρ)] and Tr[P1E(ρ)], are exactly the probabilities of no bit-flip error and a bit-flip error on qubit A, respectively. Since we already know the {χmm} from the population measurement, we can determine Re(χ03) and Im(χ12). After measuring ZAZB the system is in either of the states ρ±1=P±1E(ρ)P±1/Tr[P±1E(ρ)]. Next we measure the expectation value of an element of the normalizer N(S) of the stabilizer group S: N(S)={gPn:[S,g]=0} (the normalizer preserves the stabilizer under conjugation). For example, we can measure N = XAXB, and obtain the measurement results
Tr[Nρ+1]=(χ00χ33)N+2iIm(χ03)ZANTr[P+1E(ρ)]Tr[Nρ1]=(χ11χ22)N2iRe(χ12)ZANTr[P1E(ρ)],
(33)
where ⟨ZA⟩, ⟨N⟩, and ⟨ZAN⟩ are all nonzero and already known. eqs 32 and 33 provide a linear system that can be straightforwardly inverted to extract the four independent real parameters needed to calculate the coherence elements χ03 and χ12. It is simple to verify that a simultaneous measurement of the stabilizer and normalizer elements ZAZB and XAXB, is again a BSM.
In order to complete the measurement of the remaining coherence elements, we perform an appropriate basis change by preparing the input states HAHBα+⟩ and SASBHAHBα+⟩, which are the eigenvectors of the stabilizer operators XAXB and YAYB. Here H and S are single-qubit Hadamard and phase gates acting on the systems A and B. At the output, we measure the stabilizers and a corresponding normalizer element, e.g., ZAZB, which are again equivalent to a standard BSM, and can be expressed by measuring the Hermitian operators PΦ+ ± PΨ+ and PΦ ± PΨ (for the input state HAHBα+⟩), and PΦ+ ± PΨ and PΦ ± PΨ+ (for the input state SASBHAHBα+⟩). Figure 1(c) illustrates the DCQD scheme.
Overall, in DCQD we only need a single fixed measurement apparatus capable of performing a Bell state measurement, for a complete characterization of the dynamics. This measurement apparatus is used in four ensemble measurements each corresponding to a different input state. Figure 1(c) and Table 1 summarize the preparations required for DCQD in the single qubit case. This table implies that the required resources in DCQD are as follows: (a) preparation of a maximally entangled state (for population characterization), (b) preparation of three other known (nonmaximally) entangled states (for coherence characterization), and (c) a fixed Bell-state analyzer.
Table 1. Input States and Measurements for Direct Characterization of Quantum Dynamics (DCQD) for a Single Qubita
 MeasurementOutput
Input stateStabilizerNormalizerBell state measurementχmn
+ZAZB, XAXBN/APΨ±, PΦ±χ00, χ11, χ22, χ33
α+ZAZBXAXBPΦ+ ± PΦ, PΨ+ ± PΨχ03, χ12
α+XXAXBZAZBPΦ+ ± PΨ+, PΦ ± PΨχ01, χ23
α+YYAYBZAZBPΦ+ ± PΨ, PΦ ± PΨ+χ02, χ13
a

Here |Φα+⟩ = α|00⟩ + β|11⟩ (), |Φα+X(Y) = α|++⟩X(Y) + β|−–⟩X(Y) ( and Im(αβ*) ≠ 0) and {|0⟩, |1⟩}, {|±⟩X}, {|±⟩Y} are eigenstates of the Pauli operators Z, X, and Y. The fourth column shows the Bell state measurement (BSM) equivalent of the stabilizer + normalizer measurements. The last column is the output: all elements of the process matrix χ.

We have assumed no SPAM errors. However, this assumption can be relaxed in certain situations, in particular when the imperfections are already known. (125,126)
The DCQD scheme has been generalized to higher-dimensional quantum systems, which includes the multiqubit case. (123) The primary advantage of DCQD over all separable QPT methods (in particular, SQPT and separable AAPT; recall that nonseparable AAPT requires many-body interactions, which are experimentally unrealizable) is that the minimal number of required experimental configurations is reduced quadratically from D4n to D2n, for n qudits of dimension D each. For example, the number of experimental configurations required for systems of three or four physical qubits is reduced from ∼ 5 × 103 and ∼ 6.5 × 104 in SQPT to 64 and 256 in DCQD, respectively. (26) Moreover, only a single ensemble measurement is required for a commuting set of observables, as was recently rediscovered in the context of entanglement-enhanced learning scheme of Pauli channels. (99)

3.3. Gate Set Tomography

The main disadvantage of QPT is its sensitivity to SPAM errors. Building on quantum state and process tomography, gate set tomography (GST) was introduced for estimation of gate performance in quantum processors while avoiding the impact of SPAM errors, specifically calibration of gate errors and measurements. (31−33) Since its introduction, there have been many experimental implementations that have assessed gate errors by utilizing GST. (128−137)
Unlike quantum process tomography, GST estimates the performance of a gate set without relying on prior calibration of a system, making GST a more generic method for gate performance characterization. Like QPT, GST differs from RB in that it provides a complete characterization of gates, whereas RB provides a single parameter. As an example, RB is insensitive to coherent errors, whereas GST attempts to amplify the effect of coherent gate errors such as rotation and phase errors. (138)
GST is used for estimating the performance of the reconstructed gates after applying the gates defined in a gate set, as we describe below. To start the derivation of the GST method, we need to consider the probability of measuring the system along a certain axis using a measurement operator Mi from a POVM (see Section 2.1) performed after an arbitrary single gate G is applied to a system in the state ρ, which can be written as
pi=Mi|G|ρ=Tr(MiGρ)
(34)
where |·⟩⟩ (⟨⟨·|) represents the superket (bra) operators in Hilbert-Schmidt space.
The superoperator notation is commonly used to simplify manipulations with quantum maps. (139) Using this formalism density matrices ρ acting on a Hilbert space of dimension D are “vectorized” and denoted as |ρ⟩⟩, a vector in a Hilbert-Schmidt space of dimension D2. The superket |ρ⟩⟩ is a vector created from the matrix representation of an operator by consecutively stacking its columns. Similarly, quantum operations are represented as matrices of dimension of D2 × D2 (in the same basis as |ρ⟩⟩) with the inner product defined as ⟨⟨A|B⟩⟩ = Tr(AB), where A and B can be density operators.
Assuming a processor can perform Nρ state preparation gates and NG distinct gates with NM measurement operators, where the i’th measurement can result in NM(i) distinct outcomes, we can write down the gate set of the system as
G={{|ρ(j)}j=1Nρ;{Gj}j=1NG;Mj(m)|m=1,j=1NM(j),NM}
(35)
This gate set consists of all possible operations that one can perform on the system, where all the gates are dependent on each other. eq 34 implies that the measurement probability for any combination of measurement, gate, and initial state is invariant under a similarity transformation S, i.e.,
Mj|Mj|S1|ρ(j)S|ρ(j)GSGS1
(36)
The outcome of a gate set G remains invariant up to an arbitrary invertible S, which suggests choosing a “gauge” S that results in the highest fidelity after reconstructing the gate set.
To make GST independent of the prior calibration of the system and make it robust against SPAM errors, we need to introduce fiducial gates and rewrite the state preparation and measurement gates in eq 35 as
|ρ(j)Fj|ρ(j)Mi|Mi|Fi
(37)
where Fi is the fiducial circuit constructed using the gates {Gj} in the gate set. Figure 2 shows the circuit required to perform GST on a system.

Figure 2

Figure 2. Gate set tomography circuit consisting of the fiducial gates Fi (Fj) as well as the measurement operators M and the native gate Gk performed on the initial state ρ. Figure adapted from ref (141). Copyright 2021 Quantum.

The most important caveat here is that one needs to make sure that these fiducial circuits form an informationally complete set of state preparation and measurement operators, i.e., measurements that allow for a complete reconstruction of the quantum state. (140)
As an example, we can form a gate set such that {G}={I,X,Y}={G0,G1,G2} with fiducials {F} = {I,X,Y,XX}, where X and Y represent a π/2 rotation about the Bloch sphere x and y axis, respectively. The choice of XX=X instead of X is made here to reduce the experimental cost of implementing the protocol. The key point in this example is that we can span the Bloch sphere (informational completeness) after applying the fiducial gates to the states and measurement operators.
In the final step of GST, we need to define and optimize a likelihood function such that it reproduces the desired gates based on the experimental data. To this end, we first define the outcome probability of applying the GST protocol on the system of interest as (97)
pijk(θ)=Mi(θ)|Fi(θ)Gk(θ)Fj(θ)|ρ(θ)
(38)
where we combined eqs 34 and 37. Here θ⃗ is a vector consisting of a set of parameters that need to be optimized using the maximum likelihood estimation (MLE) method, i.e., the elements of the T matrix from the Cholesky decomposition introduced below.
The likelihood function can be simplified and written in a normalized form as
L(G)=ijk(mijkpijk(θ))2/σijk2
(39)
with mijk representing the measured values in the experiment corresponding to the probabilities pijk, and σijk2 is the variance in the measurement outcome mijk. The remaining task is to minimize the likelihood function [eq 39] using the process matrix χ.
We can write eq 3 in terms of the process matrix χG for a gate superoperator G as
G(ρ)=i,j=0D21(χG)ijEiρEj
(40)
where {Ei} is a Hermitian operator basis for B(H). Since χG is positive semidefinite we can perform a Cholesky decomposition and rewrite it as χG = TT where T is a lower triangular matrix composed of complex elements on the off-diagonal and real elements on the diagonal. (142)
The trace-preservation property of the density matrix Tr(G(ρ))=Tr(ρ) leads to a completeness relation that can be used to reduce the number of free parameters in χG through D2 linear equalities:
i,j=0D21χijTr(EiEkEj)=δk0,k=0,...,D21
(41)
As mentioned above, θ⃗ is composed of the matrix elements of T, now seen to be (D4D2)/2.
Rewriting the likelihood function in terms of the process matrices using eq 40 we obtain
pijk(θ)=Mi(θ)|Fi(θ)Gk(θ)Fj(θ)|ρ(θ)=Tr[MiFi{Gk(Fjρ)}]=mnrstu(χFj)tu(χGk)rs(χFi)mn×Tr(MiEtErEmρEnEsEu)
(42)
By substituting this expression into eq 39, observing that since Mi is a POVM element we have IMi ≥ 0, along with the completeness relation [eq 41] and the unit-trace condition of the density matrix Tr(ρ) = 1, one can solve the MLE problem by varying the elements of the vector θ⃗ (elements of the T matrix). Note that as should be clear from the example we discussed above, the process matrices χG and χF are related and need to be defined in terms of each other. Moreover, the measurement Mi and state ρ also need to be parametrized similarly to χG in terms of their Cholesky decompositions. Optimization of the parameters of θ⃗ after minimizing the likelihood function [eq 39], allows us to reconstruct the entire gate set as well as the process matrices χG and χF using the Cholesky decomposition mentioned above.
In summary, the following steps are needed to implement the GST protocol:
1.

Initialize the qubit in a pure state: ρ = |0⟩⟨0|.

2.

For a particular choice of i, j, and k apply the gate sequence FiGkFj.

3.

Measure the POVM Mi. Typically in experiments, Mi = |0⟩⟨0|.

4.

Repeat steps 1–3 many times (typically ∼2000). Average and record the results to obtain mijk.

5.

Repeat steps 1–4 for all i, j, and k.

6.

Solve the MLE problem using the likelihood function defined in eq 39 to extract the optimized vector θ⃗ to reconstruct the entire gate set G, as well as the process matrices using the aforementioned Cholesky decomposition.

Another variant of this protocol, known as long-sequence GST, has been developed to amplify and highlight the source of errors. (141) The main difference in this method is that we apply the gates Gk multiple times to amplify the role of coherent errors. The sequence applied is Fi(Gk)lFj and l can be varied to estimate the error rates caused by the miscalibration in gates.
In summary, GST provides a comprehensive measure of gate performance in quantum processors. It gives all the details one needs to study and calibrate gates. However, despite recent efforts to simplify the implementation of GST in experiments, (143) GST has major drawbacks such as its inability to capture non-Markovian noise (144) and being extremely resource intensive. For example, to reconstruct the complete gate set for a single qubit one needs to conduct 80 distinct experiments with various gate sequences, with this number increasing to over 4,000 for 2-qubit gates. (97) Note that, as mentioned above, each experiment needs to be averaged over ∼ 2000 runs to minimize the sampling error, making the GST protocol even more resource-intensive. This can make GST impractical in actual experiments, a consequence of the overcompleteness of this method. Recent work has shown that the number of experiments in the two-qubit case can be reduced by a factor of ∼ 8 to elicit only the right amount of information required for gate calibration. (145)

3.4. Direct Fidelity Estimation

Direct fidelity estimation (DFE) (35,36) can be used for estimating how close an implementation of a unitary gate is to the ideal gate, via the average output fidelity metric for quantum maps. That is, DFE provides a single number between 0 and 1 to judge the given gate implementation, where 1 represents the ideal gate implementation. The advantages of this method are that it does not make any assumptions about the underlying noise model and is relatively inexpensive and easy to implement. However, one crucial drawback is that DFE assumes negligible SPAM errors.
DFE essentially provides a Monte Carlo technique to estimate the Uhlmann fidelity between a given and a target pure state. If the given state is ρB(H) and the target pure state is σB(H), then the Uhlmann fidelity [recall eq 6] reduces to
F(σ,ρ)=Tr(σρ)
(43)
Let {Pk} denote the set of Pauli group operators (including identity) in B(H), normalized so that Tr(PkPl) = kl, where D=dim(H). We can then expand
σ=1DkσkPk,σk=Tr(σPk)
(44)
and similarly for ρ. Next, we note that we can recast the fidelity as an expectation over an appropriately defined random variable X that can take values
xk=ρkσk
(45)
with probability
pk=P(X=xk)=σk2D
(46)
To see this, note that
Tr(σρ)=1D2k,lσkρlTr(PkPl)=1Dkσkρk=kpkxk=X
(47)
Note that eq 46 denotes a normalized probability distribution. The expectation ⟨X⟩ can then be approximated using Monte Carlo techniques, assuming that we are able to sample from the distribution P. The rigorous bound on the error of the Monte Carlo procedure is given by the following result (adapted from Theorem 1 of ref (36).):
Theorem 1 Given that
• {i1, ..., iN} are N indices sampled from the probability distribution P
each Tr(ρPik) is estimated using N2[k] samples
the fidelity estimate F̅ can be obtained within an error margin of the true fidelity F(σ, ρ) given by
P(|FF¯|<ϵ1+ϵ2)1Nϵ12+2exp[ϵ22N22(k=1N1σik2N2[k])1]
(48)
where σi = Tr(σPi). Here ϵ1 and ϵ2 are the errors associated with the Monte Carlo estimate and experimental estimate of Tr(ρPik), respectively.
As in Sections 2.1 and 3.2.2, we can invoke the Choi-Jamiołkowski isomorphism (67,68) between quantum processes and quantum states. We can then extend DFE to benchmarking gates. As in eq 4, define the Choi states ρU=(IU)(|ϕϕ|) and ρU~=(IU~)(|ϕϕ|), where U represents the ideal unitary gate operation and U~ represents the actual (noisy) experimental implementation. Then the benchmark metric that is calculable via DFE is the average output fidelity F¯(U,U~), defined as the average of the process fidelities [eq 6] of U and U~ over all possible pure states chosen uniformly at random. It can be shown that (146)
F¯(U,U~)=DFproc(ρU,ρU~)+1D+1
(49)
where the expression on the right is calculable using DFE. A naive implementation would require constructing the maximally entangled state |ϕ⟩. However, one can sidestep the issue of preparing |ϕ⟩ in experiments by simplifying the expression for Fproc(ρU,ρU~):
Fproc(ρU,ρU~)=1D2j,k(ρU)jk(ρU~)jk=1D4j,kTr(U(PjT)Pk)Tr(U~(PjT)Pk)=1D4j,kTr(U(Pj)Pk)Tr(U~(Pj)Pk)
(50)
where the first line follows from eq 47 with (ρE)jk=Tr(ρEPjPk), the second line due to the identity Tr[(OE)(|ϕϕ|)]=1DTr(E(OT))), and the third line from the property of Pauli matrices: PT = ±PT. The expression then no longer involves the entangled state |ϕ⟩. The corresponding experimental procedure is schematically illustrated in Figure 3 and becomes
1.

Determine the tuple of Pauli operators (Pj, Pk) for which Tr(U(Pj)Pk)0. Call this set A.

2.

Sample a tuple of Pauli operators, each of length log2(D), from the set A according to the probability distribution P(Pj,Pk)=Tr(U(Pj)Pk)2/D4. This is done via importance sampling. Note that here it is relatively easy to sample because it is possible to calculate the entire probability distribution.

3.

Prepare the D eigenstates for Pj, which require only local operations.

(a)

For each eigenstate, apply U~, and calculate the expectation of Pk (which also requires only local operations and measurements). Call the estimate for the expectation ãi, and the eigenvalue of the eigenstate λi.

(b)

The estimate for Tr(U~(Pj)Pk) is then ∑iλiãi. Store Tr(U~(Pj)Pk)Tr(U(Pj)Pk) as the estimate for the sample.

4.

Repeat the procedure N times. The average of the sample estimates gives F(ρU,ρU~).

Figure 3

Figure 3. Schematic of DFE: To estimate the state fidelity, expectations of random Pauli strings Pj are calculated for the input state ρ. The distribution of the random Pauli strings depends on the state with respect to which the fidelity is being calculated.

Alternatively, DFE can also be used for calculating the entanglement fidelity given by Fe=Tr(UU~)/D2 where U and U~ are both now treated as matrices acting via left multiplication to give the output state under the corresponding channel. (36)
Theoretically, the number of samples required for DFE scales exponentially with the number of qubits n in the worst case. However, for most states of interest, the number of samples required is only polynomial in n. In practice, DFE is efficient: a variant of it was demonstrated on 7 qubits, (147) while vanilla DFE has been used for a benchmarking implementation of a Toffoli gate using superconducting qubits. (148) However, DFE has been limited in its adaptation to benchmarking due to its susceptibility to SPAM errors; in this sense, alternatives such as RB (Section 3.1) that are robust to SPAM errors are preferred. (53,149,150)
DFE has been extended to simpler and more efficient measurement schemes for fidelity estimation, and tested on multiqubit entangled states on both trapped-ion (151−153) and superconducting quantum devices. (151) Other uses of DFE include tuning control for logic gates in error correcting codes, (154) and calculating minimum gate fidelities. (155)

3.5. Process Fidelity Estimation

An alternative and generalization of DFE is the Process Fidelity Estimation (PFE) method introduced in ref (34). This method is also concerned with estimating the process fidelity between an ideal and actual, noisy process, but generalizes DFE to also account for measurements. As such, it is applicable to entire quantum algorithms.
Consider a projective measurement in the σz-basis described by Π(ρ) = ∑kΠk(ρ) = ∑kπkρπk (where πk = |k⟩⟨k| and {πk} are a complete set of orthogonal projectors), an ideal unitary map U(ρ)=UρU, and its noisy implementation U~. Let the ideal map be Φ=ΠU and its noisy version be Φ~=ΠU~. We can always formally decompose the noisy unitary map as U~=ΛUU where ΛU=U~U is a U-dependent noise map that captures all the noise acting during the noisy implementation of the ideal unitary.
Here, the orthonormal basis states {|i⟩} in the maximally entangled state |ϕ=1Di=1D|i|i used to construct the Choi state [eq 4] can be chosen to be the eigenstates of the projectors {πk}, i.e., πk|i⟩ = δki|k⟩. Since ΦU=Π and Φ~U=ΠΛUΠ~, and the Uhlmann fidelity is unitarily invariant, we have
Fproc(ρΦ,ρΦ~)=Fproc(ρΠ,ρΠ~)
(51)
Note that
Π(A)=kπkAπk=kk|A|kπk
(52)
Writing eij ≡ |i⟩⟨j|, the Choi state of Π̃ is given by
ρΠ~(IΠ~)(|ϕϕ|)=1Di,j=1DeijΠ[ΛU(eij)]=1Dk=1Di,j=1Dk|ΛU(eij)]|keijπk=1Dkσk~πk
(53)
with
σk~ijk|ΛU(eij)|keij
(54)
Similarly, in the noiseless case, i.e., after dropping the tilde everywhere:
ρΠ(IΠ)(|ϕϕ|)=1Dkσkπk
(55)
where:
σki,jk|eij|keij=i,jδkiδjkeij=πk
(56)
so that
ρΠ=1DkPk,Pk=πkπk
(57)
and {Pk} is another set of orthogonal projectors: PkPl = δklPk. Therefore
ρΠ2=1D2k,l=1DPkPl=1DρΠ
(58)
so that DρΠ is the unique positive semidefinite square-root of ρΠ, i.e., ρ=DρΠ.
Thus, combining eqs 6 and 51
Fproc(Φ,Φ~)=Fproc(Π,Π~)=F(ρΠ,ρΠ~)=TrρΠρΠ~ρΠ=TrdρΠρΠ~ρΠ=1DTrkakPk
(59)
where ak = Tr(πkσ̃k). Noting that
(kakPk)2=klakalPkδkl=kakPk
(60)
we can write
Fproc(Φ,Φ~)=1DTrkakPk=1Dkak
(61)
Now, since
ak=k|σk~|k=i,jδkiδjkk|ΛU(eij)|k=k|ΛU(πk)|k
(62)
this finally yields
Fproc(Φ,Φ~)=1Dk=1Dk|ΛU(πk)|k=1Dk=1DPr(k|ΛU(πk))
(63)
where Pr(x|ρ) is the probability of measuring the output x given a state ρ. This result is analogous to the process fidelity expression associated with the DFE method, i.e., eq 50. However, the two methods clearly result in very different operational interpretations.
Since ΛU(πk)=U~[U(πk)], eq 63 has the following operational interpretation: it means that we can determine the fidelity of a noisy unitary that is followed by a measurement, by preparing the time-reversed initial states |ψk⟩ = U|k⟩, applying the noisy unitary U~ and combining the probabilities to measure the respective outputs k. However, this requires the ability to prepare the initial states |ψk⟩ with high fidelity, which is a potential drawback of the PFE method. In certain cases this only involves accurately implementable single qubit gates (34) (Hadamard and virtual phase gates).
As the number of different initial states |k⟩ increases exponentially, we can sample m different values {kl}l=1...m, where kl ∈{1, ..., 2n}, by drawing randomly from the uniform distribution over all 2n values. Note that using the square of the sample average (denoted by an overline) as an estimator for E[X]2 has an O(1/m) bias. This can be seen by noting that E[(X¯)2] = E[(1miXi)2] = E[1m2(ijXiXj+iXi] = m1mE[X]2 + 1mE[X]. However, we can construct an unbiased estimator of E[X]2 by using mm1(X¯)21m1X¯, and estimate the process fidelity via
[Fproc(Φ,Φ~)]2=[1Dk=12nPr(k|ΛU(πk))]2mm1[1ml=1mPr(kl|ΛU(πk))]21m(m1)l=1mPr(kl|ΛU(πk)))
(64)
where l runs over the samples.
This method was used recently in an experimental demonstration of the quantum Fourier transform using midcircuit measurements. (34) Given its novelty, it has not yet been implemented in other experiments.

