This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Can shallow quantum circuits scramble local noise into global white noise?

Jonathan Foldager jonf@dtu.dk Department of Applied Mathematics and Computer Science, Technical University of Denmark, Denmark Department of Materials, University of Oxford, Parks Road, Oxford OX1 3PH, United Kingdom    Bálint Koczor balint.koczor@materials.ox.ac.uk Department of Materials, University of Oxford, Parks Road, Oxford OX1 3PH, United Kingdom Quantum Motion, 9 Sterling Way, London N7 9HJ, United Kingdom
Abstract

Shallow quantum circuits are believed to be the most promising candidates for achieving early practical quantum advantage – this has motivated the development of a broad range of error mitigation techniques whose performance generally improves when the quantum state is well approximated by a global depolarising (white) noise model. While it has been crucial for demonstrating quantum supremacy that random circuits scramble local noise into global white noise—a property that has been proved rigorously—we investigate to what degree practical shallow quantum circuits scramble local noise into global white noise. We define two key metrics as (a) density matrix eigenvalue uniformity and (b) commutator norm. While the former determines the distance from white noise, the latter determines the performance of purification based error mitigation. We derive analytical approximate bounds on their scaling and find in most cases they nicely match numerical results. On the other hand, we simulate a broad class of practical quantum circuits and find that white noise is in certain cases a bad approximation posing significant limitations on the performance of some of the simpler error mitigation schemes. On a positive note, we find in all cases that the commutator norm is sufficiently small guaranteeing a very good performance of purification-based error mitigation. Lastly, we identify techniques that may decrease both metrics, such as increasing the dimensionality of the dynamical Lie algebra by gate insertions or randomised compiling.

I Introduction

Current generations of quantum hardware can already significantly outperform classical computers in random sampling tasks [1, 2] and hopefully future hardware developments will enable powerful applications in quantum machine learning [3], fundamental physics [4, 5] and in developing novel drugs and materials [6, 7, 8, 9]. The scale and precision of the technology today is, however, still below what is required for fully fault-tolerant quantum computation: Due to noise accumulation in the noisy intermediate-scale quantum (NISQ) era [10], one is thus limited to only shallow-depth quantum circuits which led to the development of a broad range of hybrid quantum-classical protocols and quantum machine learning algorithms [11, 12, 13].

The aim in this paradigm is to prevent excessive error buildup via a parameterised, shallow-depth quantum circuit and then perform a series of repeated measurements in order to extract expected values. These expected values are then post processed on a classical computer in order to update the parameters of the circuits, e.g., as part of a training procedure. A major challenge is the potential need for an excessive number of circuit repetitions which, however, can be significantly suppressed by the use of advanced training algorithms [14, 15, 16] or via classical-shadows-based protocols [17, 18, 19]. As such, the primary limitation of near-term applications is the damaging effect of gate noise on the estimated expected values which can only be reduced by advanced error mitigation techniques [20, 12].

Error mitigation comprises a broad collection of diverse techniques that generally aim to estimate precise expected values by suppressing the effect of hardware imperfections [20, 12]. Due to the diversity of techniques and due to the significant differences in the range of applicability, the need for performance metrics was recently emphasised [20]. This motivates the present work to characterise noise in typical practical circuits, e.g., in quantum simulation or in quantum machine learning, and define two key metrics that determine the performance of a broad class of error mitigation techniques: (a) eigenvalue uniformity as a closeness to global depolarising (white) noise and (b) norm of the commutator between the ideal and noisy quantum states. While (b) determines the performance of purification based error mitigation techniques [21, 22], (a) implies a good performance of all error mitigation techniques.

Our primary motivation is that gate errors in complex quantum circuits are scrambled into global white noise [1, 23]. This property has been proved for random circuits by establishing exponentially decreasing error bounds; surprisingly, in our numerical simulations we find that in many practical scenarios the same bounds apply relatively well. In particular, we find that both our metrics, (a) the distance from global-depolarising noise and (b) the commutator norm, are approximated as

f(ν)=αeξξ(1eξ)ν=αν+O(ξ),f(\nu)=\alpha\frac{e^{-\xi}\xi}{(1-e^{-\xi})\sqrt{\nu}}=\frac{\alpha}{\sqrt{\nu}}+O(\xi), (1)

where ν\nu is the number of gates in the quantum circuit, ξ\xi is the number of expected errors in the entire circuit and α\alpha is a constant. As such, if one keeps the error rate small ξ1\xi\ll 1 but increases the number of gates in a circuit then both (a) and (b) are expected to decrease. This is a highly desirable property in practice, e.g., white noise does not introduce a bias to the expected-value measurement but only a trivial, global scaling as we detail in the rest of this introduction.