3.6. Cross-Entropy Benchmarking

Cross-entropy benchmarking (XEB) is a scalable noise characterization and benchmarking scheme that estimates the fidelity F of a n-qubit circuit composed of single- and two-qubit gates. The procedure is based on sampling bitstrings from the output probability distribution of a quantum circuit of interest and was first proposed to characterize the fidelity of running random quantum circuits on a noisy quantum processor, a task known as random circuit sampling (RCS). (37) Later, it was used to produce indirect evidence of quantum supremacy. (39) Following these developments, a new gate characterization method called RCS benchmarking was built on top of XEB and RCS, where the overall noise strength of a circuit is estimated using XEB of random circuits. (156)
Consider the task of sampling bitstrings from the quantum state generated by a circuit. An ideal circuit CU, which implements a unitary U prepares an ideal pure state ρU = |ψU⟩⟨ψU|, which can be used for sampling. In reality, the bitstrings are drawn from a noisy state ρŨ prepared by the noisy circuit CŨ instead. We can model the effect of noise in terms of a simple CPTP map:
ρU~=FU|ψUψU|+(1FU)σU
(65)
where σU denotes an “error state” that captures the effect of the noise, and FU = ⟨ψUŨU⟩ is the fidelity we would like to estimate. Methods such as DFE (see Section 3.4) satisfy precisely this task but have poor worst-case scaling, requiring an exponential number of samples in the number of qubits n. Instead, to compute F, we first define the cross-entropy metric (157) between the ideal and noisy outcome probability distributions:
S(pU~,pU)=x{0,1}npU~(x)log[pU(x)]
(66)
where pU(x) = ⟨xU|x⟩ and pŨ(x) = ⟨xŨ|x⟩ denote the ideal and actual probability distribution of observing the bitstring x from the corresponding ideal and actual states ρU and ρŨ respectively. Note that pU(x) is the probability of bitstring x computed classically for the ideal quantum circuit U, while pŨ(x) is obtained from running the actual noisy quantum circuit Ũ using a quantum computer.
When the set of circuits C is random, the expectation of the cross entropy is (38)
EU[S(pU~,pU)]=EU[xpU~(x)log[pU(x)]]=EU[FUS(pU,pU)+(1FU)S(pσU,pU)]
(67)
where pσU = ⟨xU|x⟩ and we substituted eq 65 in the third line. Under the assumption that the fidelity depends only on the specific gates applied, and not their order, i.e. FU = F, solving for F yields
F=EU[S(pσU,pU)]EU[S(pU~,pU)]EU[S(pσU,pU)]EU[S(pU,pU)]
(68)
In a loose sense, this reflects an uncorrelated noise model. We further assume that for sufficiently deep circuits, σU results in a probability distribution pσU which under expectation behaves like a uniform distribution, i.e., EU[pσU] is a uniform distribution over bitstrings. This is formalized as follows:
xx|σU|xlog[pU(x)]=1Dxlog[pU(x)]+ϵ
(69)
where D = 2n and ϵ decreases sufficiently fast with the number of qubits n. This forms the central assumption in XEB methods. For a depolarizing map, the error state is σU = I/D, and eq 69 is exact. More generally, the error ϵ in eq 69 decreases exponentially with n. This can be understood as the “concentration of measure phenomenon” in high-dimensional spaces in the limit of a large number of qubits n. (158)
Eq 69 makes the expression EU[S(pσU,pU)] classically calculable. Therefore, in eq 68, only the term EU[S(pU~,pU)] needs to be calculated experimentally; the rest are calculated classically.
In practice, instead of using the logarithm to calculate the standard XEB, a linearized version of XEB is considered. The linear XEB is defined as (39)
Slinear(pU,pU~)=Dx{0,1}npU~(x)pU(x)1
(70)
Intuitively, both the standard and linear XEBs are expectations of functions (EXpU~[logpU(X)] and EXpU~[pU(X)D1], respectively) which map observed bitstrings with higher ideal probabilities to larger values. When the circuits are noiseless, in expectation, the bitstring samples are drawn from the ideal probability distribution of random circuits, which is called the Porter-Thomas distribution, (159) an exponential distribution over a random permutation of bitstrings. This is the limit where F = 1. In the opposite limit where the bitstring samples are drawn from the uniform distribution, we have pU~(x)=1D and F = 0. The same analysis as eq 68 follows through with the modified, linear XEB definition.
The above exposition describes the theory of using the XEB to calculate the fidelity F. Through heuristic arguments, (38,39,160) the unbiased linear cross entropy is considered an estimator of the fidelity itself.
As in the case of DFE (Section 3.4) and PFE [eq 64], directly computing the exponentially large sum involved in eq 70 is impractical, and instead we resort to a sampling method. To construct an unbiased estimator for the linear cross entropy, consider M bitstring samples S = {xi}i=1M drawn from the probabilities of the noisy output state of a random circuit, the linear cross-entropy fidelity estimator is given by
F^XEB=DMi=1MpU(xi)1
(71)
Reference (161) proposed the following form for an unbiased linear XEB fidelity estimator:
F^uXEB=DMi=1MpU(xi)1Dx{0,1}npU(x)21
(72)
Note that one will need to compute all 2n output probabilities in the denominator of eq 72 to compute the unbiased linear XEB fidelity estimator. For random circuits, the denominator in eq 72 approaches 1 in log depth, (162) hence the two XEB fidelity estimators, eq 71 and eq 72 give the same result as the depth increases. However, the unbiased fidelity estimator uXEB is more accurate for circuits with short depth.
In both estimators for XEB, {xi} is sampled from the noisy circuit, while the ideal circuit is classically simulated to provide the values pU(xi). A variety of different classical algorithms have been devised for this task, among which tensor networks are the most efficient. (163−165)
Thus, computing FXEB becomes infeasible for large circuit sizes that do not have tractable classical simulations. This computational hurdle motivates using the XEB measure as a way to certify a quantum computing advantage. More precisely, one can demonstrate a quantum advantage over classical algorithms, for the specific task of estimating the XEB values in the context of the computational task of sampling bitstrings from the probability distributions generated by random circuits. This idea was the basis for the first quantum supremacy experiment. (166)
Among other things, the experimental linear XEB was shown to be consistent with a simple uncorrelated noise model. (166) This fact underlines that sampling from random circuits can be used as a way to benchmark noise in quantum circuits. This is known as RCS benchmarking. (156)
In an RCS benchmarking circuit, similar to the circuit in the IRB protocol (Section 3.1.3), the gate to be characterized is applied in an alternating fashion and interleaved with Haar random single-qubit gates (approximated using unitary t-designs (167,168)), followed by measurements in the computational basis; see Figure 4. The noise introduced by the imperfect gates is assumed to be described by a Pauli noise model of the form N(ρ)=pασαρσα, where σα is an n-qubit Pauli operator and pα is the associated error probability. The RCS benchmark procedure is the following:
1.

Sample a random circuit Ci from an ensemble of random circuits RCS(n, P) with n qubits and depth P, where P denotes the layers of two-qubit gates independently drawn from the Haar measure on U(4).

2.

Estimate the fidelity of Ci (denoted P,i).

3.

Repeat step 1–2 for L instances of random circuits and obtain the average fidelity F^P=1LiLF^P,i.

4.

Repeat step 1–3 for varying depths P{1,...,P} to obtain a sequence of average fidelities {F^P}1P.

5.

Fit the average fidelities {F^P}1P to an exponentially decaying function F = AeλP, where the fitted parameter λ characterizes the overall noise strength.

Figure 4

Figure 4. RCS benchmarking circuit: layers of 2-qubit gates (white boxes) are interleaved with Haar random 1-qubit gates (blue boxes), followed by measurement in the computational basis at the end. Adapted from Figure 1 of ref (156). Copyright 2021 arXiv.

The key step in the procedure above is estimating the circuit fidelity F, which is defined as the overlap between the output state |ψU⟩ of the ideal circuit and the mixed state ρŨ generated by an experiment, i.e., FU = ⟨ψUŨU⟩. This type of fidelity estimate, in general, requires exponentially many measurements to determine the state ρŨ. To circumvent this exponential cost, the cross-entropy fidelity estimator in eq 71 is considered instead.
Under the assumption of a Pauli noise map, it has been shown, both analytically and numerically, (156) that the average fidelities of the random circuits decay exponentially as a function of circuit depth P, i.e., E[F]=eλP where λ = ∑αpα, which allows one to extract the effective noise rate of the quantum hardware λ by estimating E[F]. Similar to the RB protocol, when fitting the experimentally estimated XEB fidelities {F^P}P=1P to an exponentially decaying function, the depth-independent fitting parameters correspond to SPAM errors, which can be estimated independently. For this reason, the XEB benchmark protocol is considered robust to SPAM errors. The exponential decay characteristic and the related extraction of noise parameters can be generalized beyond Pauli noise maps. (164)
Another desirable feature of the XEB benchmark is that the fidelity estimation procedure is sample-efficient. To see this, we consider the variance of the unbiased XEB fidelity estimator, which has been shown to be given by (161)
VarC,S(F^uXEB)=O(1M+λ2(E[F])2)
(73)
This suggests that the fidelity estimator is sample-efficient (compared to, e.g., DFE) and requires M = O(1/ϵ2) measurement samples to achieve ϵ additive accuracy for each instance of random circuits. The first term in the variance expression is the leading term, especially in large-scale experiments, due to the exponential decay of E[F]. In practice, it suffices to collect a large number of samples for a few circuits to estimate E[F] with a small additive error.
Unlike the standard RB protocol described in Section 3.1, which involves Clifford gate sequences, the XEB protocol can also be applied to circuits with arbitrary non-Clifford gates. Moreover, RB does not scale easily beyond two or three qubits due to the large circuit depth it requires; XEB, in contrast, is limited only by the classical simulability of the underlying circuit and can extend to up to ∼50 qubits. (156) Hence, it can be used to benchmark broader families of quantum gates and quantum circuits. However, the XEB fidelity estimator can be insensitive to certain types of noise. As an example, under a Pauli noise model, local Pauli errors can add up or cancel as they propagate forward in time through random circuits. When the propagated error operators before the measurement are of type I or Z, they do not contribute to the XEB fidelity as they do not affect measurements in the computational basis. It has also been shown that the XEB is not a good proxy for the fidelity in the presence of time-correlated errors. (169) Overall, the XEB protocol is useful in estimating the total amount of error in a sample-efficient way under certain strong assumptions about the noise. To extract more information about the noise map, alternative noise characterization techniques are needed.

4. Deterministic Benchmarking

Click to copy section linkSection link copied!

In this section, we introduce a method we call deterministic benchmarking (DB) (170) and validate its utility through experiments conducted on a superconducting transmon qubit. (73) Unlike RB, which averages over many and lengthy random Clifford gate sequences to approximate all errors as a depolarizing map and lacks the sensitivity needed to detect small changes in decoherence times and qubit frequencies, (144,171) DB is highly sensitive to such changes. As illustrated in Figure 5(a,b), DB utilizes a carefully chosen small set of gate-pair sequences that specifically target the real noise affecting single-qubit gates. The DB protocol is significantly more efficient than the other methods we have reviewed in earlier sections but is currently limited to single-qubit gates. Indeed, we show below that DB requires only four sets of state-fidelity measurements to accurately estimate all incoherent and coherent errors relevant to single-qubit gate operations. Additionally, DB can also be used to characterize the bath temperature. Moreover, we show that DB also enables exploration of how coherent and incoherent errors interact to cause biased noise, whose impact on fault tolerance has been thoroughly studied. (83,86)

Figure 5

Figure 5. Schematic representations of the gate sequences for (a) randomized benchmarking (RB) and (b) deterministic benchmarking (DB), respectively. RB uses N random Clifford gates Ci, followed by a single recovery gate Cr = CN···C1. DB prepares an initial state via U|0⟩, applies n repetitions of a fixed, deterministic two-pulse sequence P1P2, and unprepares the initial state via U.

The gate sequences used in DB resemble the simplest dynamical decoupling (DD) pulse sequences introduced decades ago in nuclear magnetic resonance. (172−174) But, while DD is nowadays typically used to suppress incoherent and coherent noise during idling─whether between gates (34,175−179) or as a result of crosstalk, (180−186) the goal of DB is instead to characterize the noise arising during gate operation.

4.1. Qubit Model

We base our presentation on transmon qubits, (73) which we model as driven two-level systems and consider in the drive frame under the rotating wave approximation. (71) However, the DB methodology is general and applies to any qubit modality.
Let Rα(θ) ≡ exp[−i(θ/2)σα], where {σα} is the set of Pauli matrices. The Hamiltonian that generates the single-qubit gates Rx(θ) and Ry(θ), including both phase and rotation errors, is given by
Hα=(ε+εerr)σα2+Δerrσz2,α{x,y}
(74)
Ideally, both the amplitude error εerr and the detuning error Δerr vanish. In reality, both are present and give rise to rotation and phase errors
δθεerrtg,δϕΔerrε¯=1θΔerrtg
(75)
respectively, with tg denoting the gate duration, θ ≡∫0tg ε(t)dt, and ε̅ = θ/tg the average pulse amplitude. We expect to see a pulse-shape dependence in the phase error δϕ, as the effective detuning error Δerr may depend on instantaneous drive amplitude ε(t). We assume square-shaped pulses for our analytical theory and cosine-shaped pulses for our numerical calculations using the Lindblad master equation (cosine pulses are used in experimental measurements).
Eq 74 is an effective model, where Δerr is an error resulting from both detuning─a discrepancy between the drive and qubit frequencies─and the ac-Stark shift induced by driving other transitions (such as the higher levels of the transmon) off-resonantly during the gate. (187) The two-level model is justifiable, given that leakage to higher levels of the transmon during our gates is very small (see Section 4.11).

4.2. Four Key Parameters

Recall from Section 2.1 that any completely positive trace-preserving (CPTP) map acting on a D-dimensional Hilbert space can be characterized using no more than D4D2 real parameters, i.e., 12 for a single qubit. While RB reduces this to a single effective parameter, DB makes a less drastic reduction, to four key parameters: the relaxation time T1, the coherence time T2 (which contains information on pure dephasing Tϕ), the rotation error δθ, and the phase error δϕ. Assuming a Markovian noise model, we show below that these parameters can be straightforwardly extracted and, at the same time, provide just the right amount of information about the errors affecting single-qubit gates to enable such gates to be accurately calibrated.
We denote XRx(π), XRx(−π), and similarly for Y and Y. To define DB, we need gate-pair sequences of the form P1P2. In particular, we consider P1P2 ∈{XX, XX, YY, YY, YY}. Note that in the absence of any errors, i.e., when εerr = Δerr = 0 and without coupling to a bath, P1P2 = ±I for this set (the global phase will not matter). In reality, P1P2 ∝̷ I and we consider the corresponding fidelity (n) = |⟨ψ|(P1P2)n|ψ⟩|2 where |ψ⟩ is the initial state and nZ+ is the number of repetitions of the P1P2 sequence; hence the fidelity is defined stroboscopically. We use a tilde to denote the experimentally measured fidelity.
The core DB protocol consists of a set of four “learning” experiments designed to extract {T1, T2, δθ, δϕ}, followed by tests involving new experimental data. In each case, the notation {P1P2;|ψ⟩} means a circuit comprising preparation of the state |ψ⟩ from the |0⟩ state, a varying number of repetitions of the P1P2 sequence, unpreparation of |ψ⟩, and finally a measurement in the σz eigenbasis. Finally, the empirical fidelity is found for each experiment as the ratio of |0⟩ outcomes to the total number of experimental shots (800 in our case). We find that in all these experiments─learning as well as testing─the fidelity can be accurately described using a simple analytical expression:
F(tn)=12(1+a)+12(1a)etn/TDcos(2ωtn)
(76)
where tn = 2ntg represents the total evolution time. The parameter TD denotes a characteristic timescale associated with incoherent noise processes, while ω is a frequency associated with coherent errors. The parameter a quantifies the deviation from the infinite temperature limit, where F()=12. Eq 76 effectively captures the interplay between coherent and incoherent noise sources. While it is phenomenological, it is also derivable from a Markovian master equation model subject to certain approximations, as shown in Section 4.9.
Finally, we note that eq 76 can be further generalized if we replace and , which adds an extra parameter. If we set ω = 0 (no coherent errors), this reduces to an RB fidelity-like expression Apn + B, where p = e–2tg/TD [recall eq 14]. RB’s ability to quantify state preparation and measurement (SPAM) errors as 1 – (A + B) is thus captured by this generalization of DB; however, we do not consider this here, as our focus is the determination of {T1, T2, δθ, δϕ}. Finally, note that coherent errors introduced by leakage do not fit the form of eq 76; we discuss this further in Section 4.11.

4.3. Protocol

We now detail the DB protocol, which comprises the following experiments or steps:
1.

{free;|1⟩}: This is a standard T1 measurement, where we do not apply any pulses (free evolution). eq 76 reduces to 12(1+a)+12(1a)et/T1 in this case which we fit to determine T1. This yields the relaxation component of the Lindbladian.

2.

{XX;|+⟩}: Fitting this experiment with eq 76 yields TD = T2, the coherence time from which we deduce the pure dephasing time Tϕ = 2T1T2/(2T1T2). This is analogous to an echo-Ramsey measurement (188) and yields the dephasing component of the Lindbladian. This experiment is also sensitive to leakage (discussed below); here we assume leakage is small.

3.

{YY;|+⟩}: This yields fidelity oscillations which we fit to find the rotation error δθ where ω ≈ δθ/(2tg) in eq 76.

4.

{XX;|+⟩}: Similarly, this yields fidelity oscillations which we fit to find the phase error δϕ where ω = δϕ/tg in eq 76. The decay times TD obtained here and in step 3 are a combination of the incoherent and coherent noise and hence different from T2 obtained in step 2. We now have all four parameters, which we use to numerically calculate predictions for the test experiments.

5.

Test 1: {YY;|+⟩} and {YY;|+⟩}: This highlights the interplay of asymmetric relaxation noise and gate operations and can also be used to characterize the qubit temperature (see Section 4.10).

6.

Test 2: Various other experiments. Here, we report tests using the UR6 dynamical decoupling sequence. (189)

Next, we demonstrate that this protocol enables a detailed assessment of both coherent and incoherent errors affecting single-qubit gates.

4.4. Experimental Results

We conduct all our experiments using a superconducting transmon qubit dispersively coupled to its readout resonator (see Appendix A for more details). We implement Ry(θ) gates by adjusting the phase of the microwave pulses used for Rx(θ) gates, thus eliminating the need for additional calibrations. All single-qubit gates are calibrated using the Derivative Removal of Adiabatic Gates (DRAG) technique (190) to minimize leakage and phase errors induced by higher levels of the transmon. To emulate drifts in experimental parameters, we first perform measurements on an outdated set of gate calibration parameters. The results are shown in Figure 6.

Figure 6

Figure 6. Empirical (, symbols), analytical fit (F, solid curves), and Lindblad-based (, dashed curves) fidelities for the initial states |1⟩ under free evolution and |+⟩ under various gate sequences P1P2. The analytical fit yields T1 = 23.36 ± 0.40 μs, T2 = 44.13 ± 2.49 μs, δθ = 0.398 ± 0.004°, and δϕ = 0.426 ± 0.004°. With the exception of the XX (orange) and YY (green) cases, the Lindblad-based results are indistinguishable from the analytical fits.

We observe that YY oscillates, indicating rotation errors δθ, as explained below. Similarly, XX oscillates due to phase errors δϕ. The XX sequence on |+⟩ is insensitive to coherent errors and, instead, XX exhibits an exponential decay characterizing the T2 time during time-dependent gate operations. The TD parameter from the XX sequence is chosen as the effective T2 since, subject to continuous time evolution under the noiseless version of this sequence, starting from |+⟩ the state always remains |+⟩. Also shown in Figure 6 are results from the YY and YY sequences. We discuss the reason for the pronounced difference they exhibit below.
We use eq 76 to fit (solid curves) each of the four experimental fidelities (symbols) in Figure 6 prescribed by steps 1–4 of the DB protocol and find close agreement.
We contrast the resulting {T1, T2, δθ, δϕ} values with the single metric rC derived from the RB protocol [eq 17]. Figure 7 shows the average RB result for 30 distinct experiments with a total length of 700 Clifford gates for the same gate parameters shown in Figure 6. On average, our RB Clifford gate set consists of approximately 80% π/2 and 20% π gates, with π/2 pulses expected to exhibit half the rotation errors and one-fourth of the phase errors. The RB curve can be fitted to a function of the form Apm + B (Figure 7). This process yields an average gate infidelity of rC = (1 – p)/2 = 0.262 ± 0.004% for the single-qubit Clifford gate set, a far less detailed characterization that misses the coherent oscillations and the distinction between T1 and T2. This also shows that RB alone is not sufficient to achieve accurate gate calibration. Note that this method is limited by the relaxation time of the qubit (T1) as well as the gate duration (tg) with the average single-qubit gate fidelity scaling as ∼ tg/T1 (191) and in our case, the gate fidelity is significantly affected by the relatively short T1 ≈ 20 μs of the device.

Figure 7

Figure 7. Randomized benchmarking measurement for the same gate calibration parameters used in Figure 6. The error bars represent the deviation of the averaged benchmarking results across 30 distinct random sequences with a maximum number of Cliffords (NCliffords) of 700, which are smaller than the marker sizes.

As we show in the next several sections, all observed features can be understood using either a simple, analytically solvable closed-system model or a numerical model based on a Lindblad master equation with time-dependent gates. We first provide the analytical fidelities in the closed system case, which exhibit oscillations for certain sequences.

4.5. Fidelity with Coherent Errors

Assuming the system is closed, eq 74 can be used to analytically compute the fidelities of the gate sequences comprising our DB protocol. The closed system fidelity FP1P2clsd(n) then reflects only the coherent errors contained within P1 and P2. For YY and XX and the initial state |+⟩, we find
FYYclsd(n)=cos2(nθerr)
(77a)
FXXclsd(n)=1(πδϕsin(nθerr)θerr)2
(77b)
where θerr(εtg+δθ)2+(πδϕ)2.
Expanding to the leading order in δϕ and δθ, with εtg = π, FYYclsd(n) simplifies to cos2(nδθ). Thus, the fidelity under the YY sequence is dominated by rotation errors δθ.
Regarding FXXclsd(n), in the small error limit, the second term in eq 77b is suppressed, and FXXclsd(n) ≈ 1. Consequently, there is no dependence on either the rotation or the phase error. Finally:
FXX¯clsd(n)=cos2(nφerr)φerrtan1(2πδϕθerrsin(θerr2)1Λ/21Λ)Λ(πδϕθerr)2(1cosθerr)
(78)
Expanding to leading order again, with εtg = π, we have FXX¯clsd(n)cos2[2nδϕ(11πδθ)]. Thus, the fidelity under the XX̅ sequence is dominated by phase errors δϕ.
These results justify our assignment of the oscillations observed in the YY and XX experiments shown in Figure 6 to rotation and phase errors, respectively.

4.6. Open System Model

The cosine dependence in eq 76 is well motivated by the closed system (coherent error) fidelities given above, corresponding to TD = and a = 0. The choice of the two incoherent parameters in the DB protocol, specifically T1 (associated with free evolution) and T2 (associated with the XX sequence), warrants further investigation through numerical simulation to better understand their contributions to the parameters TD and a in eq 76. We use a simple Lindblad master equation tρ=L(ρ) to capture the effects of nonunitary evolution, with a Lindbladian of the form
L=i[H,·]+α{1,ϕ}γα(Lα·Lα12{LαLα,·})
(79)
The Lindblad operators are L1 = σ = |0⟩⟨1| and Lϕ=σz/2 with rates γ1 = 1/T1 (relaxation) and γϕ = 1/Tϕ = (2T1T2)/(2T1T2) (dephasing).
Using the fitted parameters {T1, T2, δθ, δϕ} from Steps 1–4 of the DB protocol, we numerically solve the Lindblad equation and find good agreement with the experimental results, as shown by the dashed lines in Figure 6. Moreover, knowledge of these parameters allows us to proceed to step 5 of the protocol and predict, as a test, the outcomes of the YY and YY experiments. As shown in Figure 6, there is again good agreement between the simulations and the experimental results. This validates the choice of {T1, T2, δθ, δϕ} as an appropriate set of parameters. In contrast, rC is clearly insufficient for predicting the outcomes of the YY and YY experiments.

4.7. Sensitivity to Coherent Errors

Having confirmed that DB effectively identifies various errors, we next assess the sensitivity of both DB and RB to coherent errors. Initially, we calibrate the gates to minimize both δϕ and δθ (see Appendix B), and immediately follow this calibration by a series of DD experiments, as depicted in Figure 8(a). These experiments focus on the XX and YY sequences, which are sensitive to phase and rotation errors. As a second test (Step 6 in the DB protocol), we include a Universally Robust (URn) DD sequence, specifically designed to suppress coherent errors. (189) We then deliberately introduce coherent errors with (δθ, δϕ) values given in Figure 8. The phase error δϕ, is introduced by mistuning the DRAG weighting parameter α, which effectively varies the detuning error term Δerr in our model. The slight increase in rotation error δθ from Figure 8(c) (in the absence of δϕ) to (d) (in the presence of δϕ) can be attributed to the second order dependence of θerr on δϕ, given by 12(πδϕ)2/(π+δθ).

Figure 8

Figure 8. Experimental fidelities |+⟩ showing the very different sensitivity of DB and RB to coherent errors. (a) δθ = 0°, δϕ = 0°, (b) δθ = 0°, δϕ = 0.893 ± 0.002°, (c) δθ = 0.932 ± 0.007°, δϕ = 0°, (d) δθ = 0.995 ± 0.009°, δϕ = 0.90 ± 0.002°. These angles are obtained from fits using eq 76. Dashed curves are the numerically computed fidelities F|+⟩ from the open system Lindblad model using the DB parameters above along with the corresponding T1 and T2. The UR6 sequence suppresses even large coherent errors.

In each of cases (a)-(d) we also conduct RB experiments with the same values of Δerr and δθ. More precisely, when we add coherent errors to the RB gates, they are added to both R(π/2) and R(π) in such a way that for R(π/2) the rotation errors are 1/2 that of R(π) and the phase errors for R(π/2) are 1/4 of that of R(π).
As anticipated, immediately following gate calibration using the parameters extracted from the DB protocol, neither XX nor YY causes oscillations, as shown in Figure 8(a), and the coherent error is negligible. However, after deliberately introducing phase errors, significant oscillations appear in XX̅, as depicted in Figure 8(b), while YY simply decays exponentially without exhibiting any oscillations. In contrast, introducing a rotation error in Figure 8(c) reveals clear oscillations in YY but none in XX. When both errors are introduced, Figure 8(d) displays clear oscillations for both sequences.
The UR6 sequence never exhibits oscillations, confirming that coherent errors are their source. The Lindblad model with parameters obtained via the DB protocol (dashed curves) is again in good agreement with the experimental fidelities, further confirming the utility of DB. In contrast, the error rate derived from RB, rC , is unchanged when a phase error (δϕ ≈ 0.90°) is introduced and increases only slightly (≈ 0.02%) with a rotation error (δθ ≈ 0.93°). We see that small coherent errors can cause large, rapid oscillations in sequence fidelity even when the RB fidelity is almost unchanged. This illustrates RB’s limitations in detecting specific types of coherent errors.
Finally, note that in the presence of only detuning error Δerr in eq 74, the Rabi rate increases from ε to ε2+Δerr2. Thus, for a gate time tg, there would be an additional induced rotation error δθind=(ε2+Δerr2ε)tgΔerr2tg/(2ε) (assuming a square pulse), even without a drive amplitude error. Using Δerr = ε̅δϕ and ε̅ = θ/tg, this gives
δθind12ϕδϕ2
(80)
From the data in Figure 8, we have at most δϕ = 0.90° with a nominal rotation angle of θ = π, giving δθind ≈ 0.02° which is negligible. We therefore ignore the induced rotation error in our analysis.

4.8. Interplay of T1 Asymmetry and Gates

An intriguing observation arising from Figure 6 is the strong asymmetry in the decay of the YY and YY fidelity curves. Transmon qubits operate at frequencies in the 3 – 6 GHz range and temperatures around 10 mK (∼ 0.2 GHz). In this regime, the thermal energy is negligible, and the ground state |0⟩ is approximately the thermal state of the transmon qubit, i.e., excitation from |0⟩ to |1⟩ is negligible and relaxation dominates. Tracing the Bloch sphere trajectories corresponding to different sequences reveals that YY confines the state to the southern (excited state) hemisphere, while YY limits it to the northern (ground state) hemisphere. Therefore, the former is more susceptible to relaxation (T1) errors, which explains the faster decay and lower saturation fidelity of YY compared to YY. This asymmetry is also well captured by the time-dependent Lindblad model, as shown by the dashed curves in Figure 6.
To demonstrate this effect, we performed additional experiments in the same outdated gate calibration cycle used in Figure 6. We applied the XX sequence to |±i⟩ (eigenstates of σy), as well as to |ψπ/4,3π/4 and |ψ3π/4,3π/4, where |ψθ,ϕ⟩ = cos(θ/2)|0⟩ + e sin(θ/2)|1⟩. We observe a clear asymmetry between these pairs of orthogonal states under XX, as shown in Figure 9(a). We further confirm this asymmetry using YY applied to |±⟩ (eigenstates of σx) and {|ψπ/4,3π/4⟩, |ψ3π/4,–3π/4⟩, finding the same asymmetric decay patterns both experimentally and numerically in Figure 9(b). These results illustrate the dependence of relaxation during time-dependent gates on the initial state and its trajectory under different pulse sequences. Numerical modeling of this asymmetry enables qubit temperature calibration. (see Section 4.10).

Figure 9

Figure 9. Asymmetry in the decay pattern of DB sequences due to the interplay of gates and T1. Fidelity decay for (a) XX and (b) YY applied to two pairs of orthogonal initial states. Dashed curves represent the Lindblad master equation simulation results with parameters obtained by fitting eq 76.

4.9. Derivations: Fidelities of the YY and XX Sequences

In this section, we present a detailed analytical derivation of the fidelity subject to Markovian decoherence and coherent errors. We aim to derive our phenomenological fidelity expression, eq 76, as well the specific sequence dependence of the decay constant TD, the coherent oscillation frequency ω, and the SPAM parameters A and B.
We assume a Markovian environment which we model via the Lindblad master equation with time-dependent driving:
ρ˙=i[H(t),ρ]+LD(ρ)LD(ρ)=αγα(LαρLα12{LαLα,ρ})
(81)
We recognize that this is a phenomenological model and that rigorously derived master equations for time-dependent driving have a different structure, where the Lindblad operators Lα become time-dependent. (192−198) However, this simplified model is analytically solvable and suffices for our purposes.

4.9.1. Background

For analysis purposes, we work in the Bloch vector picture, which substitutes ρ = 12(I + vxσx + vyσy + vzσz), H = 12(hxσx + hyσy + hzσz), where v = (vx, vy, vz) ∈ R3, h = (hx, hy, hz) ∈ R3. The dissipator LD is replaced by (R, c) with Rij12Tr(σiLD(σj)) and ci12Tr(σiLD(I)). The Lindblad master equation then is = h × v + Rv + c (see, e.g., ref (199) for derivations of all these results). The corresponding fidelity with the |+⟩ state is given by
F(t)=12(1+vx(t))
(82)
We model our system as a qubit with an energy gap ω01 coupled to a thermal environment at inverse temperature β along with dephasing noise. The combined noise model translates to R = diag(−γ2, −γ2, −γ1) and c = (0, 0, η γ1) where
γ2=12γ1+γϕ
(83)
and
η=1exp(βω01)1+exp(βω01)[0,1]
(84)

4.9.2. YY Sequence

For a square YY pulse, h(t) = (0, εtot, Δerr) where εtot = ε + εerr. Then the Bloch vector equation becomes = Gv + c, where
G=(γ2ΔerrεtotΔerrγ20εtot0γ1)
(85)
We first solve the case where Δerr = 0 (δϕ = 0). Define
δG=(0Δerr0Δerr00000)
(86)
Exact Solution for Δerr = 0
The Bloch vector equation for Δerr = 0 becomes
v˙=(GδG)v+c
(87)
In this equation, vy is decoupled from the other two components and to solve for vx(t) and vz(t), requires exponentiation of a 2 × 2 matrix. Given that v(0) = (1, 0, 0), the relevant solutions for the Bloch vector components take the following form:
where
v=ηεtotγ1εtot2+γ1γ2γ*=12(γ1+γ2)ω*=εtot214(γ1γ2)2
(88)
Along with eq 82, eq 88 is close to, but not quite eq 76.
Perturbation Theory for Δerr = 0
In the regime where εεerr, γ1, γ2, the quantities in eq 88 simplify to
vηγ1ε,ω*εtot=ε+εerr
(89)
to first order. Moreover, since we are interested in the fidelities after each DD sequence, i.e., at t = 2ntg (nN), sin(2ωntg) ≈ sin(2nδθ) and cos(2ωntg) ≈ cos(2nδθ). The term corresponding to sin(2ωntg) vanishes to first order initially because it is multiplied by another first-order term and later because it is exponentially suppressed. This results in a simplified expression for vx(2ntg):
vx(2ntg)=v+(1v)e2ntgγ*cos(2nδθ)
(90)
The fidelity of the |+⟩ state as a function of YY repetitions n, Fδϕ=0(n), is therefore given by
Fδϕ=0(tn)=12(1+v)+12(1v)eγ*tncos(2nδθ)
(91)
Eq 91 is identical in form to eq 76, with a = v ≈ γ1/ε, TD = 1/γ* and 2ωtgδθ. Note that the oscillation period of 2nδθ matches that of the fidelity in the closed system case as discussed below eq 77b.
Perturbation Theory for Δerr ≠ 0
If the solution of the Bloch vector equation given in eq 87 is denoted by v0(t), then in the case where Δerr ≠ 0, let the deviation of the actual Bloch vector trajectory v(t) from v0(t) be given by ξ(t)R3, i.e., v(t) = v0(t) + ξ(t). Substituting this form into the Bloch vector equation = Gv + c gives
ξ˙=Gξ+δGv0
(92)
The solution for ξ(t) is then given by
ξ(t)=0teG(ts)δGv0(s)ds
(93)
Taking the norm of both sides and moving the norm inside the integral, we obtain
ξ(t)0teG(ts)eGδGv0(s)ds
(94)
Using eq 86, ∥δGv0(s)∥ simplifies to |Δerrv0,x(s)| since v0,y(s) = 0. From eq 88, ∥v0,x(s)∥ in turn can be strictly upper bounded by |v| + ke–γs for some constant k. A smaller value of k will result in a tighter bound. In the approximation regime discussed above, k can conveniently be chosen as 2.
If G = SDS–1 where D is the diagonal matrix of G’s eigenvalues λi and S is the similarity matrix with the corresponding eigenvectors of G as its columns, then
eG(ts)SeD(ts)S1κeλ(ts)
(95)
where κ=SS1 and λ=maxiR(λi). (200) This gives the inequality
ξ(t)<{κ|Δerr|(1eλtλ|v|+keγ*teλtλγ*)ifλγ*κ|Δerr|(1eλtλ|v|+kteλt)otherwise
(96)
Note that the first term is Oerrv/λ).
From first-order eigenvalue perturbation theory, we have
δλi=wi·δGviwi·vi
(97)
where wi and vi are the left and right eigenvectors of G in eq 85 with eigenvalues {λi} = {γ ± , −γ2}. Therefore, using eq 86:
λ1,2=(γ*±iω*)+O(Δerr)λ3=γ2+O(Δerr)
(98)
Under the additional assumption Δerr1,2 = O(1), the first summand in eq 96 simplifies to O(v) = O1/ε). Combining this with the fact that the second summand decays exponentially with time, ξ(t) is a small enough quantity such that v(t) can be well approximated by v0(t). Therefore, the resulting fidelity expression FYY(n) in a Δerr ≠ 0 system is still well captured by eq 91. It matches eq 76 with an error |FYY(n) – Fδϕ=0(n)| < 12ξ(2ntg), with the factor 1/2 due to eq 82. In big-O notation,
|FYY(n)Fδϕ=0(n)|=O(γ1/ε)=O(γ1tg)
(99)

4.9.3. XX̅ Sequence

We solve for a simplified model in which Tϕ = 2T1, that is, γ1 = γ2: all the components of the Bloch vector decay uniformly to the steady state (0, 0, η). The resulting Bloch vector equations are
v˙(t)={G+v(t)+c,if(2k)tg<t(2k+1)tg,kZGv(t)+c,if(2k1)tg<t(2k)tg,kZ
(100)
where c = (0, 0, η γ), γ = γ1 = γ2, and the matrices G± are
G±=(γΔerr0Δerrγεtot0±εtotγ)
(101)
The solutions for the Bloch vector after each XX sequence (which corresponds to a duration of 2tg) are then given recursively by
v(2(n+1)tg)=Av(2ntg)ζ
(102)
where ζR3, and
A=eGtgeG+tgζ=(IeGtg)(G)1c+eGtg(IeG+tg)(G+)1c
(103)
Unraveling the recursion yields
v(2ntg)=Anv(0)k=0n1Akζ
(104)
We are interested in calculating only vx(2ntg):
vx(2ntg)=(1,0,0)·Anv(0)k=0n1(1,0,0)·Akζ
(105)
The required elements of the matrix An can be written as
(An)11=12(p2n+q2n)(An)12=iω(1+eiωtg)2χ(p2nq2n)(An)13=εtot(1eiωtg)2χ(p2nq2n)
(106)
where
ωεtot2+Δerr2χ4eiωtgω2+(1eiωtg)2Δerr2peγtgiωtg2ω2((1+e2iωtg)Δerr2+2eiωtgεtot2+(1eiωtg)Δerrχ)qeγtgiωtg2ω2((1+e2iωtg)Δerr2+2eiωtgεtot2(1eiωtg)Δerrχ)
Substituting these into eq 105 gives
vx(2ntg)=12(p2n+q2n)+1p2n2(1p)(ζxi(1+eiωtg)ωχζy(1eiωtg)εtotχζz)+1q2n2(1q)(ζx+i(1+eiωtg)ωχζy+(1eiωtg)εtotχζz)
where ζ carries the temperature dependence. Under first order approximations, χ ≈ 2eiω′tg/2ω′, ω′εtot, and
peγtgiϑqeγtg+iϑϑtan1(2sin(ωtg2)Δerrω)
(107)
The values of ζ’s components are lengthy expressions and are given in Appendix C so as not to interrupt the flow. Under first-order approximations, ζx ≈ 0, (1 + eiω′tgy/χ ≈ 0 and ζz ≈ 0. Substituting these values gives the following final expression:
vx(2ntg)12(p2n+q2n)=e2nγtgcos2nϑ
(108)
The expression for the fidelity with the |+⟩ state is, therefore
F(tn)=12+12eγtncos(2nϑ)
This expression matches eq 76 with 2ωtg = ϑ, TD = 1/γ and a = 0. Moreover, the oscillation frequency matches the closed system fidelity oscillation frequency [eq 78] up to the first order (noting that θerr = ω′tg).
Solving the general case where γ1 ≠ γ2 is difficult: we cannot apply perturbative techniques directly to bound the deviation of the Bloch vector trajectory in a {γ1 ≠ γ2, Δerr ≠ 0} environment (where vx exhibits oscillating exponential decay) from a Bloch vector trajectory in a {γ1 ≠ γ2, Δerr = 0} environment (where vx exhibits exponential decay). We rely instead on numerical simulations of the Lindblad equation.

4.10. Finite Temperature

We discuss numerical, experimental, and analytical approaches for accounting for finite-temperature (T > 0) effects.

4.10.1. Numerical

In our simulations, we take the zero temperature limit, which fits our experimental data well and is typically valid for transmons. However, it is straightforward to generalize to a nonzero temperature. For zero temperature, we use a Lindblad operator L1 = σ = |0⟩⟨1| with rate γ1 = 1/T1. For a nonzero temperature T, we replace this with two operators L+ = σ = |0⟩⟨1| and L = σ+ = |1⟩⟨0|, with rates γ and γ+, respectively. The rates obey γ + γ+ = γ1 = 1/T1 and the detailed balance condition γ+ = exp(−ℏω/kBT), (64) where ω is the qubit frequency. Varying T in our numerical model thus allows us to fit dynamics at any temperature. In particular, relaxation-sensitive sequences, such as YY̅ applied to |+⟩, saturate to a fidelity that strongly depends on the device temperature.

4.10.2. Experimental

The effect of finite temperature can also be seen in our DB protocol. In general, state preparation and measurement (SPAM) errors will cause the initial state fidelity to be less than 1. We can calibrate these errors out by defining the last point in the T1 sequence (i.e., the thermally equilibrated state) to be a fidelity of 0, and the first point (i.e., the population-inverted thermal state) to be a fidelity of 1. Such a definition accounts for all temperature effects in state preparation and all measurement errors, rescaling all fidelities by the same multiplicative factor, and thus leaving the fitting and analysis unaffected. If we then compare the saturation fidelities of the XX, YY, and YY sequences to the midpoint of the T1 fidelities, we see that they differ in a way that depends on the temperature, all saturating to a fidelity of 0.5 in the limit of infinite temperature. We can thus use DB to extract the device temperature, although using the XX sequence (with saturation close to the mixed state) will be less sensitive to small temperature variations than using YY and YY, or typical Rabi-based methods. (201) Nevertheless, our original 4-step protocol captures the effect of finite temperature. Note that this technique is insensitive to measurement error, which can be significant and would disrupt temperature measurements based on single-shot state-assignment readout.

4.10.3. Analytical

Given a in eq 76 and square pulses, in fact, the temperature can be computed without simulation of the Lindblad master equation. This can be shown in the following way: suppose that the Bloch vector equation is the same as eq 100, but with G+ and G replaced with G1 and G2, respectively. G1 and G2 correspond to the two pulses chosen for the DD sequence; e.g., in an XX sequence, G1 = G2 = G+. Note that G1 and G2 carry no temperature dependence. Then, similar to the analysis in Section 4.9.3, the Bloch vector after n pulse-pairs is given by
v(2ntg)=Bnv(0)ηk=0n1Bkζ
(109)
where similar to eq 103,
B=eG2tgeG1tgζ=(IeG2tg)(G2)1c+eG2tg(IeG1tg)(G1)1c
and c = (0, 0, γ1). The fidelity is then
F(n)=1+v(0)·Bnv(0)ηk=0n1v(0)·Bkζ2
(110)
where again η is the thermal factor defined in eq 84. For n, ∥Bn∥→ 0 from eq 95 and
F()=12(1ηv(0)·(IB)1ζ)
(111)
where we used ∑k=0 Bk = (IB)−1. Substituting F() = (1 + a)/2 from eq 76 gives
a=ηv(0)·(BI)1ζ
(112)
The LHS of eq 112 is measured experimentally, while the inner product in the RHS is computed using the parameters {T1, T2, εtot, Δerr}.
Note that a is maximized when η = 1, i.e., at zero temperature [eq 84]. Accordingly, substituting aT=0 = (v(0)·(BI)−1ζ) into eq 112 and rearranging eq 84 gives the temperature as
kBT=ω01(ln(aT=0+aaT=0a))1
(113)
As mentioned above, for transmons ω01 ∈ [3 – 6]GHz and T ∼ 10 mK, so aaT=0. However, for devices where finite temperature effects are significant, a < aT=0; moreover a ≈ 0 for T.
Since the experimental error in calculating a is largely unaffected by the type of pulse sequence (i.e., by the value of a itself), it is advisable to select pulse sequences that have a higher value of |aT=0| to minimize the fractional uncertainty in temperature T. E.g., a YY sequence which gives a simulated value of |aYY| would be preferred over a YY sequence which gives |aYY| if |aYY| > |aYY|.

4.11. Leakage

Leakage to a state outside the qubit subspace is a coherent error as long as the leakage state’s coherence is long-lived compared to the gate duration. The coherent rotation to the leakage state will lead to oscillations of the fidelity. These oscillations will, in principle, show up with any pulse sequence, but they are difficult to distinguish from the oscillations due to rotation errors in the YY sequence or phase errors in the XX sequence. For this reason it is best to look at the XX sequence, which has no oscillations in the absence of leakage. In this case the functional form eq 76 no longer holds, as the amplitude of the oscillations and the amplitude of the decay are decoupled from each other. Because leakage is a dynamical effect that depends on the pulse shape, and because the leakage state may influence the readout signal, no simple analytical formula can capture the effect of leakage on fidelity. Qualitatively, however, it is easy to see that more leakage leads to larger and faster fidelity oscillations. For a simple example, see Figure 10. Here we have done a closed-system simulation of a qubit with a leakage state that is −150 MHz detuned from the qubit transition and has a drive matrix element 2 times larger─i.e., a qutrit similar to a transmon. (202,203) The simulated pulses have cosine envelopes with durations of 10 or 20 ns (orange squares and blue circles, respectively). When prepared in |+⟩ and driven with the XX sequence, fidelity oscillations appear, with amplitude and frequency that increase as the gate time is decreased (gate bandwidth is increased). Note that these are pulses without DRAG (190) and so leakage can be quite significant. Adding DRAG with appropriate parameters can remove these oscillations. In our experiments the pulses were much slower relative to the anharmonicity and DRAG removed any residual leakage, so we ignore leakage in the experimental analysis.

Figure 10

Figure 10. Simulated survival fidelity of the |+⟩ state after repetitions of the XX sequence, where each gate is a cosine pulse length of 20 ns (blue circles) or 10 ns (orange squares). A third state in the simulation represents leakage, with a detuning (anharmonicity) of −150 MHz relative to the qubit transition. Both the amplitude and frequency of oscillations increase as the gate duration is decreased and leakage becomes more significant.

Because the XX sequence on |+⟩ is unaffected by rotation and phase errors, fidelity oscillations in XX provide a direct diagnostic of the amount of leakage. This is particularly useful in the case where the leakage state is unknown and so measurement of the leaked population may not be possible. While no easy analytical solution presents itself, it should be possible to numerically model a gate sequence and fit the measured data to extract parameters of the leakage state.

4.12. DB Summary and Future Research Directions

DB seeks to strike a balance between the amount of information collected about the noise and efficiency of the process to obtain that information. Moreover, DB highlights noise bias, which might benefit the development of compilation strategies (204) provided circuits can be compiled so as to minimize relaxation by steering quantum trajectories away from excited states. The benefits could extend to resource estimation for fault-tolerant quantum algorithms (205,206) and biased noise error correction. (86) The implementation of the DB protocol is straightforward and adaptable beyond superconducting qubits. We anticipate that DB will contribute to the benchmarking and enhancement of quantum gates across various platforms. A natural open problem is to extend DB to two-qubit gates and identify the corresponding minimal extended parameter set that describes incoherent and coherent errors.

5. Conclusions and Outlook

Click to copy section linkSection link copied!

Gate benchmarking techniques are crucial for practical quantum computing, fulfilling two key roles: (1) offering realistic estimates of attainable performance with existing quantum hardware, and (2) supplying valuable metrics for assessing and monitoring advancements in various gate and circuit implementations. These methods differ widely in their assumptions about the computation and error models, their implementation’s simplicity and cost, and the level of detail they provide about the noise and errors impacting the gates.
This review provides a comprehensive analysis of key benchmarking methodologies for assessing noise in quantum gates and circuits, emphasizing the trade-offs inherent in each strategy. Among the techniques explored are Randomized Benchmarking (RB), well-known for its efficient and scalable computation of average gate fidelity; Quantum Process Tomography, which, while entailing greater implementation costs, yields a complete noise characterization of gates but is sensitive to SPAM; Gate Set Tomography, which has all the ingredients of quantum process tomography but includes state preparation and measurement errors explicitly in its model; Process Fidelity Estimation and Direct Fidelity Estimation, which deliver practical fidelity measures with reduced resource demands and are applicable to both circuits and individual gates; as well as Cross Entropy Benchmarking and Random Circuit Sampling, which are predominantly employed for evaluating complete quantum circuits, particularly in the context of demonstrating quantum supremacy. We also presented a new Deterministic Benchmarking (DB) protocol and demonstrated it using a superconducting transmon qubit. DB is deterministic and efficient: it employs a small and fixed set of simple pulse-pair sequences. The protocol quantifies both incoherent and coherent gate errors that elude RB and does so without averaging over random circuits, in contrast to all the other methods we surveyed, with the exception of QPT.
As quantum computing technology reaches maturity, benchmarking will become a valuable tool for fine-tuning these systems. Currently, however, it is essential for identifying noise and decoherence sources, as well as aiding in the development and optimization of quantum information devices until they achieve reliability and functionality. Ongoing enhancement and advancement of benchmarking methods are pivotal for raising the fidelity of quantum logic gates and circuits to meet the demands of scalable quantum computing architectures.
Ultimately, combining refined noise characterization and error mitigation holds the key to unlocking challenging quantum chemistry problems that extend beyond the scope of classical computation. (21) In the near term, these improvements enable the design of noise-resilient ansätze for VQE, thereby raising the ceiling on the system sizes that can be reliably tackled on noisy intermediate-scale quantum (NISQ) devices. (20) Over the longer horizon, as fault-tolerance thresholds are approached and eventually surpassed, the careful measurement of device-specific noise will continue to inform quantum error correction procedures, bridging the gap between hardware capabilities and the exacting precision required for practical chemical applications. By continuing to improve noise characterization and error mitigation, we inch closer to a regime where quantum hardware can deliver transformative insights into the molecular processes and material properties that govern our physical world.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
    • Eli M. Levenson-Falk - Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United StatesCenter for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United StatesMing Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United StatesOrcidhttps://orcid.org/0000-0002-0019-9465 Email: [email protected]
    • Daniel A. Lidar - Department of Chemistry, University of Southern California, Los Angeles, California 90089, United StatesMing Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United StatesDepartment of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United StatesCenter for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United StatesOrcidhttps://orcid.org/0000-0002-1671-1515 Email: [email protected]
  • Authors
    • Vinay Tripathi - Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United StatesCenter for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United StatesOrcidhttps://orcid.org/0000-0003-1418-9046
    • Daria Kowsari - Department of Physics & Astronomy, University of Southern California, Los Angeles, California 90089, United StatesCenter for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United States
    • Kumar Saurav - Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United StatesMing Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United States
    • Haimeng Zhang - Center for Quantum Information Science and Technology, University of Southern California, Los Angeles, California 90089, United StatesMing Hsieh Department of Electrical & Computer Engineering, University of Southern California, Los Angeles, California 90089, United States
  • Author Contributions

    V.T. and D.K. contributed equally to this work.

  • Notes
    The authors declare no competing financial interest.

Biographies

Click to copy section linkSection link copied!

Vinay Tripathi received his PhD in Physics from the University of Southern California in 2024, focusing on noise characterization and suppression in superconducting quantum processors. Upon graduation, he joined the IBM Quantum team as a Research Scientist, contributing to advancements in quantum computing hardware and noise mitigation techniques.

Daria Kowsari received his PhD in Physics from Washington University in St. Louis in 2023. Upon graduation, he joined USC as a postdoctoral scholar. His research focuses on the realization of novel open quantum systems utilizing qubit processors as well as developing new methods to characterize the errors in quantum gates and circuits.

Kumar Saurav is a PhD student in the ECE department of the University of Southern California and is a Viterbi Graduate Fellowship recipient. Prior to USC, Saurav completed his Bachelor’s degree with Honors in Computer Science from IIT Bombay, following which he worked as a Software Engineer at Google for two years. His current research interests mainly include quantum control and device physics.

Haimeng Zhang received her PhD in Electrical Engineering from the University of Southern California in 2023. Her research focuses on noise characterization and suppression in superconducting qubits. After graduation, she joined IBM as a quantum algorithm engineer, where she works on benchmarking transpilation methods and developing error suppression and mitigation techniques to improve the implementation of quantum algorithms on quantum computers.

Eli M. Levenson-Falk is an Associate Professor in the Department of Physics & Astronomy and the Ming Hsieh Department of Electrical & Computer Engineering at the University of Southern California. He is the principal investigator of the Levenson-Falk Lab (LFL) which conducts experimental research using superconducting circuits. Eli received his PhD in Physics from the University of California Berkeley in 2013. He is a 2018 Air Force Office of Scientific Research Young Investigator, a 2021 Office of Naval Research Young Investigator, and a 2021 Cottrell Scholar.

Daniel A. Lidar is the holder of the Viterbi Professorship of Engineering at the University of Southern California, and a Professor of Electrical & Computer Engineering, Chemistry, and Physics & Astronomy. He holds a PhD in Physics from the Hebrew University of Jerusalem. He did his postdoctoral work in Theoretical Chemistry at UC Berkeley. Prior to joining USC in 2005 he was a faculty member at the University of Toronto. His main research interest is quantum information processing, where he works on quantum control, quantum error correction, the theory of open quantum systems, and quantum algorithms. He is the Director of the USC Center for Quantum Information Science and Technology, the Director of the USC-IBM Quantum Innovation Center, and the co-Director (Scientific Director) of the USC Center for Quantum Computing. He is a recipient of a Sloan Research Fellowship, a Guggenheim Fellowship and is a Fellow of the AAAS, APS, and IEEE.