In the present work we simulate a broad range of quantum circuits often used in practice and identify scenarios where this approximation holds well, by means of gate parameters and circuit structures are sufficiently random. We also identify strategies that improve scrambling local gate noise into global white noise, such as inserting additional gates into a circuit to increase the dimensionality of its Lie algebra [24]. In most cases, however, we conclude that white noise is not necessarily a good approximation due to the large prefactor α\alpha in Eq. 1. Thus the performance of some error mitigation techniques that rely on a global-depolarising noise assumption is limited. On the other hand, we find that in all cases the commutator norm, our other key metric, is smaller by at least 1-2 orders of magnitude guaranteeing a very good performance of purification-based error mitigation techniques.

Our work is structured as follows. In the rest of this introduction we briefly review global depolarising noise and how it can be exploited in error mitigation, and then briefly review purification-based error mitigation techniques and their performance. In Section II we introduce theoretical notions and finally in Section III we present our simulation results.

I.1 Global depolarisation and error mitigation

In the NISQ-era, we don’t have comprehensive solutions to error correction, which has led the field to develop error mitigation techniques. These techniques aim to extract expected values Oideal:=tr[Oρid]\langle O\rangle_{ideal}:=\tr[O\rho_{id}] of observables—typically some Hamiltonian of interest—with respect to an ideal noiseless quantum state ρid\rho_{id}.

A very simple error model, the global depolarising noise channel, has been very often considered as a relatively good approximation to complex quantum circuits. For qubit states, the channel mixes the ideal, noise-free state with the maximally mixed state Id/d\mathrm{Id}/d of dimension d=2Nd=2^{N} as

ρwn:=ηρid+(1η)Id/d.\rho_{wn}:=\eta\rho_{id}+(1-\eta)\mathrm{Id}/d. (2)

Here ηF\eta\approx F is a probability that approximates the fidelity as F=η+(1η)/dF=\eta+(1-\eta)/d. The white noise channel has been commonly used in the literature for modelling errors in near-term quantum computers [25] and, in particular, it has been shown to be a very good approximation to noise in random circuits [1, 23]. White noise is extremely convenient as it lets the user extract, after rescaling by η\eta, the ideal expected value of any traceless Hermitian observable OO via

Oideal=tr[Oρwn]/η.\langle O\rangle_{ideal}=\tr[O\rho_{wn}]/\eta. (3)

Of course, for small fidelities η1\eta\ll 1 the expected value tr[Oρwn]\tr[O\rho_{wn}] requires a significantly increased sampling to suppress shot noise. In this model, the scaling factor η\eta is a global property and can be estimated experimentally, e.g., via randomised measurements [25], via extrapolation [26] or via learning-based techniques [27].

Global depolarisation, however, may not be sufficiently accurate model to capture more subtle effects of gate noise and thus rescaling an experimentally estimated expected value yields a biased estimate of the ideal one as Obias:=tr[Oρ]/ηOideal\langle O\rangle_{bias}:=\tr[O\rho]/\eta-\langle O\rangle_{ideal}. The bias here Obias\langle O\rangle_{bias} is not a global property, i.e., it is specific to each observable, and requires the use of more advanced error mitigation techniques to suppress.

Intuitively, one expects the bias is small for quantum states that are well approximated by a global depolarising model as ρρwn\rho\approx\rho_{wn} and, indeed, we find a general upper bound in terms of the trace distance as

|Obias|=|tr[Oρ]tr[Oρwn]|ηOρρwn1η.|\langle O\rangle_{bias}|=\frac{|\tr[O\rho]-\tr[O\rho_{wn}]|}{\eta}\leq\frac{\lVert O\rVert_{\infty}\lVert\rho-\rho_{wn}\rVert_{1}}{\eta}. (4)

Here O\lVert O\rVert_{\infty} is the operator norm as the absolute largest eigenvalue of the traceless OO, refer to ref. [28] for a proof. As such, a small trace distance guarantees a small bias and thus indirectly determines the performance of all error mitigation techniques – and further protocols [29, 19].

In this work, we characterise how close noisy quantum states ρ\rho in practical applications approach white noise states ρwn\rho_{wn} and consider various types of variational quantum circuits that are typical for NISQ applications. When the above trace distance is small then it guarantees a small bias in expected values which allows us to nearly trivially mitigate the effect of gate noise, i.e., via a simple rescaling.