Acknowledgments

Click to copy section linkSection link copied!

We thank Namit Anand, Nic Ezzell and Alexandru Paler for useful discussions. Devices were fabricated and provided by the Superconducting Qubits at Lincoln Laboratory (SQUILL) Foundry at MIT Lincoln Laboratory, with funding from the Laboratory for Physical Sciences (LPS) Qubit Collaboratory. This research was supported by the National Science Foundation, the Quantum Leap Big Idea under Grant No. OMA-1936388, the U. S. Army Research Laboratory and the U. S. Army Research Office contract/grant number W911NF2310255, and the Intelligence Advanced Research Projects Activity (IARPA) under Cooperative Agreement No. W911NF-23-2-0216.

Appendix

Click to copy section linkSection link copied!

This appendix describes technical details relevant to the Deterministic Benchmarking protocol.

A. Device Design and Experimental Setup

The device used in the experiment consists of a simple grounded Xmon geometry (207) dispersively coupled to its co-planar waveguide (CPW) readout resonator. It is nominally identical to the device reported in Ref (95)., and the design is posted on the open-source SQuADDS database. (208) We make use of a separate weakly-coupled microwave line to drive the qubit, which results in reasonably fast gate times (80 ns) with an average gate fidelity of 99.74%. The device layout was created using the Qiskit Metal open-source package (209) with the finite-element (FE) simulations taking place in the Ansys HFSS simulator to find the various parameters of the device, i.e., the qubit and cavity frequencies (ωq and ωc), as well as the electric field distribution across the device of interest. The FE simulation results are then analyzed using the pyEPR package (210) to extract the qubit anharmonicity (η) and the qubit-cavity dispersive shift (χqc). Lastly, the resonator line width (κ) is extracted using the HFSS driven-modal scattering simulations. Table 2 shows the measured parameters of the device used in this paper.
Table 2. Measured Parameters of the Device Used in the Experiment
ωq/2π|η|/2πχqc/2πωc/2π
4.37 GHz215 MHz230 kHz6.95 GHz
κ/2πT1T2*
150 kHz20 μs32 μs
The device is fabricated and packaged in a copper box. We then mount and thermalize the box in the lowest stage of a dilution refrigerator (DR), which has a base temperature of ∼ 15mK, surrounded by an additional copper shielding as well as a Cryoperm to further isolate the chip from the external infrared radiation and magnetic fields. The readout and drive lines each have 50 and 60 dB of attenuation to reach the single-photon power regime on the chip. Moreover, we use an 8 GHz low-pass filter on each line to protect the device from high-frequency noise, as well as high-frequency Eccosorb filters mounted close to the device in the DR to repel the infrared radiation from the higher stages of the DR. The readout line is protected by using four cryogenic microwave circulators, with the signal being amplified using both a traveling wave parametric amplifier (TWPA) at the mixing chamber and a high-electron-mobility transistor (HEMT) amplifier at the 4K stage to achieve single-shot readout.

B. Conventional Single-Qubit Gate Calibration

In our experiments, the Hx Hamiltonian [eq 74] is fully calibrated and the Hy Hamiltonian is then obtained simply by applying a phase shift via an arbitrary waveform generator (AWG), which is equivalent to Hy=Rz(π2)HxRz(π2), and leaving εerr and Δerr unchanged.
We use a cosine envelope pulse shape for all the experiments presented in this work. All our pulses are 80 ns long with an additional padding of 4 ns on both sides, i.e., tg = 88 ns. We perform the following steps to calibrate the single-qubit gates used in the experiments:
1.

Find the resonance frequencies of the cavity and the qubit using conventional spectroscopy techniques.

2.

Look for a monotone signal sent at the frequency of the qubit (cavity) at the output spectrum of the microwave mixers and try to minimize the leakage and the mirror signals caused by the mixer’s electrical offset and imbalance.

3.

Perform simple Rabi experiments to find the right gate length for a specific gate amplitude. We chose a gate length of 80 ns and a cosine-shaped gate envelope for the experiments reported here. Note that there is a padding of 8 ns between consecutive pulses, and hence the effective pulse interval is 88 ns.

4.

Characterize the relaxation and decoherence rates of the device and use the Ramsey experiments to fine-tune the qubit’s frequency.

5.

Optimize the readout fidelity by using the previously calibrated Rx(π) pulses and varying the readout tone’s frequency and amplitude.

6.

Apply consecutive Rx(π) and Rx(π/2) pulses to the system initialized in the ground state, such that after each measurement the qubit ideally returns to its ground state, i.e., 4N × Rx(π/2) and 2N × Rx(π) pulses with N ∈ [1, ..., 20]. Using this method we can observe the ground state fidelity of the qubit and pick an amplitude that pins the qubit to its ground state without exhibiting any oscillations. As described above, calibrating the pulses along the x-axis automatically calibrates them along the y-axis as well.

7.

Optimize the DRAG weighting parameter (α) by applying multiple pairs of Rx(π)Rx(−π) pulses to the system initialized in its ground state, (211) observe the ground state fidelity, and choose the parameter values for which the qubit remains in its ground state.

C. ζ’s Components

The values of ζ’s components [defined in eq 103], along with their first order approximations (Δerr, γ, εerrε = π/tg) are noted below:
ζx=4ηe2γtgΔerrεtotω4(ω2+γ2)[ω2((γsinωtg2+(1eγtg)2ωcosωtg2)2+ω24((1eγtg)sinωtg2)2)Δerr2(1cos(ωtg))((γsinωtg2+12ωcosωtg2)2+ω24(2(1eγtg)cos2ωtg2))]0
ζy=ηe2γtgωεtotω4(ω2+γ2)[4Δerr2sin(ωtg)((γsinωtg2+12ωcosωtg2)2+ω24(2(1eγtg)cos2ωtg2))+γω3((eγtgcos(ωtg))2+sin2(ωtg))](1+eγtg)2γε
ζz=ηe2γtgω4(ω2+γ2)[8Δerr2εtot2sin2ωtg2((γsinωtg2+12ωcosωtg2)2ω24cos2ωtg2)+eγtgγω3(eγtgγω+2εtot2sin(ωtg))+Δerr2ω2(eγtg1)(eγtgω2+(2cos(ωtg)1)εtot2+Δerr2)]0

References

Click to copy section linkSection link copied!

This article references 211 other publications.

  1. 1
    Bharti, K.; Cervera-Lierta, A.; Kyaw, T. H.; Haug, T.; Alperin-Lea, S.; Anand, A.; Degroote, M.; Heimonen, H.; Kottmann, J. S.; Menke, T.; Mok, W.-K.; Sim, S.; Kwek, L.-C.; Aspuru-Guzik, A. Noisy intermediate-scale quantum algorithms. Rev. Mod. Phys. 2022, 94, 015004  DOI: 10.1103/RevModPhys.94.015004
  2. 2
    Dalzell, A. M.; McArdle, S.; Berta, M.; Bienias, P.; Chen, C.-F.; Gilyén, A.; Hann, C. T.; Kastoryano, M. J.; Khabiboulline, E. T.; Kubica, A.; Salton, G.; Wang, S.; Brandão, F. G. S. L. Quantum algorithms: A survey of applications and end-to-end complexities. 2023; https://arxiv.org/abs/2310.03011 (accessed 2025-02-23).
  3. 3
    Shor, P. Fault-tolerant quantum computation. Proceedings of 37th Conference on Foundations of Computer Science. 1996; pp 5665.
  4. 4
    Aharonov, D.; Ben-Or, M. Fault tolerant quantum computation with constant error. Proceedings of 29th Annual ACM Symposium on Theory of Computing (STOC). New York, NY, 1997; p 176.
  5. 5
    Knill, E.; Laflamme, R.; Zurek, W. H. Resilient Quantum Computation. Science 1998, 279, 342345,  DOI: 10.1126/science.279.5349.342
  6. 6
    Aliferis, P.; Gottesman, D.; Preskill, J. Quantum accuracy threshold for concatenated distance-3 codes. Quant. Inf. Comput. 2006, 6, 97,  DOI: 10.26421/QIC6.2-1
  7. 7
    Campbell, E. T.; Terhal, B. M.; Vuillot, C. Roads towards fault-tolerant universal quantum computation. Nature 2017, 549, 172,  DOI: 10.1038/nature23460
  8. 8
    Reiher, M.; Wiebe, N.; Svore, K. M.; Wecker, D.; Troyer, M. Elucidating reaction mechanisms on quantum computers. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 75557560,  DOI: 10.1073/pnas.1619152114
  9. 9
    Kassal, I.; Whitfield, J. D.; Perdomo-Ortiz, A.; Yung, M.-H.; Aspuru-Guzik, A. Simulating Chemistry Using Quantum Computers. Annu. Rev. Phys. Chem. 2011, 62, 185207,  DOI: 10.1146/annurev-physchem-032210-103512
  10. 10
    Preskill, J. Quantum Computing in the NISQ era and beyond. Quantum 2018, 2, 79,  DOI: 10.22331/q-2018-08-06-79
  11. 11
    O’Malley, P. J. J. Scalable Quantum Simulation of Molecular Energies. Physical Review X 2016, 6, 031007  DOI: 10.1103/PhysRevX.6.031007
  12. 12
    Kandala, A.; Mezzacapo, A.; Temme, K.; Takita, M.; Brink, M.; Chow, J. M.; Gambetta, J. M. Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets. Nature 2017, 549, 242,  DOI: 10.1038/nature23879
  13. 13
    Lidar, D. A.; Wang, H. Calculating the thermal rate constant with exponential speedup on a quantum computer. Phys. Rev. E 1999, 59, 24292438,  DOI: 10.1103/PhysRevE.59.2429
  14. 14
    Aspuru-Guzik, A.; Dutoi, A. D.; Love, P. J.; Head-Gordon, M. Simulated Quantum Computation of Molecular Energies. Science 2005, 309, 1704,  DOI: 10.1126/science.1113479
  15. 15
    Peruzzo, A.; McClean, J.; Shadbolt, P.; Yung, M.-H.; Zhou, X.-Q.; Love, P. J.; Aspuru-Guzik, A.; O’Brien, J. L. A variational eigenvalue solver on a photonic quantum processor. Nat. Commun. 2014, 5, 4213,  DOI: 10.1038/ncomms5213
  16. 16
    McClean, J. R.; Romero, J.; Babbush, R.; Aspuru-Guzik, A. The theory of variational hybrid quantum-classical algorithms. New J. Phys. 2016, 18, 023023  DOI: 10.1088/1367-2630/18/2/023023
  17. 17
    McArdle, S.; Endo, S.; Aspuru-Guzik, A.; Benjamin, S. C.; Yuan, X. Quantum computational chemistry. Rev. Mod. Phys. 2020, 92, 015003  DOI: 10.1103/RevModPhys.92.015003
  18. 18
    Tilly, J.; Chen, H.; Cao, S.; Picozzi, D.; Setia, K.; Li, Y.; Grant, E.; Wossnig, L.; Rungger, I.; Booth, G. H.; Tennyson, J. The Variational Quantum Eigensolver: A review of methods and best practices. Phys. Rep. 2022, 986, 1128,  DOI: 10.1016/j.physrep.2022.08.003
  19. 19
    Kitaev, A. Y. Quantum measurements and the Abelian Stabilizer Problem. 1995; https://arxiv.org/abs/quant-ph/9511026 (accessed 2025-02-23).
  20. 20
    Cao, Y.; Romero, J.; Olson, J. P.; Degroote, M.; Johnson, P. D.; Kieferová, M.; Kivlichan, I. D.; Menke, T.; Peropadre, B.; Sawaya, N. P. D.; Sim, S.; Veis, L.; Aspuru-Guzik, A. Quantum Chemistry in the Age of Quantum Computing. Chem. Rev. 2019, 119, 1085610915,  DOI: 10.1021/acs.chemrev.8b00803
  21. 21
    Lee, S. Evaluating the evidence for exponential quantum advantage in ground-state quantum chemistry. Nat. Commun. 2023, 14, 1952,  DOI: 10.1038/s41467-023-37587-6
  22. 22
    Huggins, W. J.; McClean, J. R.; Rubin, N. C.; Jiang, Z.; Wiebe, N.; Whaley, K. B.; Babbush, R. Efficient and noise resilient measurements for quantum chemistry on near-term quantum computers. npj Quantum Information 2021, 7, 23,  DOI: 10.1038/s41534-020-00341-7
  23. 23
    Hashim, A. . A Practical Introduction to Benchmarking and Characterization of Quantum Computers. 2024; https://arxiv.org/abs/2408.12064 (accessed 2025-02-23).
  24. 24
    Chuang, I. L.; Nielsen, M. A. Prescription for experimental determination of the dynamics of a quantum black box. J. Mod. Opt. 1997, 44, 2455,  DOI: 10.1080/09500349708231894
  25. 25
    Poyatos, J. F.; Cirac, J. I.; Zoller, P. Complete Characterization of a Quantum Process: The Two-Bit Quantum Gate. Phys. Rev. Lett. 1997, 78, 390393,  DOI: 10.1103/PhysRevLett.78.390
  26. 26
    Mohseni, M.; Rezakhani, A. T.; Lidar, D. A. Quantum-process tomography: Resource analysis of different strategies. Phys. Rev. A 2008, 77, 032322  DOI: 10.1103/PhysRevA.77.032322
  27. 27
    Emerson, J.; Alicki, R.; Życzkowski, K. Scalable noise estimation with random unitary operators. Journal of Optics B: Quantum and Semiclassical Optics 2005, 7, S347,  DOI: 10.1088/1464-4266/7/10/021
  28. 28
    Knill, E.; Leibfried, D.; Reichle, R.; Britton, J.; Blakestad, R. B.; Jost, J. D.; Langer, C.; Ozeri, R.; Seidelin, S.; Wineland, D. J. Randomized benchmarking of quantum gates. Phys. Rev. A 2008, 77, 012307  DOI: 10.1103/PhysRevA.77.012307
  29. 29
    Magesan, E.; Gambetta, J. M.; Emerson, J. Scalable and Robust Randomized Benchmarking of Quantum Processes. Phys. Rev. Lett. 2011, 106, 180504,  DOI: 10.1103/PhysRevLett.106.180504
  30. 30
    Helsen, J.; Roth, I.; Onorati, E.; Werner, A. H.; Eisert, J. General Framework for Randomized Benchmarking. PRX Quantum 2022, 3, 020357  DOI: 10.1103/PRXQuantum.3.020357
  31. 31
    Merkel, S. T.; Gambetta, J. M.; Smolin, J. A.; Poletto, S.; Córcoles, A. D.; Johnson, B. R.; Ryan, C. A.; Steffen, M. Self-consistent quantum process tomography. Phys. Rev. A 2013, 87, 062119  DOI: 10.1103/PhysRevA.87.062119
  32. 32
    Blume-Kohout, R.; Gamble, J. K.; Nielsen, E.; Mizrahi, J.; Sterk, J. D.; Maunz, P. Robust, self-consistent, closed-form tomography of quantum logic gates on a trapped ion qubit. 2013; https://arxiv.org/abs/1310.4492 (accessed 2025-02-23).
  33. 33
    Kim, D.; Ward, D. R.; Simmons, C. B.; Gamble, J. K.; Blume-Kohout, R.; Nielsen, E.; Savage, D. E.; Lagally, M. G.; Friesen, M.; Coppersmith, S. N.; Eriksson, M. A. Microwave-driven coherent operation of a semiconductor quantum dot charge qubit. Nat. Nanotechnol. 2015, 10, 243247,  DOI: 10.1038/nnano.2014.336
  34. 34
    Bäumer, E.; Tripathi, V.; Seif, A.; Lidar, D.; Wang, D. S. Quantum Fourier Transform Using Dynamic Circuits. Phys. Rev. Lett. 2024, 133, 150602150609,  DOI: 10.1103/PhysRevLett.133.150602
  35. 35
    da Silva, M. P.; Landon-Cardinal, O.; Poulin, D. Practical characterization of quantum devices without tomography. Phys. Rev. Lett. 2011, 107, 210404210409,  DOI: 10.1103/PhysRevLett.107.210404
  36. 36
    Flammia, S. T.; Liu, Y.-K. Direct Fidelity Estimation from Few Pauli Measurements. Phys. Rev. Lett. 2011, 106, 230501,  DOI: 10.1103/PhysRevLett.106.230501
  37. 37
    Neill, C. A blueprint for demonstrating quantum supremacy with superconducting qubits. Science 2018, 360, 195199,  DOI: 10.1126/science.aao4309
  38. 38
    Boixo, S.; Isakov, S. V.; Smelyanskiy, V. N.; Babbush, R.; Ding, N.; Jiang, Z.; Bremner, M. J.; Martinis, J. M.; Neven, H. Characterizing quantum supremacy in near-term devices. Nature Physics 2018, 14, 595600
  39. 39
    Arute, F. Quantum supremacy using a programmable superconducting processor. Nature 2019, 574, 505510
  40. 40
    Kueng, R.; Long, D. M.; Doherty, A. C.; Flammia, S. T. Comparing Experiments to the Fault-Tolerance Threshold. Phys. Rev. Lett. 2016, 117, 170502,  DOI: 10.1103/PhysRevLett.117.170502
  41. 41
    Barnes, J. P.; Trout, C. J.; Lucarelli, D.; Clader, B. D. Quantum error-correction failure distributions: Comparison of coherent and stochastic error models. Phys. Rev. A 2017, 95, 062338  DOI: 10.1103/PhysRevA.95.062338
  42. 42
    Ng, H. K.; Preskill, J. Fault-tolerant quantum computation versus Gaussian noise. Phys. Rev. A 2009, 79, 032318. DOI: 10.1103/PhysRevA.79.032318
  43. 43
    Wallman, J.; Granade, C.; Harper, R.; Flammia, S. T. Estimating the coherence of noise. New J. Phys. 2015, 17, 113020,  DOI: 10.1088/1367-2630/17/11/113020
  44. 44
    Sanders, Y. R.; Wallman, J. J.; Sanders, B. C. Bounding quantum gate error rate based on reported average fidelity. New J. Phys. 2016, 18, 012002  DOI: 10.1088/1367-2630/18/1/012002
  45. 45
    Hashim, A.; Seritan, S.; Proctor, T.; Rudinger, K.; Goss, N.; Naik, R. K.; Kreikebaum, J. M.; Santiago, D. I.; Siddiqi, I. Benchmarking quantum logic operations relative to thresholds for fault tolerance. npj Quantum Information 2023, 9, 109,  DOI: 10.1038/s41534-023-00764-y
  46. 46
    Wallman, J. J.; Emerson, J. Noise tailoring for scalable quantum computation via randomized compiling. Phys. Rev. A 2016, 94, 052325052334,  DOI: 10.1103/PhysRevA.94.052325
  47. 47
    Hashim, A.; Naik, R. K.; Morvan, A.; Ville, J.-L.; Mitchell, B.; Kreikebaum, J. M.; Davis, M.; Smith, E.; Iancu, C.; O’Brien, K. P.; Hincks, I.; Wallman, J. J.; Emerson, J.; Siddiqi, I. Randomized Compiling for Scalable Quantum Computing on a Noisy Superconducting Quantum Processor. Phys. Rev. X 2021, 11, 041039  DOI: 10.1103/PhysRevX.11.041039
  48. 48
    van den Berg, E.; Minev, Z. K.; Kandala, A.; Temme, K. Probabilistic error cancellation with sparse Pauli–Lindblad models on noisy quantum processors. Nat. Phys. 2023, 19, 11161121,  DOI: 10.1038/s41567-023-02042-2
  49. 49
    Seif, A.; Liao, H.; Tripathi, V.; Krsulich, K.; Malekakhlagh, M.; Amico, M.; Jurcevic, P.; Javadi-Abhari, A. Suppressing Correlated Noise in Quantum Computers via Context-Aware Compiling. 51st Annual International Symposium on Computer Architecture (ISCA) 2024, 310324
  50. 50
    Gross, J. A.; Genois, E.; Debroy, D. M.; Zhang, Y.; Mruczkiewicz, W.; Cian, Z.-P.; Jiang, Z. Characterizing Coherent Errors using Matrix-Element Amplification. npj Quantum Information 2024, 10, 123,  DOI: 10.1038/s41534-024-00917-7
  51. 51
    Bravyi, S.; Englbrecht, M.; König, R.; Peard, N. Correcting coherent errors with surface codes. npj Quantum Information 2018, 4, 55,  DOI: 10.1038/s41534-018-0106-y
  52. 52
    Barenco, A.; Bennett, C. H.; Cleve, R.; DiVincenzo, D. P.; Margolus, N.; Shor, P.; Sleator, T.; Smolin, J. A.; Weinfurter, H. Elementary gates for quantum computation. Phys. Rev. A 1995, 52, 34573467,  DOI: 10.1103/PhysRevA.52.3457
  53. 53
    Erhard, A.; Wallman, J. J.; Postler, L.; Meth, M.; Stricker, R.; Martinez, E. A.; Schindler, P.; Monz, T.; Emerson, J.; Blatt, R. Characterizing large-scale quantum computers via cycle benchmarking. Nat. Commun. 2019, 10, 5347,  DOI: 10.1038/s41467-019-13068-7
  54. 54
    McKay, D. C.; Wood, C. J.; Sheldon, S.; Chow, J. M.; Gambetta, J. M. Efficient Z gates for quantum computing. Phys. Rev. A 2017, 96, 022330  DOI: 10.1103/PhysRevA.96.022330
  55. 55
    Cross, A. W.; Bishop, L. S.; Sheldon, S.; Nation, P. D.; Gambetta, J. M. Validating quantum computers using randomized model circuits. Phys. Rev. A 2019, 100, 032328  DOI: 10.1103/PhysRevA.100.032328
  56. 56
    Dankert, C.; Cleve, R.; Emerson, J.; Livine, E. Exact and approximate unitary 2-designs and their application to fidelity estimation. Phys. Rev. A 2009, 80, 012304  DOI: 10.1103/PhysRevA.80.012304
  57. 57
    Proctor, T.; Rudinger, K.; Young, K.; Nielsen, E.; Blume-Kohout, R. Measuring the capabilities of quantum computers. Nat. Phys. 2022, 18, 7579,  DOI: 10.1038/s41567-021-01409-7
  58. 58
    Cincio, L.; Subası, Y.; Sornborger, A. T; Coles, P. J Learning the quantum algorithm for state overlap. New J. Phys. 2018, 20, 113022,  DOI: 10.1088/1367-2630/aae94a
  59. 59
    Reich, D. M.; Gualdi, G.; Koch, C. P. Optimal Strategies for Estimating the Average Fidelity of Quantum Gates. Phys. Rev. Lett. 2013, 111, 200401200406,  DOI: 10.1103/PhysRevLett.111.200401
  60. 60
    Torlai, G.; Mazzola, G.; Carrasquilla, J.; Troyer, M.; Melko, R.; Carleo, G. Neural-network quantum state tomography. Nat. Phys. 2018, 14, 447450,  DOI: 10.1038/s41567-018-0048-5
  61. 61
    Koutny, D.; Motka, L.; Hradil, Z.; Rehacek, J.; Sanchez-Soto, L. L. Neural-network quantum state tomography. Phys. Rev. A 2022, 106. DOI: 10.1103/PhysRevA.106.012409
  62. 62
    Gu, Y.; Zhuang, W.-F.; Chai, X.; Liu, D. E. Benchmarking universal quantum gates via channel spectrum. Nat. Commun. 2023, 14, 5880,  DOI: 10.1038/s41467-023-41598-8
  63. 63
    McKay, D. C.; Hincks, I.; Pritchett, E. J.; Carroll, M.; Govia, L. C. G.; Merkel, S. T. Benchmarking Quantum Processor Performance at Scale. 2023; https://arxiv.org/abs/2311.05933 (accessed: 2025-02-23).
  64. 64
    Breuer, H.-P.; Petruccione, F. The Theory of Open Quantum Systems; Oxford University Press: Oxford, 2002.
  65. 65
    Kraus, K. States, Effects and Operations; Fundamental Notions of Quantum Theory; Academic, 1983.
  66. 66
    Nielsen, M. A.; Chuang, I. L. Quantum computation and quantum information; Cambridge University Press, 2010.
  67. 67
    Jamiołkowski, A. Linear transformations which preserve trace and positive semidefiniteness of operators. Reports on Mathematical Physics 1972, 3, 275278,  DOI: 10.1016/0034-4877(72)90011-0
  68. 68
    Choi, M.-D. Completely positive linear maps on complex matrices. Linear Algebra and its Applications 1975, 10, 285290,  DOI: 10.1016/0024-3795(75)90075-0
  69. 69
    Uhlmann, A. The “transition probability”in the state space of a ∗-algebra. Reports on Mathematical Physics 1976, 9, 273279,  DOI: 10.1016/0034-4877(76)90060-4
  70. 70
    Knill, E.; Laflamme, R. Theory of quantum error-correcting codes. Phys. Rev. A 1997, 55, 900911,  DOI: 10.1103/PhysRevA.55.900
  71. 71
    Tripathi, V.; Chen, H.; Levenson-Falk, E.; Lidar, D. A. Modeling Low- and High-Frequency Noise in Transmon Qubits with Resource-Efficient Measurement. PRX Quantum 2024, 5, 010320  DOI: 10.1103/PRXQuantum.5.010320
  72. 72
    Byrd, M. S.; Lidar, D. A.; Wu, L.-A.; Zanardi, P. Universal leakage elimination. Phys. Rev. A 2005, 71, 052301  DOI: 10.1103/PhysRevA.71.052301
  73. 73
    Koch, J.; Yu, T. M.; Gambetta, J.; Houck, A. A.; Schuster, D. I.; Majer, J.; Blais, A.; Devoret, M. H.; Girvin, S. M.; Schoelkopf, R. J. Charge-insensitive qubit design derived from the Cooper pair box. Phys. Rev. A 2007, 76, 042319  DOI: 10.1103/PhysRevA.76.042319
  74. 74
    Wu, L.-A.; Walther, P.; Lidar, D. A. No-go theorem for passive single-rail linear optical quantum computing. Sci. Rep. 2013, 3, 1394,  DOI: 10.1038/srep01394
  75. 75
    Levine, H. Demonstrating a Long-Coherence Dual-Rail Erasure Qubit Using Tunable Transmons. Physical Review X 2024, 14, 011051011072,  DOI: 10.1103/PhysRevX.14.011051
  76. 76
    Grünzweig, T.; Hilliard, A.; McGovern, M.; Andersen, M. F. Near-deterministic preparation of a single atom in an optical microtrap. Nat. Phys. 2010, 6, 951954,  DOI: 10.1038/nphys1778
  77. 77
    Wu, Y.; Kolkowitz, S.; Puri, S.; Thompson, J. D. Erasure conversion for fault-tolerant quantum computing in alkaline earth Rydberg atom arrays. Nat. Commun. 2022, 13, 4657,  DOI: 10.1038/s41467-022-32094-6
  78. 78
    Varbanov, B. M.; Battistel, F.; Tarasinski, B. M.; Ostroukh, V. P.; O’Brien, T. E.; DiCarlo, L.; Terhal, B. M. Leakage detection for a transmon-based surface code. npj Quantum Information 2020, 6, 102,  DOI: 10.1038/s41534-020-00330-w
  79. 79
    McEwen, M. Removing leakage-induced correlated errors in superconducting quantum error correction. Nat. Commun. 2021, 12, 1761,  DOI: 10.1038/s41467-021-21982-y
  80. 80
    Gourlay, I.; Snowdon, J. F. Concatenated Coding in the Presence of Dephasing. Phys. Rev. A 2000, 62, 022308  DOI: 10.1103/PhysRevA.62.022308
  81. 81
    Evans, Z. W. E.; Stephens, A. M.; Cole, J. H.; Hollenberg, L. C. L. Error Correction Optimisation in the Presence of X/Z Asymmetry. 2007.
  82. 82
    Stephens, A. M.; Evans, Z. W. E.; Devitt, S. J.; Hollenberg, L. C. L. Asymmetric Quantum Error Correction via Code Conversion. Phys. Rev. A 2008, 77, 062335  DOI: 10.1103/PhysRevA.77.062335
  83. 83
    Aliferis, P.; Preskill, J. Fault-tolerant quantum computation against biased noise. Phys. Rev. A 2008, 78, 052331  DOI: 10.1103/PhysRevA.78.052331
  84. 84
    Stephens, A. M.; Munro, W. J.; Nemoto, K. High-Threshold Topological Quantum Error Correction against Biased Noise. Phys. Rev. A 2013, 88, 060301  DOI: 10.1103/PhysRevA.88.060301
  85. 85
    Brooks, P.; Preskill, J. Fault-Tolerant Quantum Computation with Asymmetric Bacon-Shor Codes. Phys. Rev. A 2013, 87, 032310  DOI: 10.1103/PhysRevA.87.032310
  86. 86
    Bonilla Ataides, J. P.; Tuckett, D. K.; Bartlett, S. D.; Flammia, S. T.; Brown, B. J. The XZZX surface code. Nat. Commun. 2021, 12, 2172,  DOI: 10.1038/s41467-021-22274-1
  87. 87
    Chruscinski, D.; Kossakowski, A. Non-Markovian Quantum Dynamics: Local versus Nonlocal. Phys. Rev. Lett. 2010, 104, 070406  DOI: 10.1103/PhysRevLett.104.070406
  88. 88
    Rivas, Á.; Huelga, S. F.; Plenio, M. B. Quantum non-Markovianity: characterization, quantification and detection. Rep. Prog. Phys. 2014, 77, 094001  DOI: 10.1088/0034-4885/77/9/094001
  89. 89
    Shabani, A.; Lidar, D. A. Completely positive post-Markovian master equation via a measurement approach. Phys. Rev. A 2005, 71, 020101  DOI: 10.1103/PhysRevA.71.020101
  90. 90
    Zhang, H.; Pokharel, B.; Levenson-Falk, E.; Lidar, D. Predicting Non-Markovian Superconducting-Qubit Dynamics from Tomographic Reconstruction. Physical Review Applied 2022, 17, 054018  DOI: 10.1103/PhysRevApplied.17.054018
  91. 91
    Zhang, J.; Vala, J.; Sastry, S.; Whaley, K. B. Geometric theory of nonlocal two-qubit operations. Phys. Rev. A 2003, 67, 042313  DOI: 10.1103/PhysRevA.67.042313
  92. 92
    Sete, E. A.; Didier, N.; Chen, A. Q.; Kulshreshtha, S.; Manenti, R.; Poletto, S. Parametric-Resonance Entangling Gates with a Tunable Coupler. Phys. Rev. Appl. 2021, 16, 024050  DOI: 10.1103/PhysRevApplied.16.024050
  93. 93
    Sete, E. A.; Tripathi, V.; Valery, J. A.; Lidar, D.; Mutus, J. Y. Error Budget of a Parametric Resonance Entangling Gate with a Tunable Coupler. Physical Review Applied 2024, 22, 014059  DOI: 10.1103/PhysRevApplied.22.014059
  94. 94
    Tripathi, V.; Khezri, M.; Korotkov, A. N. Operation and intrinsic error budget of a two-qubit cross-resonance gate. Phys. Rev. A 2019, 100, 012301  DOI: 10.1103/PhysRevA.100.012301
  95. 95
    Gaikwad, C.; Kowsari, D.; Brame, C.; Song, X.; Zhang, H.; Esposito, M.; Ranadive, A.; Cappelli, G.; Roch, N.; Levenson-Falk, E. M.; Murch, K. W. Entanglement Assisted Probe of the Non-Markovian to Markovian Transition in Open Quantum System Dynamics. Phys. Rev. Lett. 2024, 132, 200401,  DOI: 10.1103/PhysRevLett.132.200401
  96. 96
    Dubi, Y.; Di Ventra, M. Relaxation times in an open interacting two-qubit system. Phys. Rev. A 2009, 79, 012328  DOI: 10.1103/PhysRevA.79.012328
  97. 97
    Greenbaum, D. Introduction to Quantum Gate Set Tomography. 2015, arXiv:1509.02921 [quant-ph] (accessed: 2025-02-23).
  98. 98
    Brown, W. G.; Eastin, B. Randomized Benchmarking with Restricted Gate Sets. Phys. Rev. A 2018, 97, 062323  DOI: 10.1103/PhysRevA.97.062323
  99. 99
    Seif, A.; Chen, S.; Majumder, S.; Liao, H.; Wang, D. S.; Malekakhlagh, M.; Javadi-Abhari, A.; Jiang, L.; Minev, Z. K. Entanglement-enhanced learning of quantum processes at scale. 2024; https://arxiv.org/abs/2408.03376 (accessed: 2025-02-23).
  100. 100
    Khan, M. Q.; Dong, W.; Norris, L. M.; Viola, L. Multiaxis quantum noise spectroscopy robust to errors in state preparation and measurement. Physical Review Applied 2024, 22, 024074  DOI: 10.1103/PhysRevApplied.22.024074
  101. 101
    Merkel, S. T.; Gambetta, J. M.; Smolin, J. A.; Poletto, S.; Córcoles, A. D.; Johnson, B. R.; Ryan, C. A.; Steffen, M. Self-consistent quantum process tomography. Phys. Rev. A 2013, 87, 062119  DOI: 10.1103/PhysRevA.87.062119
  102. 102
    Magesan, E.; Gambetta, J. M.; Emerson, J. Characterizing quantum gates via randomized benchmarking. Phys. Rev. A 2012, 85, 042311  DOI: 10.1103/PhysRevA.85.042311
  103. 103
    Aaronson, S.; Gottesman, D. Improved simulation of stabilizer circuits. Phys. Rev. A 2004, 70, 052328  DOI: 10.1103/PhysRevA.70.052328
  104. 104
    Nielsen, M. A. A simple formula for the average gate fidelity of a quantum dynamical operation. Phys. Lett. A 2002, 303, 249252,  DOI: 10.1016/S0375-9601(02)01272-0
  105. 105
    Viola, L.; Knill, E. Random Decoupling Schemes for Quantum Dynamical Control and Error Suppression. Phys. Rev. Lett. 2005, 94, 060502  DOI: 10.1103/PhysRevLett.94.060502
  106. 106
    Santos, L. F.; Viola, L. Dynamical control of qubit coherence: Random versus deterministic schemes. Phys. Rev. A 2005, 72, 062303  DOI: 10.1103/PhysRevA.72.062303
  107. 107
    Santos, L. F.; Viola, L. Enhanced Convergence and Robust Performance of Randomized Dynamical Decoupling. Phys. Rev. Lett. 2006, 97, 150501,  DOI: 10.1103/PhysRevLett.97.150501
  108. 108
    Magesan, E.; Gambetta, J. M.; Johnson, B. R.; Ryan, C. A.; Chow, J. M.; Merkel, S. T.; da Silva, M. P.; Keefe, G. A.; Rothwell, M. B.; Ohki, T. A.; Ketchen, M. B.; Steffen, M. Efficient Measurement of Quantum Gate Error by Interleaved Randomized Benchmarking. Phys. Rev. Lett. 2012, 109, 080505  DOI: 10.1103/PhysRevLett.109.080505
  109. 109
    Gambetta, J. M.; Córcoles, A. D.; Merkel, S. T.; Johnson, B. R.; Smolin, J. A.; Chow, J. M.; Ryan, C. A.; Rigetti, C.; Poletto, S.; Ohki, T. A.; Ketchen, M. B.; Steffen, M. Characterization of Addressability by Simultaneous Randomized Benchmarking. Phys. Rev. Lett. 2012, 109, 240504,  DOI: 10.1103/PhysRevLett.109.240504
  110. 110
    Wallman, J. J.; Barnhill, M.; Emerson, J. Robust characterization of leakage errors. New J. Phys. 2016, 18, 043021  DOI: 10.1088/1367-2630/18/4/043021
  111. 111
    Wood, C. J.; Gambetta, J. M. Quantification and characterization of leakage errors. Phys. Rev. A 2018, 97, 032306  DOI: 10.1103/PhysRevA.97.032306
  112. 112
    Fogarty, M. A.; Veldhorst, M.; Harper, R.; Yang, C. H.; Bartlett, S. D.; Flammia, S. T.; Dzurak, A. S. Nonexponential fidelity decay in randomized benchmarking with low-frequency noise. Phys. Rev. A 2015, 92, 022326  DOI: 10.1103/PhysRevA.92.022326
  113. 113
    Wallman, J. J. Randomized benchmarking with gate-dependent noise. Quantum 2018, 2, 47,  DOI: 10.22331/q-2018-01-29-47
  114. 114
    Proctor, T.; Seritan, S.; Rudinger, K.; Nielsen, E.; Blume-Kohout, R.; Young, K. Scalable Randomized Benchmarking of Quantum Computers Using Mirror Circuits. Phys. Rev. Lett. 2022, 129, 150502,  DOI: 10.1103/PhysRevLett.129.150502
  115. 115
    Helsen, J.; Xue, X.; Vandersypen, L. M. K.; Wehner, S. A new class of efficient randomized benchmarking protocols. npj Quantum Information 2019, 5, 71,  DOI: 10.1038/s41534-019-0182-7
  116. 116
    Helsen, J.; Roth, I.; Onorati, E.; Werner, A.; Eisert, J. General Framework for Randomized Benchmarking. PRX Quantum 2022, 3, 020357  DOI: 10.1103/PRXQuantum.3.020357
  117. 117
    Arrighi, P.; Patricot, C. On quantum operations as quantum states. Annals of Physics 2004, 311, 2652,  DOI: 10.1016/j.aop.2003.11.005
  118. 118
    D’Ariano, G. M.; Lo Presti, P. Quantum Tomography for Measuring Experimentally the Matrix Elements of an Arbitrary Quantum Operation. Phys. Rev. Lett. 2001, 86, 41954198,  DOI: 10.1103/PhysRevLett.86.4195
  119. 119
    Leung, D. W. Choi’s proof as a recipe for quantum process tomography. Journal of Mathematical Physics 2003, 44, 528,  DOI: 10.1063/1.1518554
  120. 120
    Altepeter, J. B.; Branning, D.; Jeffrey, E.; Wei, T. C.; Kwiat, P. G.; Thew, R. T.; O’Brien, J. L.; Nielsen, M. A.; White, A. G. Ancilla-Assisted Quantum Process Tomography. Phys. Rev. Lett. 2003, 90, 193601,  DOI: 10.1103/PhysRevLett.90.193601
  121. 121
    D’Ariano, G. M.; Lo Presti, P. Imprinting Complete Information about a Quantum Channel on its Output State. Phys. Rev. Lett. 2003, 91, 047902  DOI: 10.1103/PhysRevLett.91.047902
  122. 122
    Mohseni, M.; Lidar, D. A. Direct Characterization of Quantum Dynamics. Phys. Rev. Lett. 2006, 97, 170501,  DOI: 10.1103/PhysRevLett.97.170501
  123. 123
    Mohseni, M.; Lidar, D. A. Direct characterization of quantum dynamics: General theory. Phys. Rev. A 2007, 75, 062331  DOI: 10.1103/PhysRevA.75.062331
  124. 124
    Mohseni, M.; Rezakhani, A. T.; Aspuru-Guzik, A. Direct estimation of single- and two-qubit Hamiltonians and relaxation rates. Phys. Rev. A 2008, 77, 042320. DOI: 10.1103/PhysRevA.77.042320
  125. 125
    Graham, T. M.; Barreiro, J. T.; Mohseni, M.; Kwiat, P. G. Hyperentanglement-Enabled Direct Characterization of Quantum Dynamics. Phys. Rev. Lett. 2013, 110, 060404  DOI: 10.1103/PhysRevLett.110.060404
  126. 126
    Nigg, D.; Barreiro, J. T.; Schindler, P.; Mohseni, M.; Monz, T.; Chwalla, M.; Hennrich, M.; Blatt, R. Experimental Characterization of Quantum Dynamics Through Many-Body Interactions. Phys. Rev. Lett. 2013, 110, 060403  DOI: 10.1103/PhysRevLett.110.060403
  127. 127
    Gottesman, D. Class of quantum error-correcting codes saturating the quantum Hamming bound. Phys. Rev. A 1996, 54, 1862,  DOI: 10.1103/PhysRevA.54.1862
  128. 128
    Blume-Kohout, R.; Gamble, J. K.; Nielsen, E.; Rudinger, K.; Mizrahi, J.; Fortier, K.; Maunz, P. Demonstration of qubit operations below a rigorous fault tolerance threshold with gate set tomography. Nat. Commun. 2017, 8, 14485,  DOI: 10.1038/ncomms14485
  129. 129
    Rol, M. A.; Bultink, C. C.; O’Brien, T. E.; de Jong, S. R.; Theis, L. S.; Fu, X.; Luthi, F.; Vermeulen, R. F. L.; de Sterke, J. C.; Bruno, A.; Deurloo, D.; Schouten, R. N.; Wilhelm, F. K.; DiCarlo, L. Restless Tuneup of High-Fidelity Qubit Gates. Phys. Rev. Appl. 2017, 7, 041001  DOI: 10.1103/PhysRevApplied.7.041001
  130. 130
    Mavadia, S.; Edmunds, C. L.; Hempel, C.; Ball, H.; Roy, F.; Stace, T. M.; Biercuk, M. J. Experimental quantum verification in the presence of temporally correlated noise. npj Quantum Information 2018, 4, 7,  DOI: 10.1038/s41534-017-0052-0
  131. 131
    Hempel, C.; Maier, C.; Romero, J.; McClean, J.; Monz, T.; Shen, H.; Jurcevic, P.; Lanyon, B. P.; Love, P.; Babbush, R.; Aspuru-Guzik, A.; Blatt, R.; Roos, C. F. Quantum Chemistry Calculations on a Trapped-Ion Quantum Simulator. Phys. Rev. X 2018, 8, 031022  DOI: 10.1103/PhysRevX.8.031022
  132. 132
    Song, C.; Cui, J.; Wang, H.; Hao, J.; Feng, H.; Li, Y. Quantum computation with universal error mitigation on a superconducting quantum processor. Science Advances 2019, 5, eaaw5686  DOI: 10.1126/sciadv.aaw5686
  133. 133
    Cincio, L.; Rudinger, K.; Sarovar, M.; Coles, P. J. Machine Learning of Noise-Resilient Quantum Circuits. PRX Quantum 2021, 2, 010324  DOI: 10.1103/PRXQuantum.2.010324
  134. 134
    Hong, S. S.; Papageorge, A. T.; Sivarajah, P.; Crossman, G.; Didier, N.; Polloreno, A. M.; Sete, E. A.; Turkowski, S. W.; da Silva, M. P.; Johnson, B. R. Demonstration of a parametrically activated entangling gate protected from flux noise. Phys. Rev. A 2020, 101, 012302  DOI: 10.1103/PhysRevA.101.012302
  135. 135
    Joshi, M. K.; Elben, A.; Vermersch, B.; Brydges, T.; Maier, C.; Zoller, P.; Blatt, R.; Roos, C. F. Quantum Information Scrambling in a Trapped-Ion Quantum Simulator with Tunable Range Interactions. Phys. Rev. Lett. 2020, 124, 240505,  DOI: 10.1103/PhysRevLett.124.240505
  136. 136
    Zhang, S.; Lu, Y.; Zhang, K.; Chen, W.; Li, Y.; Zhang, J.-N.; Kim, K. Error-mitigated quantum gates exceeding physical fidelities in a trapped-ion system. Nat. Commun. 2020, 11, 587,  DOI: 10.1038/s41467-020-14376-z
  137. 137
    Ware, M.; Ribeill, G.; Ristè, D.; Ryan, C. A.; Johnson, B.; da Silva, M. P. Experimental Pauli-frame randomization on a superconducting qubit. Phys. Rev. A 2021, 103, 042604  DOI: 10.1103/PhysRevA.103.042604
  138. 138
    Dehollain, J. P.; Muhonen, J. T.; Blume-Kohout, R.; Rudinger, K. M.; Gamble, J. K.; Nielsen, E.; Laucht, A.; Simmons, S.; Kalra, R.; Dzurak, A. S.; Morello, A. Optimization of a solid-state electron spin qubit using gate set tomography. New J. Phys. 2016, 18, 103018,  DOI: 10.1088/1367-2630/18/10/103018
  139. 139
    Kitaev, A.; Shen, A.; Vyalyi, M. Classical and Quantum Computation; Graduate studies in mathematics; American Mathematical Society, 2002.
  140. 140
    Scott, A. J. Tight informationally complete quantum measurements. Journal of Physics A: Mathematical and General 2006, 39, 13507,  DOI: 10.1088/0305-4470/39/43/009
  141. 141
    Nielsen, E.; Gamble, J. K.; Rudinger, K.; Scholten, T.; Young, K.; Blume-Kohout, R. Gate Set Tomography. Quantum 2021, 5, 557,  DOI: 10.22331/q-2021-10-05-557
  142. 142
    Householder, A. The Theory of Matrices in Numerical Analysis; Dover Books on Mathematics; Dover Publications, 2013.
  143. 143
    Nielsen, E.; Rudinger, K.; Proctor, T.; Russo, A.; Young, K.; Blume-Kohout, R. Probing quantum processor performance with pyGSTi. Quantum Science and Technology 2020, 5, 044002  DOI: 10.1088/2058-9565/ab8aa4
  144. 144
    Proctor, T.; Revelle, M.; Nielsen, E.; Rudinger, K.; Lobser, D.; Maunz, P.; Blume-Kohout, R.; Young, K. Detecting and tracking drift in quantum information processors. Nat. Commun. 2020, 11, 5396,  DOI: 10.1038/s41467-020-19074-4
  145. 145
    Rudinger, K. M.; Ostrove, C. I.; Seritan, S. K.; Grace, M. D.; Nielsen, E.; Blume-Kohout, R. J.; Young, K. C. Two-Qubit Gate Set Tomography with Fewer Circuits. 2023; https://arxiv.org/abs/2307.15767 (accessed: 2025-02-23).
  146. 146
    Horodecki, M.; Horodecki, P.; Horodecki, R. General teleportation channel, singlet fraction, and quasidistillation. Phys. Rev. A 1999, 60, 18881898,  DOI: 10.1103/PhysRevA.60.1888
  147. 147
    Lu, D.; Li, H.; Trottier, D.-A.; Li, J.; Brodutch, A.; Krismanich, A. P.; Ghavami, A.; Dmitrienko, G. I.; Long, G.; Baugh, J.; Laflamme, R. Experimental Estimation of Average Fidelity of a Clifford Gate on a 7-qubit Quantum Processor. Phys. Rev. Lett. 2015, 114, 140505,  DOI: 10.1103/PhysRevLett.114.140505
  148. 148
    Fedorov, A.; Steffen, L.; Baur, M.; da Silva, M. P.; Wallraff, A. Implementation of a Toffoli Gate with Superconducting Circuits. Nature 2012, 481, 170172,  DOI: 10.1038/nature10713
  149. 149
    Wallman, J. J.; Flammia, S. T. Randomized benchmarking with confidence. New J. Phys. 2014, 16, 103032,  DOI: 10.1088/1367-2630/16/10/103032
  150. 150
    Roth, I.; Kueng, R.; Kimmel, S.; Liu, Y.-K.; Gross, D.; Eisert, J.; Kliesch, M. Recovering quantum gates from few average gate fidelities. Phys. Rev. Lett. 2018, 121, 170502,  DOI: 10.1103/PhysRevLett.121.170502
  151. 151
    Kalev, A.; Kyrillidis, A.; Linke, N. M. Validating and certifying stabilizer states. Phys. Rev. A 2019, 99, 042337  DOI: 10.1103/PhysRevA.99.042337
  152. 152
    Seshadri, A.; Ringbauer, M.; Spainhour, J.; Monz, T.; Becker, S. Theory of versatile fidelity estimation with confidence. Phys. Rev. A 2024, 110, 012431  DOI: 10.1103/PhysRevA.110.012431
  153. 153
    Seshadri, A.; Ringbauer, M.; Spainhour, J.; Blatt, R.; Monz, T.; Becker, S. Versatile Fidelity Estimation with Confidence. Phys. Rev. Lett. 2024, 133, 020402  DOI: 10.1103/PhysRevLett.133.020402
  154. 154
    Kosut, R. L.; Rabitz, H.; Grace, M. D. Adaptive quantum control via direct fidelity estimation and indirect model-based parametric process tomography. 52nd IEEE Conference on Decision and Control. Firenze, 2013; pp 12471252.
  155. 155
    Lu, Y.; Sim, J. Y.; Suzuki, J.; Englert, B.-G.; Ng, H. K. Direct estimation of minimum gate fidelity. Phys. Rev. A 2020, 102, 022410  DOI: 10.1103/PhysRevA.102.022410
  156. 156
    Liu, Y.; Otten, M.; Bassirianjahromi, R.; Jiang, L.; Fefferman, B. Benchmarking near-term quantum computers via random circuit sampling. 2022; http://arxiv.org/abs/2105.05232, arXiv: 2105.05232 [quant-ph] (accessed: 2025-02-23).
  157. 157
    Elements of Information Theory; John Wiley & Sons, Ltd, 2005; Chapter 4, pp 71101.
  158. 158
    Ledoux, M. The Concentration of Measure Phenomenon; American Mathematical Soc., 2001; Google-Books-ID: rRKChz4Om6gC.
  159. 159
    Boixo, S.; Smelyanskiy, V. N.; Neven, H. Fourier analysis of sampling from noisy chaotic quantum circuits. 2017; http://arxiv.org/abs/1708.01875, arXiv: 1708.01875 [quant-ph] (accessed: 2025-02-23).
  160. 160
    Choi, J.; Shaw, A. L.; Madjarov, I. S.; Xie, X.; Finkelstein, R.; Covey, J. P.; Cotler, J. S.; Mark, D. K.; Huang, H.-Y.; Kale, A.; Pichler, H.; Brandão, F. G. S. L.; Choi, S.; Endres, M. Preparing random states and benchmarking with many-body quantum chaos. Nature 2023, 613, 468473,  DOI: 10.1038/s41586-022-05442-1
  161. 161
    Rinott, Y.; Shoham, T.; Kalai, G. Statistical Aspects of the Quantum Supremacy Demonstration. Statistical Science 2022, 37, 322347
  162. 162
    Dalzell, A. M.; Hunter-Jones, N.; Brandão, F. G. S. L. Random Quantum Circuits Anticoncentrate in Log Depth. PRX Quantum 2022, 3, 010333
  163. 163
    Huang, C.; Zhang, F.; Newman, M.; Cai, J.; Gao, X.; Tian, Z.; Wu, J.; Xu, H.; Yu, H.; Yuan, B.; Szegedy, M.; Shi, Y.; Chen, J. Classical Simulation of Quantum Supremacy Circuits. arXiv 2020 (accessed 2025-02-23). DOI: 10.48550/arXiv.2005.06787
  164. 164
    Zlokapa, A.; Villalonga, B.; Boixo, S.; Lidar, D. A. Boundaries of quantum supremacy via random circuit sampling. npj Quantum Information 2023, 9, 36,  DOI: 10.1038/s41534-023-00703-x
  165. 165
    Kechedzhi, K.; Isakov, S. V.; Mandrà, S.; Villalonga, B.; Mi, X.; Boixo, S.; Smelyanskiy, V. Effective quantum volume, fidelity and computational cost of noisy quantum processing experiments. Future Generation Computer Systems 2024, 153, 431441,  DOI: 10.1016/j.future.2023.12.002
  166. 166
    Arute, F. Quantum supremacy using a programmable superconducting processor. Nature 2019, 574, 505510,  DOI: 10.1038/s41586-019-1666-5
  167. 167
    Harrow, A. W.; Mehraban, S. Approximate Unitary t-Designs by Short Random Quantum Circuits Using Nearest-Neighbor and Long-Range Gates. Communications in Mathematical Physics 2023, 401, 15311626,  DOI: 10.1007/s00220-023-04675-z
  168. 168
    Haferkamp, J.; Montealegre-Mora, F.; Heinrich, M.; Eisert, J.; Gross, D.; Roth, I. Efficient Unitary Designs with a System-Size Independent Number of Non-Clifford Gates. Communications in Mathematical Physics 2023, 397, 9951041,  DOI: 10.1007/s00220-022-04507-6
  169. 169
    Gao, X.; Kalinowski, M.; Chou, C.-N.; Lukin, M. D.; Barak, B.; Choi, S. Limitations of Linear Cross-Entropy as a Measure for Quantum Advantage. PRX Quantum 2024, 5, 010334  DOI: 10.1103/PRXQuantum.5.010334
  170. 170
    Tripathi, V.; Kowsari, D.; Saurav, K.; Zhang, H.; M.Levenson-Falk, E.; Lidar, D. A. Benchmarking quantum gates and circuits. 2024; https://arxiv.org/abs/2407.09942 (accessed: 2025-02-23).
  171. 171
    Burnett, J. J.; Bengtsson, A.; Scigliuzzo, M.; Niepce, D.; Kudra, M.; Delsing, P.; Bylander, J. Decoherence benchmarking of superconducting qubits. npj Quantum Information 2019, 5, 54,  DOI: 10.1038/s41534-019-0168-5
  172. 172
    Hahn, E. L. Spin Echoes. Phys. Rev. 1950, 80, 580594,  DOI: 10.1103/PhysRev.80.580
  173. 173
    Carr, H. Y.; Purcell, E. M. Effects of Diffusion on Free Precession in Nuclear Magnetic Resonance Experiments. Phys. Rev. 1954, 94, 630638,  DOI: 10.1103/PhysRev.94.630
  174. 174
    Meiboom, S.; Gill, D. Modified Spin?Echo Method for Measuring Nuclear Relaxation Times. Rev. Sci. Instrum. 1958, 29, 688691,  DOI: 10.1063/1.1716296
  175. 175
    Viola, L.; Lloyd, S. Dynamical suppression of decoherence in two-state quantum systems. Phys. Rev. A 1998, 58, 27332744,  DOI: 10.1103/PhysRevA.58.2733
  176. 176
    Uhrig, G. S. Keeping a Quantum Bit Alive by Optimized π-Pulse Sequences. Phys. Rev. Lett. 2007, 98, 100504,  DOI: 10.1103/PhysRevLett.98.100504
  177. 177
    Suter, D.; Álvarez, G. A. Colloquium: Protecting quantum information against environmental noise. Rev. Mod. Phys. 2016, 88, 041001  DOI: 10.1103/RevModPhys.88.041001
  178. 178
    Jurcevic, P. Demonstration of Quantum Volume 64 on a Superconducting Quantum Computing System. Quantum Sci. Technol. 2021, 6, 025020  DOI: 10.1088/2058-9565/abe519
  179. 179
    Pokharel, B.; Lidar, D. A. Demonstration of Algorithmic Quantum Speedup. Phys. Rev. Lett. 2023, 130, 210602,  DOI: 10.1103/PhysRevLett.130.210602
  180. 180
    Tripathi, V.; Chen, H.; Khezri, M.; Yip, K.-W.; Levenson-Falk, E.; Lidar, D. A. Suppression of Crosstalk in Superconducting Qubits Using Dynamical Decoupling. Phys. Rev. Appl. 2022, 18, 024068  DOI: 10.1103/PhysRevApplied.18.024068
  181. 181
    Zhou, Z.; Sitler, R.; Oda, Y.; Schultz, K.; Quiroz, G. Quantum Crosstalk Robust Quantum Control. Phys. Rev. Lett. 2023, 131, 210802,  DOI: 10.1103/PhysRevLett.131.210802
  182. 182
    Bäumer, E.; Tripathi, V.; Wang, D. S.; Rall, P.; Chen, E. H.; Majumder, S.; Seif, A.; Minev, Z. K. Efficient Long-Range Entanglement using Dynamic Circuits. 2023.
  183. 183
    Shirizly, L.; Misguich, G.; Landa, H. Dissipative Dynamics of Graph-State Stabilizers with Superconducting Qubits. Phys. Rev. Lett. 2024, 132, 010601  DOI: 10.1103/PhysRevLett.132.010601
  184. 184
    Niu, S.; Todri-Sanial, A.; Bronn, N. T. Multi-qubit dynamical decoupling for enhanced crosstalk suppression. Quantum Science and Technology 2024, 9, 045003  DOI: 10.1088/2058-9565/ad5a37
  185. 185
    Evert, B.; Izquierdo, Z. G.; Sud, J.; Hu, H.-Y.; Grabbe, S.; Rieffel, E. G.; Reagor, M. J.; Wang, Z. Syncopated Dynamical Decoupling for Suppressing Crosstalk in Quantum Circuits. 2024; https://arxiv.org/abs/2403.07836 (accessed: 2025-02-23).
  186. 186
    Brown, A. F.; Lidar, D. A. Efficient Chromatic-Number-Based Multi-Qubit Decoherence and Crosstalk Suppression. 2024. https://arxiv.org/abs/2406.13901(accessed: 2025-02-23).
  187. 187
    Martinis, J. M.; Geller, M. R. Fast adiabatic qubit gates using only σz control. Phys. Rev. A 2014, 90, 022307  DOI: 10.1103/PhysRevA.90.022307
  188. 188
    Krantz, P.; Kjaergaard, M.; Yan, F.; Orlando, T. P.; Gustavsson, S.; Oliver, W. D. A quantum engineer’s guide to superconducting qubits. Applied Physics Reviews 2019, 6, 021318  DOI: 10.1063/1.5089550
  189. 189
    Genov, G. T.; Schraft, D.; Vitanov, N. V.; Halfmann, T. Arbitrarily Accurate Pulse Sequences for Robust Dynamical Decoupling. Phys. Rev. Lett. 2017, 118, 133202,  DOI: 10.1103/PhysRevLett.118.133202
  190. 190
    Motzoi, F.; Gambetta, J. M.; Rebentrost, P.; Wilhelm, F. K. Simple Pulses for Elimination of Leakage in Weakly Nonlinear Qubits. Phys. Rev. Lett. 2009, 103, 110501,  DOI: 10.1103/PhysRevLett.103.110501
  191. 191
    Chow, J. M.; Gambetta, J. M.; Tornberg, L.; Koch, J.; Bishop, L. S.; Houck, A. A.; Johnson, B. R.; Frunzio, L.; Girvin, S. M.; Schoelkopf, R. J. Randomized Benchmarking and Process Tomography for Gate Errors in a Solid-State Qubit. Phys. Rev. Lett. 2009, 102, 090502  DOI: 10.1103/PhysRevLett.102.090502
  192. 192
    Albash, T.; Boixo, S.; Lidar, D. A.; Zanardi, P. Quantum adiabatic Markovian master equations. New J. of Phys. 2012, 14, 123016,  DOI: 10.1088/1367-2630/14/12/123016
  193. 193
    Dann, R.; Levy, A.; Kosloff, R. Time-dependent Markovian quantum master equation. Phys. Rev. A 2018, 98, 052129  DOI: 10.1103/PhysRevA.98.052129
  194. 194
    Campos Venuti, L.; Lidar, D. A. Error reduction in quantum annealing using boundary cancellation: Only the end matters. Phys. Rev. A 2018, 98, 022315  DOI: 10.1103/PhysRevA.98.022315
  195. 195
    Mozgunov, E.; Lidar, D. Completely positive master equation for arbitrary driving and small level spacing. Quantum 2020, 4, 227,  DOI: 10.22331/q-2020-02-06-227
  196. 196
    Nathan, F.; Rudner, M. S. Universal Lindblad equation for open quantum systems. Phys. Rev. B 2020, 102, 115109,  DOI: 10.1103/PhysRevB.102.115109
  197. 197
    Davidović, D. Completely Positive, Simple, and Possibly Highly Accurate Approximation of the Redfield Equation. Quantum 2020, 4, 326,  DOI: 10.22331/q-2020-09-21-326
  198. 198
    Davidović, D. Geometric-arithmetic master equation in large and fast open quantum systems. Journal of Physics A: Mathematical and Theoretical 2022, 55, 455301,  DOI: 10.1088/1751-8121/ac9f30
  199. 199
    Kasatkin, V.; Gu, L.; Lidar, D. A. Which differential equations correspond to the Lindblad equation?. Physical Review Research 2023, 5, 043163  DOI: 10.1103/PhysRevResearch.5.043163
  200. 200
    Perko, L. Differential Equations and Dynamical Systems; Springer New York: New York, NY, 2001; pp 163.
  201. 201
    Geerlings, K.; Leghtas, Z.; Pop, I. M.; Shankar, S.; Frunzio, L.; Schoelkopf, R. J.; Mirrahimi, M.; Devoret, M. H. Demonstrating a Driven Reset Protocol for a Superconducting Qubit. Phys. Rev. Lett. 2013, 110, 120501,  DOI: 10.1103/PhysRevLett.110.120501
  202. 202
    Goss, N.; Morvan, A.; Marinelli, B.; Mitchell, B. K.; Nguyen, L. B.; Naik, R. K.; Chen, L.; Jünger, C.; Kreikebaum, J. M.; Santiago, D. I.; Wallman, J. J.; Siddiqi, I. High-fidelity qutrit entangling gates for superconducting circuits. Nat. Commun. 2022, 13, 7481,  DOI: 10.1038/s41467-022-34851-z
  203. 203
    Tripathi, V.; Goss, N.; Vezvaee, A.; Nguyen, L. B.; Siddiqi, I.; Lidar, D. A. Qudit Dynamical Decoupling on a Superconducting Quantum Processor. Phys. Rev. Lett. 2025, 134, 050601  DOI: 10.1103/PhysRevLett.134.050601
  204. 204
    Krishnan Vijayan, M.; Paler, A.; Gavriel, J.; Myers, C. R.; Rohde, P. P.; Devitt, S. J. Compilation of algorithm-specific graph states for quantum circuits. Quantum Science and Technology 2024, 9, 025005  DOI: 10.1088/2058-9565/ad1f39
  205. 205
    Beverland, M. E.; Murali, P.; Troyer, M.; Svore, K. M.; Hoefler, T.; Kliuchnikov, V.; Low, G. H.; Soeken, M.; Sundaram, A.; Vaschillo, A. Assessing requirements to scale to practical quantum advantage. 2022. https://arxiv.org/abs/2211.07629 (accessed 2025-02-23).
  206. 206
    Agrawal, A. A.; Job, J.; Wilson, T. L.; Saadatmand, S. N.; Hodson, M. J.; Mutus, J. Y.; Caesura, A.; Johnson, P. D.; Elenewski, J. E.; Morrell, K. J.; Kemper, A. F. Quantifying fault tolerant simulation of strongly correlated systems using the Fermi-Hubbard model. 2024; https://arxiv.org/abs/2406.06511(accessed: 2025-02-23).
  207. 207
    Barends, R. Coherent Josephson Qubit Suitable for Scalable Quantum Integrated Circuits. Phys. Rev. Lett. 2013, 111, 080502  DOI: 10.1103/PhysRevLett.111.080502
  208. 208
    Shanto, S.; Kuo, A.; Miyamoto, C.; Zhang, H.; Maurya, V.; Vlachos, E.; Hecht, M.; Shum, C. W.; Levenson-Falk, E. SQuADDS: A Validated Design Database and Simulation Workflow for Superconducting Qubit Design. Quantum 2024, 8, 1465,  DOI: 10.22331/q-2024-09-09-1465
  209. 209
    Shah, P. A. . qiskit-community/qiskit-metal: Qiskit Metal 0.1.5a1 (2023). 2023;  DOI: 10.5281/zenodo.8061781 (accessed 2025-02-23).
  210. 210
    Minev, Z. K.; Leghtas, Z.; Reinhold, P.; Mundhada, S. O.; Diringer, A.; Hillel, D. C.; Wang, D. Z.-R.; Facchini, M.; Shah, P. A.; Devoret, M. pyEPR: The energy-participation-ratio (EPR) open- source framework for quantum device design. 2021;  DOI: 10.5281/zenodo.4744448 , https://github.com/zlatko-minev/pyEPR; https://pyepr-docs.readthedocs.io/en/latest/ (accessed 2025-02-23).
  211. 211
    Chen, Z. Measuring and Suppressing Quantum State Leakage in a Superconducting Qubit. Phys. Rev. Lett. 2016, 116, 020501  DOI: 10.1103/PhysRevLett.116.020501

Cited By

Click to copy section linkSection link copied!

This article has not yet been cited by other publications.

Chemical Reviews

Cite this: Chem. Rev. 2025, XXXX, XXX, XXX-XXX
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.chemrev.4c00870
Published May 5, 2025

© 2025 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0 .

Article Views

1476

Altmetric

-

Citations

-
Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. Three different QPT methods. (a) Schematic of SQPT. An ensemble of states {ρj} is prepared and each state is subjected to the map E followed by measurements {Em}. (b) Schematic of separable AAPT. A “faithful” input state ρ is subjected to the map EI. The operators {Ej} are measured on the joint system, which results in the required joint probability distributions or expectation values. (c) Schematic of DCQD. The system and the ancilla are prepared in one of the input states as in Table 1 and are subjected to the map EI. The joint system is measured in the Bell-state basis (Bell State Measurement).

    Figure 2

    Figure 2. Gate set tomography circuit consisting of the fiducial gates Fi (Fj) as well as the measurement operators M and the native gate Gk performed on the initial state ρ. Figure adapted from ref (141). Copyright 2021 Quantum.

    Figure 3

    Figure 3. Schematic of DFE: To estimate the state fidelity, expectations of random Pauli strings Pj are calculated for the input state ρ. The distribution of the random Pauli strings depends on the state with respect to which the fidelity is being calculated.

    Figure 4

    Figure 4. RCS benchmarking circuit: layers of 2-qubit gates (white boxes) are interleaved with Haar random 1-qubit gates (blue boxes), followed by measurement in the computational basis at the end. Adapted from Figure 1 of ref (156). Copyright 2021 arXiv.

    Figure 5

    Figure 5. Schematic representations of the gate sequences for (a) randomized benchmarking (RB) and (b) deterministic benchmarking (DB), respectively. RB uses N random Clifford gates Ci, followed by a single recovery gate Cr = CN···C1. DB prepares an initial state via U|0⟩, applies n repetitions of a fixed, deterministic two-pulse sequence P1P2, and unprepares the initial state via U.

    Figure 6

    Figure 6. Empirical (, symbols), analytical fit (F, solid curves), and Lindblad-based (, dashed curves) fidelities for the initial states |1⟩ under free evolution and |+⟩ under various gate sequences P1P2. The analytical fit yields T1 = 23.36 ± 0.40 μs, T2 = 44.13 ± 2.49 μs, δθ = 0.398 ± 0.004°, and δϕ = 0.426 ± 0.004°. With the exception of the XX (orange) and YY (green) cases, the Lindblad-based results are indistinguishable from the analytical fits.

    Figure 7

    Figure 7. Randomized benchmarking measurement for the same gate calibration parameters used in Figure 6. The error bars represent the deviation of the averaged benchmarking results across 30 distinct random sequences with a maximum number of Cliffords (NCliffords) of 700, which are smaller than the marker sizes.

    Figure 8

    Figure 8. Experimental fidelities |+⟩ showing the very different sensitivity of DB and RB to coherent errors. (a) δθ = 0°, δϕ = 0°, (b) δθ = 0°, δϕ = 0.893 ± 0.002°, (c) δθ = 0.932 ± 0.007°, δϕ = 0°, (d) δθ = 0.995 ± 0.009°, δϕ = 0.90 ± 0.002°. These angles are obtained from fits using eq 76. Dashed curves are the numerically computed fidelities F|+⟩ from the open system Lindblad model using the DB parameters above along with the corresponding T1 and T2. The UR6 sequence suppresses even large coherent errors.

    Figure 9

    Figure 9. Asymmetry in the decay pattern of DB sequences due to the interplay of gates and T1. Fidelity decay for (a) XX and (b) YY applied to two pairs of orthogonal initial states. Dashed curves represent the Lindblad master equation simulation results with parameters obtained by fitting eq 76.

    Figure 10

    Figure 10. Simulated survival fidelity of the |+⟩ state after repetitions of the XX sequence, where each gate is a cosine pulse length of 20 ns (blue circles) or 10 ns (orange squares). A third state in the simulation represents leakage, with a detuning (anharmonicity) of −150 MHz relative to the qubit transition. Both the amplitude and frequency of oscillations increase as the gate duration is decreased and leakage becomes more significant.

  • References


    This article references 211 other publications.

    1. 1
      Bharti, K.; Cervera-Lierta, A.; Kyaw, T. H.; Haug, T.; Alperin-Lea, S.; Anand, A.; Degroote, M.; Heimonen, H.; Kottmann, J. S.; Menke, T.; Mok, W.-K.; Sim, S.; Kwek, L.-C.; Aspuru-Guzik, A. Noisy intermediate-scale quantum algorithms. Rev. Mod. Phys. 2022, 94, 015004  DOI: 10.1103/RevModPhys.94.015004
    2. 2
      Dalzell, A. M.; McArdle, S.; Berta, M.; Bienias, P.; Chen, C.-F.; Gilyén, A.; Hann, C. T.; Kastoryano, M. J.; Khabiboulline, E. T.; Kubica, A.; Salton, G.; Wang, S.; Brandão, F. G. S. L. Quantum algorithms: A survey of applications and end-to-end complexities. 2023; https://arxiv.org/abs/2310.03011 (accessed 2025-02-23).
    3. 3
      Shor, P. Fault-tolerant quantum computation. Proceedings of 37th Conference on Foundations of Computer Science. 1996; pp 5665.
    4. 4
      Aharonov, D.; Ben-Or, M. Fault tolerant quantum computation with constant error. Proceedings of 29th Annual ACM Symposium on Theory of Computing (STOC). New York, NY, 1997; p 176.
    5. 5
      Knill, E.; Laflamme, R.; Zurek, W. H. Resilient Quantum Computation. Science 1998, 279, 342345,  DOI: 10.1126/science.279.5349.342
    6. 6
      Aliferis, P.; Gottesman, D.; Preskill, J. Quantum accuracy threshold for concatenated distance-3 codes. Quant. Inf. Comput. 2006, 6, 97,  DOI: 10.26421/QIC6.2-1
    7. 7
      Campbell, E. T.; Terhal, B. M.; Vuillot, C. Roads towards fault-tolerant universal quantum computation. Nature 2017, 549, 172,  DOI: 10.1038/nature23460
    8. 8
      Reiher, M.; Wiebe, N.; Svore, K. M.; Wecker, D.; Troyer, M. Elucidating reaction mechanisms on quantum computers. Proc. Natl. Acad. Sci. U. S. A. 2017, 114, 75557560,  DOI: 10.1073/pnas.1619152114
    9. 9
      Kassal, I.; Whitfield, J. D.; Perdomo-Ortiz, A.; Yung, M.-H.; Aspuru-Guzik, A. Simulating Chemistry Using Quantum Computers. Annu. Rev. Phys. Chem. 2011, 62, 185207,  DOI: 10.1146/annurev-physchem-032210-103512
    10. 10
      Preskill, J. Quantum Computing in the NISQ era and beyond. Quantum 2018, 2, 79,  DOI: 10.22331/q-2018-08-06-79
    11. 11
      O’Malley, P. J. J. Scalable Quantum Simulation of Molecular Energies. Physical Review X 2016, 6, 031007  DOI: 10.1103/PhysRevX.6.031007
    12. 12
      Kandala, A.; Mezzacapo, A.; Temme, K.; Takita, M.; Brink, M.; Chow, J. M.; Gambetta, J. M. Hardware-efficient variational quantum eigensolver for small molecules and quantum magnets. Nature 2017, 549, 242,  DOI: 10.1038/nature23879
    13. 13
      Lidar, D. A.; Wang, H. Calculating the thermal rate constant with exponential speedup on a quantum computer. Phys. Rev. E 1999, 59, 24292438,  DOI: 10.1103/PhysRevE.59.2429
    14. 14
      Aspuru-Guzik, A.; Dutoi, A. D.; Love, P. J.; Head-Gordon, M. Simulated Quantum Computation of Molecular Energies. Science 2005, 309, 1704,  DOI: 10.1126/science.1113479
    15. 15
      Peruzzo, A.; McClean, J.; Shadbolt, P.; Yung, M.-H.; Zhou, X.-Q.; Love, P. J.; Aspuru-Guzik, A.; O’Brien, J. L. A variational eigenvalue solver on a photonic quantum processor. Nat. Commun. 2014, 5, 4213,  DOI: 10.1038/ncomms5213
    16. 16
      McClean, J. R.; Romero, J.; Babbush, R.; Aspuru-Guzik, A. The theory of variational hybrid quantum-classical algorithms. New J. Phys. 2016, 18, 023023  DOI: 10.1088/1367-2630/18/2/023023
    17. 17
      McArdle, S.; Endo, S.; Aspuru-Guzik, A.; Benjamin, S. C.; Yuan, X. Quantum computational chemistry. Rev. Mod. Phys. 2020, 92, 015003  DOI: 10.1103/RevModPhys.92.015003
    18. 18
      Tilly, J.; Chen, H.; Cao, S.; Picozzi, D.; Setia, K.; Li, Y.; Grant, E.; Wossnig, L.; Rungger, I.; Booth, G. H.; Tennyson, J. The Variational Quantum Eigensolver: A review of methods and best practices. Phys. Rep. 2022, 986, 1128,  DOI: 10.1016/j.physrep.2022.08.003
    19. 19
      Kitaev, A. Y. Quantum measurements and the Abelian Stabilizer Problem. 1995; https://arxiv.org/abs/quant-ph/9511026 (accessed 2025-02-23).
    20. 20
      Cao, Y.; Romero, J.; Olson, J. P.; Degroote, M.; Johnson, P. D.; Kieferová, M.; Kivlichan, I. D.; Menke, T.; Peropadre, B.; Sawaya, N. P. D.; Sim, S.; Veis, L.; Aspuru-Guzik, A. Quantum Chemistry in the Age of Quantum Computing. Chem. Rev. 2019, 119, 1085610915,  DOI: 10.1021/acs.chemrev.8b00803
    21. 21
      Lee, S. Evaluating the evidence for exponential quantum advantage in ground-state quantum chemistry. Nat. Commun. 2023, 14, 1952,  DOI: 10.1038/s41467-023-37587-6
    22. 22
      Huggins, W. J.; McClean, J. R.; Rubin, N. C.; Jiang, Z.; Wiebe, N.; Whaley, K. B.; Babbush, R. Efficient and noise resilient measurements for quantum chemistry on near-term quantum computers. npj Quantum Information 2021, 7, 23,  DOI: 10.1038/s41534-020-00341-7
    23. 23
      Hashim, A. . A Practical Introduction to Benchmarking and Characterization of Quantum Computers. 2024; https://arxiv.org/abs/2408.12064 (accessed 2025-02-23).
    24. 24
      Chuang, I. L.; Nielsen, M. A. Prescription for experimental determination of the dynamics of a quantum black box. J. Mod. Opt. 1997, 44, 2455,  DOI: 10.1080/09500349708231894
    25. 25
      Poyatos, J. F.; Cirac, J. I.; Zoller, P. Complete Characterization of a Quantum Process: The Two-Bit Quantum Gate. Phys. Rev. Lett. 1997, 78, 390393,  DOI: 10.1103/PhysRevLett.78.390
    26. 26
      Mohseni, M.; Rezakhani, A. T.; Lidar, D. A. Quantum-process tomography: Resource analysis of different strategies. Phys. Rev. A 2008, 77, 032322  DOI: 10.1103/PhysRevA.77.032322
    27. 27
      Emerson, J.; Alicki, R.; Życzkowski, K. Scalable noise estimation with random unitary operators. Journal of Optics B: Quantum and Semiclassical Optics 2005, 7, S347,  DOI: 10.1088/1464-4266/7/10/021
    28. 28
      Knill, E.; Leibfried, D.; Reichle, R.; Britton, J.; Blakestad, R. B.; Jost, J. D.; Langer, C.; Ozeri, R.; Seidelin, S.; Wineland, D. J. Randomized benchmarking of quantum gates. Phys. Rev. A 2008, 77, 012307  DOI: 10.1103/PhysRevA.77.012307
    29. 29
      Magesan, E.; Gambetta, J. M.; Emerson, J. Scalable and Robust Randomized Benchmarking of Quantum Processes. Phys. Rev. Lett. 2011, 106, 180504,  DOI: 10.1103/PhysRevLett.106.180504
    30. 30
      Helsen, J.; Roth, I.; Onorati, E.; Werner, A. H.; Eisert, J. General Framework for Randomized Benchmarking. PRX Quantum 2022, 3, 020357  DOI: 10.1103/PRXQuantum.3.020357
    31. 31
      Merkel, S. T.; Gambetta, J. M.; Smolin, J. A.; Poletto, S.; Córcoles, A. D.; Johnson, B. R.; Ryan, C. A.; Steffen, M. Self-consistent quantum process tomography. Phys. Rev. A 2013, 87, 062119  DOI: 10.1103/PhysRevA.87.062119
    32. 32
      Blume-Kohout, R.; Gamble, J. K.; Nielsen, E.; Mizrahi, J.; Sterk, J. D.; Maunz, P. Robust, self-consistent, closed-form tomography of quantum logic gates on a trapped ion qubit. 2013; https://arxiv.org/abs/1310.4492 (accessed 2025-02-23).
    33. 33
      Kim, D.; Ward, D. R.; Simmons, C. B.; Gamble, J. K.; Blume-Kohout, R.; Nielsen, E.; Savage, D. E.; Lagally, M. G.; Friesen, M.; Coppersmith, S. N.; Eriksson, M. A. Microwave-driven coherent operation of a semiconductor quantum dot charge qubit. Nat. Nanotechnol. 2015, 10, 243247,  DOI: 10.1038/nnano.2014.336
    34. 34
      Bäumer, E.; Tripathi, V.; Seif, A.; Lidar, D.; Wang, D. S. Quantum Fourier Transform Using Dynamic Circuits. Phys. Rev. Lett. 2024, 133, 150602150609,  DOI: 10.1103/PhysRevLett.133.150602
    35. 35
      da Silva, M. P.; Landon-Cardinal, O.; Poulin, D. Practical characterization of quantum devices without tomography. Phys. Rev. Lett. 2011, 107, 210404210409,  DOI: 10.1103/PhysRevLett.107.210404
    36. 36
      Flammia, S. T.; Liu, Y.-K. Direct Fidelity Estimation from Few Pauli Measurements. Phys. Rev. Lett. 2011, 106, 230501,  DOI: 10.1103/PhysRevLett.106.230501
    37. 37
      Neill, C. A blueprint for demonstrating quantum supremacy with superconducting qubits. Science 2018, 360, 195199,  DOI: 10.1126/science.aao4309
    38. 38
      Boixo, S.; Isakov, S. V.; Smelyanskiy, V. N.; Babbush, R.; Ding, N.; Jiang, Z.; Bremner, M. J.; Martinis, J. M.; Neven, H. Characterizing quantum supremacy in near-term devices. Nature Physics 2018, 14, 595600
    39. 39
      Arute, F. Quantum supremacy using a programmable superconducting processor. Nature 2019, 574, 505510
    40. 40
      Kueng, R.; Long, D. M.; Doherty, A. C.; Flammia, S. T. Comparing Experiments to the Fault-Tolerance Threshold. Phys. Rev. Lett. 2016, 117, 170502,  DOI: 10.1103/PhysRevLett.117.170502
    41. 41
      Barnes, J. P.; Trout, C. J.; Lucarelli, D.; Clader, B. D. Quantum error-correction failure distributions: Comparison of coherent and stochastic error models. Phys. Rev. A 2017, 95, 062338  DOI: 10.1103/PhysRevA.95.062338
    42. 42
      Ng, H. K.; Preskill, J. Fault-tolerant quantum computation versus Gaussian noise. Phys. Rev. A 2009, 79, 032318. DOI: 10.1103/PhysRevA.79.032318
    43. 43
      Wallman, J.; Granade, C.; Harper, R.; Flammia, S. T. Estimating the coherence of noise. New J. Phys. 2015, 17, 113020,  DOI: 10.1088/1367-2630/17/11/113020
    44. 44
      Sanders, Y. R.; Wallman, J. J.; Sanders, B. C. Bounding quantum gate error rate based on reported average fidelity. New J. Phys. 2016, 18, 012002  DOI: 10.1088/1367-2630/18/1/012002
    45. 45
      Hashim, A.; Seritan, S.; Proctor, T.; Rudinger, K.; Goss, N.; Naik, R. K.; Kreikebaum, J. M.; Santiago, D. I.; Siddiqi, I. Benchmarking quantum logic operations relative to thresholds for fault tolerance. npj Quantum Information 2023, 9, 109,  DOI: 10.1038/s41534-023-00764-y
    46. 46
      Wallman, J. J.; Emerson, J. Noise tailoring for scalable quantum computation via randomized compiling. Phys. Rev. A 2016, 94, 052325052334,  DOI: 10.1103/PhysRevA.94.052325
    47. 47
      Hashim, A.; Naik, R. K.; Morvan, A.; Ville, J.-L.; Mitchell, B.; Kreikebaum, J. M.; Davis, M.; Smith, E.; Iancu, C.; O’Brien, K. P.; Hincks, I.; Wallman, J. J.; Emerson, J.; Siddiqi, I. Randomized Compiling for Scalable Quantum Computing on a Noisy Superconducting Quantum Processor. Phys. Rev. X 2021, 11, 041039  DOI: 10.1103/PhysRevX.11.041039
    48. 48
      van den Berg, E.; Minev, Z. K.; Kandala, A.; Temme, K. Probabilistic error cancellation with sparse Pauli–Lindblad models on noisy quantum processors. Nat. Phys. 2023, 19, 11161121,  DOI: 10.1038/s41567-023-02042-2
    49. 49
      Seif, A.; Liao, H.; Tripathi, V.; Krsulich, K.; Malekakhlagh, M.; Amico, M.; Jurcevic, P.; Javadi-Abhari, A. Suppressing Correlated Noise in Quantum Computers via Context-Aware Compiling. 51st Annual International Symposium on Computer Architecture (ISCA) 2024, 310324
    50. 50
      Gross, J. A.; Genois, E.; Debroy, D. M.; Zhang, Y.; Mruczkiewicz, W.; Cian, Z.-P.; Jiang, Z. Characterizing Coherent Errors using Matrix-Element Amplification. npj Quantum Information 2024, 10, 123,  DOI: 10.1038/s41534-024-00917-7
    51. 51
      Bravyi, S.; Englbrecht, M.; König, R.; Peard, N. Correcting coherent errors with surface codes. npj Quantum Information 2018, 4, 55,  DOI: 10.1038/s41534-018-0106-y
    52. 52
      Barenco, A.; Bennett, C. H.; Cleve, R.; DiVincenzo, D. P.; Margolus, N.; Shor, P.; Sleator, T.; Smolin, J. A.; Weinfurter, H. Elementary gates for quantum computation. Phys. Rev. A 1995, 52, 34573467,  DOI: 10.1103/PhysRevA.52.3457
    53. 53
      Erhard, A.; Wallman, J. J.; Postler, L.; Meth, M.; Stricker, R.; Martinez, E. A.; Schindler, P.; Monz, T.; Emerson, J.; Blatt, R. Characterizing large-scale quantum computers via cycle benchmarking. Nat. Commun. 2019, 10, 5347,  DOI: 10.1038/s41467-019-13068-7
    54. 54
      McKay, D. C.; Wood, C. J.; Sheldon, S.; Chow, J. M.; Gambetta, J. M. Efficient Z gates for quantum computing. Phys. Rev. A 2017, 96, 022330  DOI: 10.1103/PhysRevA.96.022330
    55. 55
      Cross, A. W.; Bishop, L. S.; Sheldon, S.; Nation, P. D.; Gambetta, J. M. Validating quantum computers using randomized model circuits. Phys. Rev. A 2019, 100, 032328  DOI: 10.1103/PhysRevA.100.032328
    56. 56
      Dankert, C.; Cleve, R.; Emerson, J.; Livine, E. Exact and approximate unitary 2-designs and their application to fidelity estimation. Phys. Rev. A 2009, 80, 012304  DOI: 10.1103/PhysRevA.80.012304
    57. 57
      Proctor, T.; Rudinger, K.; Young, K.; Nielsen, E.; Blume-Kohout, R. Measuring the capabilities of quantum computers. Nat. Phys. 2022, 18, 7579,  DOI: 10.1038/s41567-021-01409-7
    58. 58
      Cincio, L.; Subası, Y.; Sornborger, A. T; Coles, P. J Learning the quantum algorithm for state overlap. New J. Phys. 2018, 20, 113022,  DOI: 10.1088/1367-2630/aae94a
    59. 59
      Reich, D. M.; Gualdi, G.; Koch, C. P. Optimal Strategies for Estimating the Average Fidelity of Quantum Gates. Phys. Rev. Lett. 2013, 111, 200401200406,  DOI: 10.1103/PhysRevLett.111.200401
    60. 60
      Torlai, G.; Mazzola, G.; Carrasquilla, J.; Troyer, M.; Melko, R.; Carleo, G. Neural-network quantum state tomography. Nat. Phys. 2018, 14, 447450,  DOI: 10.1038/s41567-018-0048-5
    61. 61
      Koutny, D.; Motka, L.; Hradil, Z.; Rehacek, J.; Sanchez-Soto, L. L. Neural-network quantum state tomography. Phys. Rev. A 2022, 106. DOI: 10.1103/PhysRevA.106.012409
    62. 62
      Gu, Y.; Zhuang, W.-F.; Chai, X.; Liu, D. E. Benchmarking universal quantum gates via channel spectrum. Nat. Commun. 2023, 14, 5880,  DOI: 10.1038/s41467-023-41598-8
    63. 63
      McKay, D. C.; Hincks, I.; Pritchett, E. J.; Carroll, M.; Govia, L. C. G.; Merkel, S. T. Benchmarking Quantum Processor Performance at Scale. 2023; https://arxiv.org/abs/2311.05933 (accessed: 2025-02-23).
    64. 64
      Breuer, H.-P.; Petruccione, F. The Theory of Open Quantum Systems; Oxford University Press: Oxford, 2002.
    65. 65
      Kraus, K. States, Effects and Operations; Fundamental Notions of Quantum Theory; Academic, 1983.
    66. 66
      Nielsen, M. A.; Chuang, I. L. Quantum computation and quantum information; Cambridge University Press, 2010.
    67. 67
      Jamiołkowski, A. Linear transformations which preserve trace and positive semidefiniteness of operators. Reports on Mathematical Physics 1972, 3, 275278,  DOI: 10.1016/0034-4877(72)90011-0
    68. 68
      Choi, M.-D. Completely positive linear maps on complex matrices. Linear Algebra and its Applications 1975, 10, 285290,  DOI: 10.1016/0024-3795(75)90075-0
    69. 69
      Uhlmann, A. The “transition probability”in the state space of a ∗-algebra. Reports on Mathematical Physics 1976, 9, 273279,  DOI: 10.1016/0034-4877(76)90060-4
    70. 70
      Knill, E.; Laflamme, R. Theory of quantum error-correcting codes. Phys. Rev. A 1997, 55, 900911,  DOI: 10.1103/PhysRevA.55.900
    71. 71
      Tripathi, V.; Chen, H.; Levenson-Falk, E.; Lidar, D. A. Modeling Low- and High-Frequency Noise in Transmon Qubits with Resource-Efficient Measurement. PRX Quantum 2024, 5, 010320  DOI: 10.1103/PRXQuantum.5.010320
    72. 72
      Byrd, M. S.; Lidar, D. A.; Wu, L.-A.; Zanardi, P. Universal leakage elimination. Phys. Rev. A 2005, 71, 052301  DOI: 10.1103/PhysRevA.71.052301
    73. 73
      Koch, J.; Yu, T. M.; Gambetta, J.; Houck, A. A.; Schuster, D. I.; Majer, J.; Blais, A.; Devoret, M. H.; Girvin, S. M.; Schoelkopf, R. J. Charge-insensitive qubit design derived from the Cooper pair box. Phys. Rev. A 2007, 76, 042319  DOI: 10.1103/PhysRevA.76.042319
    74. 74
      Wu, L.-A.; Walther, P.; Lidar, D. A. No-go theorem for passive single-rail linear optical quantum computing. Sci. Rep. 2013, 3, 1394,  DOI: 10.1038/srep01394
    75. 75
      Levine, H. Demonstrating a Long-Coherence Dual-Rail Erasure Qubit Using Tunable Transmons. Physical Review X 2024, 14, 011051011072,  DOI: 10.1103/PhysRevX.14.011051
    76. 76
      Grünzweig, T.; Hilliard, A.; McGovern, M.; Andersen, M. F. Near-deterministic preparation of a single atom in an optical microtrap. Nat. Phys. 2010, 6, 951954,  DOI: 10.1038/nphys1778
    77. 77
      Wu, Y.; Kolkowitz, S.; Puri, S.; Thompson, J. D. Erasure conversion for fault-tolerant quantum computing in alkaline earth Rydberg atom arrays. Nat. Commun. 2022, 13, 4657,  DOI: 10.1038/s41467-022-32094-6
    78. 78
      Varbanov, B. M.; Battistel, F.; Tarasinski, B. M.; Ostroukh, V. P.; O’Brien, T. E.; DiCarlo, L.; Terhal, B. M. Leakage detection for a transmon-based surface code. npj Quantum Information 2020, 6, 102,  DOI: 10.1038/s41534-020-00330-w
    79. 79
      McEwen, M. Removing leakage-induced correlated errors in superconducting quantum error correction. Nat. Commun. 2021, 12, 1761,  DOI: 10.1038/s41467-021-21982-y
    80. 80
      Gourlay, I.; Snowdon, J. F. Concatenated Coding in the Presence of Dephasing. Phys. Rev. A 2000, 62, 022308  DOI: 10.1103/PhysRevA.62.022308
    81. 81
      Evans, Z. W. E.; Stephens, A. M.; Cole, J. H.; Hollenberg, L. C. L. Error Correction Optimisation in the Presence of X/Z Asymmetry. 2007.
    82. 82
      Stephens, A. M.; Evans, Z. W. E.; Devitt, S. J.; Hollenberg, L. C. L. Asymmetric Quantum Error Correction via Code Conversion. Phys. Rev. A 2008, 77, 062335  DOI: 10.1103/PhysRevA.77.062335
    83. 83
      Aliferis, P.; Preskill, J. Fault-tolerant quantum computation against biased noise. Phys. Rev. A 2008, 78, 052331  DOI: 10.1103/PhysRevA.78.052331
    84. 84
      Stephens, A. M.; Munro, W. J.; Nemoto, K. High-Threshold Topological Quantum Error Correction against Biased Noise. Phys. Rev. A 2013, 88, 060301  DOI: 10.1103/PhysRevA.88.060301
    85. 85
      Brooks, P.; Preskill, J. Fault-Tolerant Quantum Computation with Asymmetric Bacon-Shor Codes. Phys. Rev. A 2013, 87, 032310  DOI: 10.1103/PhysRevA.87.032310
    86. 86
      Bonilla Ataides, J. P.; Tuckett, D. K.; Bartlett, S. D.; Flammia, S. T.; Brown, B. J. The XZZX surface code. Nat. Commun. 2021, 12, 2172,  DOI: 10.1038/s41467-021-22274-1
    87. 87
      Chruscinski, D.; Kossakowski, A. Non-Markovian Quantum Dynamics: Local versus Nonlocal. Phys. Rev. Lett. 2010, 104, 070406  DOI: 10.1103/PhysRevLett.104.070406
    88. 88
      Rivas, Á.; Huelga, S. F.; Plenio, M. B. Quantum non-Markovianity: characterization, quantification and detection. Rep. Prog. Phys. 2014, 77, 094001  DOI: 10.1088/0034-4885/77/9/094001
    89. 89
      Shabani, A.; Lidar, D. A. Completely positive post-Markovian master equation via a measurement approach. Phys. Rev. A 2005, 71, 020101  DOI: 10.1103/PhysRevA.71.020101
    90. 90
      Zhang, H.; Pokharel, B.; Levenson-Falk, E.; Lidar, D. Predicting Non-Markovian Superconducting-Qubit Dynamics from Tomographic Reconstruction. Physical Review Applied 2022, 17, 054018  DOI: 10.1103/PhysRevApplied.17.054018
    91. 91
      Zhang, J.; Vala, J.; Sastry, S.; Whaley, K. B. Geometric theory of nonlocal two-qubit operations. Phys. Rev. A 2003, 67, 042313  DOI: 10.1103/PhysRevA.67.042313
    92. 92
      Sete, E. A.; Didier, N.; Chen, A. Q.; Kulshreshtha, S.; Manenti, R.; Poletto, S. Parametric-Resonance Entangling Gates with a Tunable Coupler. Phys. Rev. Appl. 2021, 16, 024050  DOI: 10.1103/PhysRevApplied.16.024050
    93. 93
      Sete, E. A.; Tripathi, V.; Valery, J. A.; Lidar, D.; Mutus, J. Y. Error Budget of a Parametric Resonance Entangling Gate with a Tunable Coupler. Physical Review Applied 2024, 22, 014059  DOI: 10.1103/PhysRevApplied.22.014059
    94. 94
      Tripathi, V.; Khezri, M.; Korotkov, A. N. Operation and intrinsic error budget of a two-qubit cross-resonance gate. Phys. Rev. A 2019, 100, 012301  DOI: 10.1103/PhysRevA.100.012301
    95. 95
      Gaikwad, C.; Kowsari, D.; Brame, C.; Song, X.; Zhang, H.; Esposito, M.; Ranadive, A.; Cappelli, G.; Roch, N.; Levenson-Falk, E. M.; Murch, K. W. Entanglement Assisted Probe of the Non-Markovian to Markovian Transition in Open Quantum System Dynamics. Phys. Rev. Lett. 2024, 132, 200401,  DOI: 10.1103/PhysRevLett.132.200401
    96. 96
      Dubi, Y.; Di Ventra, M. Relaxation times in an open interacting two-qubit system. Phys. Rev. A 2009, 79, 012328  DOI: 10.1103/PhysRevA.79.012328
    97. 97
      Greenbaum, D. Introduction to Quantum Gate Set Tomography. 2015, arXiv:1509.02921 [quant-ph] (accessed: 2025-02-23).
    98. 98
      Brown, W. G.; Eastin, B. Randomized Benchmarking with Restricted Gate Sets. Phys. Rev. A 2018, 97, 062323  DOI: 10.1103/PhysRevA.97.062323
    99. 99
      Seif, A.; Chen, S.; Majumder, S.; Liao, H.; Wang, D. S.; Malekakhlagh, M.; Javadi-Abhari, A.; Jiang, L.; Minev, Z. K. Entanglement-enhanced learning of quantum processes at scale. 2024; https://arxiv.org/abs/2408.03376 (accessed: 2025-02-23).
    100. 100
      Khan, M. Q.; Dong, W.; Norris, L. M.; Viola, L. Multiaxis quantum noise spectroscopy robust to errors in state preparation and measurement. Physical Review Applied 2024, 22, 024074  DOI: 10.1103/PhysRevApplied.22.024074
    101. 101
      Merkel, S. T.; Gambetta, J. M.; Smolin, J. A.; Poletto, S.; Córcoles, A. D.; Johnson, B. R.; Ryan, C. A.; Steffen, M. Self-consistent quantum process tomography. Phys. Rev. A 2013, 87, 062119  DOI: 10.1103/PhysRevA.87.062119
    102. 102
      Magesan, E.; Gambetta, J. M.; Emerson, J. Characterizing quantum gates via randomized benchmarking. Phys. Rev. A 2012, 85, 042311  DOI: 10.1103/PhysRevA.85.042311
    103. 103
      Aaronson, S.; Gottesman, D. Improved simulation of stabilizer circuits. Phys. Rev. A 2004, 70, 052328  DOI: 10.1103/PhysRevA.70.052328
    104. 104
      Nielsen, M. A. A simple formula for the average gate fidelity of a quantum dynamical operation. Phys. Lett. A 2002, 303, 249252,  DOI: 10.1016/S0375-9601(02)01272-0
    105. 105
      Viola, L.; Knill, E. Random Decoupling Schemes for Quantum Dynamical Control and Error Suppression. Phys. Rev. Lett. 2005, 94, 060502  DOI: 10.1103/PhysRevLett.94.060502
    106. 106
      Santos, L. F.; Viola, L. Dynamical control of qubit coherence: Random versus deterministic schemes. Phys. Rev. A 2005, 72, 062303  DOI: 10.1103/PhysRevA.72.062303
    107. 107
      Santos, L. F.; Viola, L. Enhanced Convergence and Robust Performance of Randomized Dynamical Decoupling. Phys. Rev. Lett. 2006, 97, 150501,  DOI: 10.1103/PhysRevLett.97.150501
    108. 108
      Magesan, E.; Gambetta, J. M.; Johnson, B. R.; Ryan, C. A.; Chow, J. M.; Merkel, S. T.; da Silva, M. P.; Keefe, G. A.; Rothwell, M. B.; Ohki, T. A.; Ketchen, M. B.; Steffen, M. Efficient Measurement of Quantum Gate Error by Interleaved Randomized Benchmarking. Phys. Rev. Lett. 2012, 109, 080505  DOI: 10.1103/PhysRevLett.109.080505
    109. 109
      Gambetta, J. M.; Córcoles, A. D.; Merkel, S. T.; Johnson, B. R.; Smolin, J. A.; Chow, J. M.; Ryan, C. A.; Rigetti, C.; Poletto, S.; Ohki, T. A.; Ketchen, M. B.; Steffen, M. Characterization of Addressability by Simultaneous Randomized Benchmarking. Phys. Rev. Lett. 2012, 109, 240504,  DOI: 10.1103/PhysRevLett.109.240504
    110. 110
      Wallman, J. J.; Barnhill, M.; Emerson, J. Robust characterization of leakage errors. New J. Phys. 2016, 18, 043021  DOI: 10.1088/1367-2630/18/4/043021
    111. 111
      Wood, C. J.; Gambetta, J. M. Quantification and characterization of leakage errors. Phys. Rev. A 2018, 97, 032306  DOI: 10.1103/PhysRevA.97.032306
    112. 112
      Fogarty, M. A.; Veldhorst, M.; Harper, R.; Yang, C. H.; Bartlett, S. D.; Flammia, S. T.; Dzurak, A. S. Nonexponential fidelity decay in randomized benchmarking with low-frequency noise. Phys. Rev. A 2015, 92, 022326  DOI: 10.1103/PhysRevA.92.022326
    113. 113
      Wallman, J. J. Randomized benchmarking with gate-dependent noise. Quantum 2018, 2, 47,  DOI: 10.22331/q-2018-01-29-47
    114. 114
      Proctor, T.; Seritan, S.; Rudinger, K.; Nielsen, E.; Blume-Kohout, R.; Young, K. Scalable Randomized Benchmarking of Quantum Computers Using Mirror Circuits. Phys. Rev. Lett. 2022, 129, 150502,  DOI: 10.1103/PhysRevLett.129.150502
    115. 115
      Helsen, J.; Xue, X.; Vandersypen, L. M. K.; Wehner, S. A new class of efficient randomized benchmarking protocols. npj Quantum Information 2019, 5, 71,  DOI: 10.1038/s41534-019-0182-7
    116. 116
      Helsen, J.; Roth, I.; Onorati, E.; Werner, A.; Eisert, J. General Framework for Randomized Benchmarking. PRX Quantum 2022, 3, 020357  DOI: 10.1103/PRXQuantum.3.020357
    117. 117
      Arrighi, P.; Patricot, C. On quantum operations as quantum states. Annals of Physics 2004, 311, 2652,  DOI: 10.1016/j.aop.2003.11.005
    118. 118
      D’Ariano, G. M.; Lo Presti, P. Quantum Tomography for Measuring Experimentally the Matrix Elements of an Arbitrary Quantum Operation. Phys. Rev. Lett. 2001, 86, 41954198,  DOI: 10.1103/PhysRevLett.86.4195
    119. 119
      Leung, D. W. Choi’s proof as a recipe for quantum process tomography. Journal of Mathematical Physics 2003, 44, 528,  DOI: 10.1063/1.1518554
    120. 120
      Altepeter, J. B.; Branning, D.; Jeffrey, E.; Wei, T. C.; Kwiat, P. G.; Thew, R. T.; O’Brien, J. L.; Nielsen, M. A.; White, A. G. Ancilla-Assisted Quantum Process Tomography. Phys. Rev. Lett. 2003, 90, 193601,  DOI: 10.1103/PhysRevLett.90.193601
    121. 121
      D’Ariano, G. M.; Lo Presti, P. Imprinting Complete Information about a Quantum Channel on its Output State. Phys. Rev. Lett. 2003, 91, 047902  DOI: 10.1103/PhysRevLett.91.047902
    122. 122
      Mohseni, M.; Lidar, D. A. Direct Characterization of Quantum Dynamics. Phys. Rev. Lett. 2006, 97, 170501,  DOI: 10.1103/PhysRevLett.97.170501
    123. 123
      Mohseni, M.; Lidar, D. A. Direct characterization of quantum dynamics: General theory. Phys. Rev. A 2007, 75, 062331  DOI: 10.1103/PhysRevA.75.062331
    124. 124
      Mohseni, M.; Rezakhani, A. T.; Aspuru-Guzik, A. Direct estimation of single- and two-qubit Hamiltonians and relaxation rates. Phys. Rev. A 2008, 77, 042320. DOI: 10.1103/PhysRevA.77.042320
    125. 125
      Graham, T. M.; Barreiro, J. T.; Mohseni, M.; Kwiat, P. G. Hyperentanglement-Enabled Direct Characterization of Quantum Dynamics. Phys. Rev. Lett. 2013, 110, 060404  DOI: 10.1103/PhysRevLett.110.060404
    126. 126
      Nigg, D.; Barreiro, J. T.; Schindler, P.; Mohseni, M.; Monz, T.; Chwalla, M.; Hennrich, M.; Blatt, R. Experimental Characterization of Quantum Dynamics Through Many-Body Interactions. Phys. Rev. Lett. 2013, 110, 060403  DOI: 10.1103/PhysRevLett.110.060403
    127. 127
      Gottesman, D. Class of quantum error-correcting codes saturating the quantum Hamming bound. Phys. Rev. A 1996, 54, 1862,  DOI: 10.1103/PhysRevA.54.1862
    128. 128
      Blume-Kohout, R.; Gamble, J. K.; Nielsen, E.; Rudinger, K.; Mizrahi, J.; Fortier, K.; Maunz, P. Demonstration of qubit operations below a rigorous fault tolerance threshold with gate set tomography. Nat. Commun. 2017, 8, 14485,  DOI: 10.1038/ncomms14485
    129. 129
      Rol, M. A.; Bultink, C. C.; O’Brien, T. E.; de Jong, S. R.; Theis, L. S.; Fu, X.; Luthi, F.; Vermeulen, R. F. L.; de Sterke, J. C.; Bruno, A.; Deurloo, D.; Schouten, R. N.; Wilhelm, F. K.; DiCarlo, L. Restless Tuneup of High-Fidelity Qubit Gates. Phys. Rev. Appl. 2017, 7, 041001  DOI: 10.1103/PhysRevApplied.7.041001
    130. 130
      Mavadia, S.; Edmunds, C. L.; Hempel, C.; Ball, H.; Roy, F.; Stace, T. M.; Biercuk, M. J. Experimental quantum verification in the presence of temporally correlated noise. npj Quantum Information 2018, 4, 7,  DOI: 10.1038/s41534-017-0052-0
    131. 131
      Hempel, C.; Maier, C.; Romero, J.; McClean, J.; Monz, T.; Shen, H.; Jurcevic, P.; Lanyon, B. P.; Love, P.; Babbush, R.; Aspuru-Guzik, A.; Blatt, R.; Roos, C. F. Quantum Chemistry Calculations on a Trapped-Ion Quantum Simulator. Phys. Rev. X 2018, 8, 031022  DOI: 10.1103/PhysRevX.8.031022
    132. 132
      Song, C.; Cui, J.; Wang, H.; Hao, J.; Feng, H.; Li, Y. Quantum computation with universal error mitigation on a superconducting quantum processor. Science Advances 2019, 5, eaaw5686  DOI: 10.1126/sciadv.aaw5686
    133. 133
      Cincio, L.; Rudinger, K.; Sarovar, M.; Coles, P. J. Machine Learning of Noise-Resilient Quantum Circuits. PRX Quantum 2021, 2, 010324  DOI: 10.1103/PRXQuantum.2.010324
    134. 134
      Hong, S. S.; Papageorge, A. T.; Sivarajah, P.; Crossman, G.; Didier, N.; Polloreno, A. M.; Sete, E. A.; Turkowski, S. W.; da Silva, M. P.; Johnson, B. R. Demonstration of a parametrically activated entangling gate protected from flux noise. Phys. Rev. A 2020, 101, 012302  DOI: 10.1103/PhysRevA.101.012302
    135. 135
      Joshi, M. K.; Elben, A.; Vermersch, B.; Brydges, T.; Maier, C.; Zoller, P.; Blatt, R.; Roos, C. F. Quantum Information Scrambling in a Trapped-Ion Quantum Simulator with Tunable Range Interactions. Phys. Rev. Lett. 2020, 124, 240505,  DOI: 10.1103/PhysRevLett.124.240505
    136. 136
      Zhang, S.; Lu, Y.; Zhang, K.; Chen, W.; Li, Y.; Zhang, J.-N.; Kim, K. Error-mitigated quantum gates exceeding physical fidelities in a trapped-ion system. Nat. Commun. 2020, 11, 587,  DOI: 10.1038/s41467-020-14376-z
    137. 137
      Ware, M.; Ribeill, G.; Ristè, D.; Ryan, C. A.; Johnson, B.; da Silva, M. P. Experimental Pauli-frame randomization on a superconducting qubit. Phys. Rev. A 2021, 103, 042604  DOI: 10.1103/PhysRevA.103.042604
    138. 138
      Dehollain, J. P.; Muhonen, J. T.; Blume-Kohout, R.; Rudinger, K. M.; Gamble, J. K.; Nielsen, E.; Laucht, A.; Simmons, S.; Kalra, R.; Dzurak, A. S.; Morello, A. Optimization of a solid-state electron spin qubit using gate set tomography. New J. Phys. 2016, 18, 103018,  DOI: 10.1088/1367-2630/18/10/103018
    139. 139
      Kitaev, A.; Shen, A.; Vyalyi, M. Classical and Quantum Computation; Graduate studies in mathematics; American Mathematical Society, 2002.
    140. 140
      Scott, A. J. Tight informationally complete quantum measurements. Journal of Physics A: Mathematical and General 2006, 39, 13507,  DOI: 10.1088/0305-4470/39/43/009
    141. 141
      Nielsen, E.; Gamble, J. K.; Rudinger, K.; Scholten, T.; Young, K.; Blume-Kohout, R. Gate Set Tomography. Quantum 2021, 5, 557,  DOI: 10.22331/q-2021-10-05-557
    142. 142
      Householder, A. The Theory of Matrices in Numerical Analysis; Dover Books on Mathematics; Dover Publications, 2013.
    143. 143
      Nielsen, E.; Rudinger, K.; Proctor, T.; Russo, A.; Young, K.; Blume-Kohout, R. Probing quantum processor performance with pyGSTi. Quantum Science and Technology 2020, 5, 044002  DOI: 10.1088/2058-9565/ab8aa4
    144. 144
      Proctor, T.; Revelle, M.; Nielsen, E.; Rudinger, K.; Lobser, D.; Maunz, P.; Blume-Kohout, R.; Young, K. Detecting and tracking drift in quantum information processors. Nat. Commun. 2020, 11, 5396,  DOI: 10.1038/s41467-020-19074-4
    145. 145
      Rudinger, K. M.; Ostrove, C. I.; Seritan, S. K.; Grace, M. D.; Nielsen, E.; Blume-Kohout, R. J.; Young, K. C. Two-Qubit Gate Set Tomography with Fewer Circuits. 2023; https://arxiv.org/abs/2307.15767 (accessed: 2025-02-23).
    146. 146
      Horodecki, M.; Horodecki, P.; Horodecki, R. General teleportation channel, singlet fraction, and quasidistillation. Phys. Rev. A 1999, 60, 18881898,  DOI: 10.1103/PhysRevA.60.1888
    147. 147
      Lu, D.; Li, H.; Trottier, D.-A.; Li, J.; Brodutch, A.; Krismanich, A. P.; Ghavami, A.; Dmitrienko, G. I.; Long, G.; Baugh, J.; Laflamme, R. Experimental Estimation of Average Fidelity of a Clifford Gate on a 7-qubit Quantum Processor. Phys. Rev. Lett. 2015, 114, 140505,  DOI: 10.1103/PhysRevLett.114.140505
    148. 148
      Fedorov, A.; Steffen, L.; Baur, M.; da Silva, M. P.; Wallraff, A. Implementation of a Toffoli Gate with Superconducting Circuits. Nature 2012, 481, 170172,  DOI: 10.1038/nature10713
    149. 149
      Wallman, J. J.; Flammia, S. T. Randomized benchmarking with confidence. New J. Phys. 2014, 16, 103032,  DOI: 10.1088/1367-2630/16/10/103032
    150. 150
      Roth, I.; Kueng, R.; Kimmel, S.; Liu, Y.-K.; Gross, D.; Eisert, J.; Kliesch, M. Recovering quantum gates from few average gate fidelities. Phys. Rev. Lett. 2018, 121, 170502,  DOI: 10.1103/PhysRevLett.121.170502
    151. 151
      Kalev, A.; Kyrillidis, A.; Linke, N. M. Validating and certifying stabilizer states. Phys. Rev. A 2019, 99, 042337  DOI: 10.1103/PhysRevA.99.042337
    152. 152
      Seshadri, A.; Ringbauer, M.; Spainhour, J.; Monz, T.; Becker, S. Theory of versatile fidelity estimation with confidence. Phys. Rev. A 2024, 110, 012431  DOI: 10.1103/PhysRevA.110.012431
    153. 153
      Seshadri, A.; Ringbauer, M.; Spainhour, J.; Blatt, R.; Monz, T.; Becker, S. Versatile Fidelity Estimation with Confidence. Phys. Rev. Lett. 2024, 133, 020402  DOI: 10.1103/PhysRevLett.133.020402
    154. 154
      Kosut, R. L.; Rabitz, H.; Grace, M. D. Adaptive quantum control via direct fidelity estimation and indirect model-based parametric process tomography. 52nd IEEE Conference on Decision and Control. Firenze, 2013; pp 12471252.
    155. 155
      Lu, Y.; Sim, J. Y.; Suzuki, J.; Englert, B.-G.; Ng, H. K. Direct estimation of minimum gate fidelity. Phys. Rev. A 2020, 102, 022410  DOI: 10.1103/PhysRevA.102.022410
    156. 156
      Liu, Y.; Otten, M.; Bassirianjahromi, R.; Jiang, L.; Fefferman, B. Benchmarking near-term quantum computers via random circuit sampling. 2022; http://arxiv.org/abs/2105.05232, arXiv: 2105.05232 [quant-ph] (accessed: 2025-02-23).
    157. 157
      Elements of Information Theory; John Wiley & Sons, Ltd, 2005; Chapter 4, pp 71101.
    158. 158
      Ledoux, M. The Concentration of Measure Phenomenon; American Mathematical Soc., 2001; Google-Books-ID: rRKChz4Om6gC.
    159. 159
      Boixo, S.; Smelyanskiy, V. N.; Neven, H. Fourier analysis of sampling from noisy chaotic quantum circuits. 2017; http://arxiv.org/abs/1708.01875, arXiv: 1708.01875 [quant-ph] (accessed: 2025-02-23).
    160. 160
      Choi, J.; Shaw, A. L.; Madjarov, I. S.; Xie, X.; Finkelstein, R.; Covey, J. P.; Cotler, J. S.; Mark, D. K.; Huang, H.-Y.; Kale, A.; Pichler, H.; Brandão, F. G. S. L.; Choi, S.; Endres, M. Preparing random states and benchmarking with many-body quantum chaos. Nature 2023, 613, 468473,  DOI: 10.1038/s41586-022-05442-1
    161. 161
      Rinott, Y.; Shoham, T.; Kalai, G. Statistical Aspects of the Quantum Supremacy Demonstration. Statistical Science 2022, 37, 322347
    162. 162
      Dalzell, A. M.; Hunter-Jones, N.; Brandão, F. G. S. L. Random Quantum Circuits Anticoncentrate in Log Depth. PRX Quantum 2022, 3, 010333
    163. 163
      Huang, C.; Zhang, F.; Newman, M.; Cai, J.; Gao, X.; Tian, Z.; Wu, J.; Xu, H.; Yu, H.; Yuan, B.; Szegedy, M.; Shi, Y.; Chen, J. Classical Simulation of Quantum Supremacy Circuits. arXiv 2020 (accessed 2025-02-23). DOI: 10.48550/arXiv.2005.06787
    164. 164
      Zlokapa, A.; Villalonga, B.; Boixo, S.; Lidar, D. A. Boundaries of quantum supremacy via random circuit sampling. npj Quantum Information 2023, 9, 36,  DOI: 10.1038/s41534-023-00703-x
    165. 165
      Kechedzhi, K.; Isakov, S. V.; Mandrà, S.; Villalonga, B.; Mi, X.; Boixo, S.; Smelyanskiy, V. Effective quantum volume, fidelity and computational cost of noisy quantum processing experiments. Future Generation Computer Systems 2024, 153, 431441,  DOI: 10.1016/j.future.2023.12.002
    166. 166
      Arute, F. Quantum supremacy using a programmable superconducting processor. Nature 2019, 574, 505510,  DOI: 10.1038/s41586-019-1666-5
    167. 167
      Harrow, A. W.; Mehraban, S. Approximate Unitary t-Designs by Short Random Quantum Circuits Using Nearest-Neighbor and Long-Range Gates. Communications in Mathematical Physics 2023, 401, 15311626,  DOI: 10.1007/s00220-023-04675-z
    168. 168
      Haferkamp, J.; Montealegre-Mora, F.; Heinrich, M.; Eisert, J.; Gross, D.; Roth, I. Efficient Unitary Designs with a System-Size Independent Number of Non-Clifford Gates. Communications in Mathematical Physics 2023, 397, 9951041,  DOI: 10.1007/s00220-022-04507-6
    169. 169
      Gao, X.; Kalinowski, M.; Chou, C.-N.; Lukin, M. D.; Barak, B.; Choi, S. Limitations of Linear Cross-Entropy as a Measure for Quantum Advantage. PRX Quantum 2024, 5, 010334  DOI: 10.1103/PRXQuantum.5.010334
    170. 170
      Tripathi, V.; Kowsari, D.; Saurav, K.; Zhang, H.; M.Levenson-Falk, E.; Lidar, D. A. Benchmarking quantum gates and circuits. 2024; https://arxiv.org/abs/2407.09942 (accessed: 2025-02-23).
    171. 171
      Burnett, J. J.; Bengtsson, A.; Scigliuzzo, M.; Niepce, D.; Kudra, M.; Delsing, P.; Bylander, J. Decoherence benchmarking of superconducting qubits. npj Quantum Information 2019, 5, 54,  DOI: 10.1038/s41534-019-0168-5
    172. 172
      Hahn, E. L. Spin Echoes. Phys. Rev. 1950, 80, 580594,  DOI: 10.1103/PhysRev.80.580
    173. 173
      Carr, H. Y.; Purcell, E. M. Effects of Diffusion on Free Precession in Nuclear Magnetic Resonance Experiments. Phys. Rev. 1954, 94, 630638,  DOI: 10.1103/PhysRev.94.630
    174. 174
      Meiboom, S.; Gill, D. Modified Spin?Echo Method for Measuring Nuclear Relaxation Times. Rev. Sci. Instrum. 1958, 29, 688691,  DOI: 10.1063/1.1716296
    175. 175
      Viola, L.; Lloyd, S. Dynamical suppression of decoherence in two-state quantum systems. Phys. Rev. A 1998, 58, 27332744,  DOI: 10.1103/PhysRevA.58.2733
    176. 176
      Uhrig, G. S. Keeping a Quantum Bit Alive by Optimized π-Pulse Sequences. Phys. Rev. Lett. 2007, 98, 100504,  DOI: 10.1103/PhysRevLett.98.100504
    177. 177
      Suter, D.; Álvarez, G. A. Colloquium: Protecting quantum information against environmental noise. Rev. Mod. Phys. 2016, 88, 041001  DOI: 10.1103/RevModPhys.88.041001
    178. 178
      Jurcevic, P. Demonstration of Quantum Volume 64 on a Superconducting Quantum Computing System. Quantum Sci. Technol. 2021, 6, 025020  DOI: 10.1088/2058-9565/abe519
    179. 179
      Pokharel, B.; Lidar, D. A. Demonstration of Algorithmic Quantum Speedup. Phys. Rev. Lett. 2023, 130, 210602,  DOI: 10.1103/PhysRevLett.130.210602
    180. 180
      Tripathi, V.; Chen, H.; Khezri, M.; Yip, K.-W.; Levenson-Falk, E.; Lidar, D. A. Suppression of Crosstalk in Superconducting Qubits Using Dynamical Decoupling. Phys. Rev. Appl. 2022, 18, 024068  DOI: 10.1103/PhysRevApplied.18.024068
    181. 181
      Zhou, Z.; Sitler, R.; Oda, Y.; Schultz, K.; Quiroz, G. Quantum Crosstalk Robust Quantum Control. Phys. Rev. Lett. 2023, 131, 210802,  DOI: 10.1103/PhysRevLett.131.210802
    182. 182
      Bäumer, E.; Tripathi, V.; Wang, D. S.; Rall, P.; Chen, E. H.; Majumder, S.; Seif, A.; Minev, Z. K. Efficient Long-Range Entanglement using Dynamic Circuits. 2023.
    183. 183
      Shirizly, L.; Misguich, G.; Landa, H. Dissipative Dynamics of Graph-State Stabilizers with Superconducting Qubits. Phys. Rev. Lett. 2024, 132, 010601  DOI: 10.1103/PhysRevLett.132.010601
    184. 184
      Niu, S.; Todri-Sanial, A.; Bronn, N. T. Multi-qubit dynamical decoupling for enhanced crosstalk suppression. Quantum Science and Technology 2024, 9, 045003  DOI: 10.1088/2058-9565/ad5a37
    185. 185
      Evert, B.; Izquierdo, Z. G.; Sud, J.; Hu, H.-Y.; Grabbe, S.; Rieffel, E. G.; Reagor, M. J.; Wang, Z. Syncopated Dynamical Decoupling for Suppressing Crosstalk in Quantum Circuits. 2024; https://arxiv.org/abs/2403.07836 (accessed: 2025-02-23).
    186. 186
      Brown, A. F.; Lidar, D. A. Efficient Chromatic-Number-Based Multi-Qubit Decoherence and Crosstalk Suppression. 2024. https://arxiv.org/abs/2406.13901(accessed: 2025-02-23).
    187. 187
      Martinis, J. M.; Geller, M. R. Fast adiabatic qubit gates using only σz control. Phys. Rev. A 2014, 90, 022307  DOI: 10.1103/PhysRevA.90.022307
    188. 188
      Krantz, P.; Kjaergaard, M.; Yan, F.; Orlando, T. P.; Gustavsson, S.; Oliver, W. D. A quantum engineer’s guide to superconducting qubits. Applied Physics Reviews 2019, 6, 021318  DOI: 10.1063/1.5089550
    189. 189
      Genov, G. T.; Schraft, D.; Vitanov, N. V.; Halfmann, T. Arbitrarily Accurate Pulse Sequences for Robust Dynamical Decoupling. Phys. Rev. Lett. 2017, 118, 133202,  DOI: 10.1103/PhysRevLett.118.133202
    190. 190
      Motzoi, F.; Gambetta, J. M.; Rebentrost, P.; Wilhelm, F. K. Simple Pulses for Elimination of Leakage in Weakly Nonlinear Qubits. Phys. Rev. Lett. 2009, 103, 110501,  DOI: 10.1103/PhysRevLett.103.110501
    191. 191
      Chow, J. M.; Gambetta, J. M.; Tornberg, L.; Koch, J.; Bishop, L. S.; Houck, A. A.; Johnson, B. R.; Frunzio, L.; Girvin, S. M.; Schoelkopf, R. J. Randomized Benchmarking and Process Tomography for Gate Errors in a Solid-State Qubit. Phys. Rev. Lett. 2009, 102, 090502  DOI: 10.1103/PhysRevLett.102.090502
    192. 192
      Albash, T.; Boixo, S.; Lidar, D. A.; Zanardi, P. Quantum adiabatic Markovian master equations. New J. of Phys. 2012, 14, 123016,  DOI: 10.1088/1367-2630/14/12/123016
    193. 193
      Dann, R.; Levy, A.; Kosloff, R. Time-dependent Markovian quantum master equation. Phys. Rev. A 2018, 98, 052129  DOI: 10.1103/PhysRevA.98.052129
    194. 194
      Campos Venuti, L.; Lidar, D. A. Error reduction in quantum annealing using boundary cancellation: Only the end matters. Phys. Rev. A 2018, 98, 022315  DOI: 10.1103/PhysRevA.98.022315
    195. 195
      Mozgunov, E.; Lidar, D. Completely positive master equation for arbitrary driving and small level spacing. Quantum 2020, 4, 227,  DOI: 10.22331/q-2020-02-06-227
    196. 196
      Nathan, F.; Rudner, M. S. Universal Lindblad equation for open quantum systems. Phys. Rev. B 2020, 102, 115109,  DOI: 10.1103/PhysRevB.102.115109
    197. 197
      Davidović, D. Completely Positive, Simple, and Possibly Highly Accurate Approximation of the Redfield Equation. Quantum 2020, 4, 326,  DOI: 10.22331/q-2020-09-21-326
    198. 198
      Davidović, D. Geometric-arithmetic master equation in large and fast open quantum systems. Journal of Physics A: Mathematical and Theoretical 2022, 55, 455301,  DOI: 10.1088/1751-8121/ac9f30
    199. 199
      Kasatkin, V.; Gu, L.; Lidar, D. A. Which differential equations correspond to the Lindblad equation?. Physical Review Research 2023, 5, 043163  DOI: 10.1103/PhysRevResearch.5.043163
    200. 200
      Perko, L. Differential Equations and Dynamical Systems; Springer New York: New York, NY, 2001; pp 163.
    201. 201
      Geerlings, K.; Leghtas, Z.; Pop, I. M.; Shankar, S.; Frunzio, L.; Schoelkopf, R. J.; Mirrahimi, M.; Devoret, M. H. Demonstrating a Driven Reset Protocol for a Superconducting Qubit. Phys. Rev. Lett. 2013, 110, 120501,  DOI: 10.1103/PhysRevLett.110.120501
    202. 202
      Goss, N.; Morvan, A.; Marinelli, B.; Mitchell, B. K.; Nguyen, L. B.; Naik, R. K.; Chen, L.; Jünger, C.; Kreikebaum, J. M.; Santiago, D. I.; Wallman, J. J.; Siddiqi, I. High-fidelity qutrit entangling gates for superconducting circuits. Nat. Commun. 2022, 13, 7481,  DOI: 10.1038/s41467-022-34851-z
    203. 203
      Tripathi, V.; Goss, N.; Vezvaee, A.; Nguyen, L. B.; Siddiqi, I.; Lidar, D. A. Qudit Dynamical Decoupling on a Superconducting Quantum Processor. Phys. Rev. Lett. 2025, 134, 050601  DOI: 10.1103/PhysRevLett.134.050601
    204. 204
      Krishnan Vijayan, M.; Paler, A.; Gavriel, J.; Myers, C. R.; Rohde, P. P.; Devitt, S. J. Compilation of algorithm-specific graph states for quantum circuits. Quantum Science and Technology 2024, 9, 025005  DOI: 10.1088/2058-9565/ad1f39
    205. 205
      Beverland, M. E.; Murali, P.; Troyer, M.; Svore, K. M.; Hoefler, T.; Kliuchnikov, V.; Low, G. H.; Soeken, M.; Sundaram, A.; Vaschillo, A. Assessing requirements to scale to practical quantum advantage. 2022. https://arxiv.org/abs/2211.07629 (accessed 2025-02-23).
    206. 206
      Agrawal, A. A.; Job, J.; Wilson, T. L.; Saadatmand, S. N.; Hodson, M. J.; Mutus, J. Y.; Caesura, A.; Johnson, P. D.; Elenewski, J. E.; Morrell, K. J.; Kemper, A. F. Quantifying fault tolerant simulation of strongly correlated systems using the Fermi-Hubbard model. 2024; https://arxiv.org/abs/2406.06511(accessed: 2025-02-23).
    207. 207
      Barends, R. Coherent Josephson Qubit Suitable for Scalable Quantum Integrated Circuits. Phys. Rev. Lett. 2013, 111, 080502  DOI: 10.1103/PhysRevLett.111.080502
    208. 208
      Shanto, S.; Kuo, A.; Miyamoto, C.; Zhang, H.; Maurya, V.; Vlachos, E.; Hecht, M.; Shum, C. W.; Levenson-Falk, E. SQuADDS: A Validated Design Database and Simulation Workflow for Superconducting Qubit Design. Quantum 2024, 8, 1465,  DOI: 10.22331/q-2024-09-09-1465
    209. 209
      Shah, P. A. . qiskit-community/qiskit-metal: Qiskit Metal 0.1.5a1 (2023). 2023;  DOI: 10.5281/zenodo.8061781 (accessed 2025-02-23).
    210. 210
      Minev, Z. K.; Leghtas, Z.; Reinhold, P.; Mundhada, S. O.; Diringer, A.; Hillel, D. C.; Wang, D. Z.-R.; Facchini, M.; Shah, P. A.; Devoret, M. pyEPR: The energy-participation-ratio (EPR) open- source framework for quantum device design. 2021;  DOI: 10.5281/zenodo.4744448 , https://github.com/zlatko-minev/pyEPR; https://pyepr-docs.readthedocs.io/en/latest/ (accessed 2025-02-23).
    211. 211
      Chen, Z. Measuring and Suppressing Quantum State Leakage in a Superconducting Qubit. Phys. Rev. Lett. 2016, 116, 020501  DOI: 10.1103/PhysRevLett.116.020501