I.2 Purification-based error mitigation and the commutator norm

Another core metric we will consider is the commutator norm between the ideal and noisy quantum states as C:=[ρid,ρ]1\mathcal{E}_{C}:=\lVert[\rho_{id},\rho]\rVert_{1}, which determines the performance of purification based error mitigation techniques [28] – a small commutator norm has significant practical implications as it guarantees that one can accurately determine expected values using the ESD/VD [21, 22] error mitigation techniques. In particular, independently preparing nn copies of the noisy quantum state and applying a derangement circuit to entangle the copies, allows one to estimate the expected value

tr[ρnO]tr[ρn]=Oideal+ESD.\frac{\tr[\rho^{n}O]}{\tr[\rho^{n}]}=\langle O\rangle_{ideal}+\mathcal{E}_{ESD}.

The approach is very NISQ-friendly [30, 31] and its approximation error ESD\mathcal{E}_{ESD} approaches in exponential order a noise floor as we increase the number of copies nn [21]; This noise floor is determined generally by the commutator norm C\mathcal{E}_{C} but in the most typical applications of preparing eigenstates, the noise floor is quadratically smaller as C2\mathcal{E}_{C}^{2} [28].

Note that this commutator can vanish even if the quantum state is very far from a white noise state, in fact it generally vanishes when ρid\rho_{id} approximates an eigenvector of ρ\rho. When a state is close to the white noise approximation then a small commutator norm is guaranteed, however, we demonstrate that the latter is a much less stringent condition and a much better approximation in practice than the former: in all instances we find that the commutator norm is significantly smaller than the trace distance from white noise states.

II Theoretical Background

In this section we introduce the main theoretical notations and recapitulate the most relevant results from the literature.

II.1 General properties of noisy quantum states

Recall that any quantum state of dimension dd can be represented via its density matrix ρ\rho that generally admits the spectral decomposition as

ρ=k=1dλk|ψkψk|,\rho=\sum_{k=1}^{d}\lambda_{k}\outerproduct{\psi_{k}}{\psi_{k}}, (5)

where we focus on NN-qubit systems of dimension d=2Nd=2^{N}. Here λk\lambda_{k} are non-negative eigenvalues and |ψk\ket{\psi_{k}} are eigenvectors. Since iλi=1\sum_{i}\lambda_{i}=1, the spectrum λ¯\underline{\lambda} is also interpreted as a probability distribution.

If ρ\rho is prepared by a perfect, noise-free unitary circuit, only one eigenvalue is different from zero and the corresponding eigenvector is the ideal quantum state as |ψid\ket{\psi_{id}}. In contrast, an imperfect circuit prepares a ρ\rho that has more than one non-zero eigenvalues and is thus a probabilistic mixture of the pure quantum states |ψk\ket{\psi_{k}}, e.g., due to interactions with a surrounding environment. In fact, noisy quantum circuits that we typically encounter in practice produce quite particular structure of the eigenvalue distribution: one dominant component that approximates the ideal quantum state |ψ1|ψid\ket{\psi_{1}}\approx\ket{\psi_{id}} mixed with an exponentially growing number of “error” eigenvectors that have small eigenvalues. White noise is the limiting case where non-dominant eigenvalues are exponentially small 1/d\propto 1/d and |ψ1|ψid\ket{\psi_{1}}\approx\ket{\psi_{id}}.

The quality of the noisy quantum state is then defined by the probability of the ideal quantum state as the fidelity F:=ψid|ρ|ψidF:=\expectationvalue{\rho}{\psi_{id}}; We show in Appendix A that for any quantum state it approaches the dominant eigenvalue λ1\lambda_{1} as

λ1=F+O(C),\lambda_{1}=F+O(\mathcal{E}_{C}), (6)

where we compute the error term analytically in terms of the commutator norm C=[ρid,ρ]1\mathcal{E}_{C}=\lVert[\rho_{id},\rho]\rVert_{1} from Section I.1. This property is actually completely general and applies to any density matrix.

II.2 Practically motivated noise models

Most typical noise models used in practice, such as local depolarising or dephasing noise, admit the following probabilistic interpretation: a noisy gate operation Φ(ρ)\Phi(\rho) can be interpreted as a mixture of the noise-free operation UU that happens with probability 1ϵ1-\epsilon and an error contribution as

Φk(ρ)=(1ϵ)UkρUk+ϵΦerr(UkρUk).\Phi_{k}(\rho)=(1-\epsilon)U_{k}\rho U_{k}^{\dagger}+\epsilon\Phi_{err}(U_{k}\rho U_{k}^{\dagger}). (7)

Here UkU_{k} is the kthk^{th} ideal quantum gate and the completely positive trace-preserving (CPTP) map Φerr\Phi_{err} happens with probability ϵ\epsilon and accounts for all error events during the execution of a gate. A quantum circuit is then a composition of a series of ν\nu such quantum gates which prepares the convex combination as

ρ=ηρid+(1η)ρerr.\rho=\eta\rho_{id}+(1-\eta)\rho_{err}. (8)

Here ρid:=|ψidψid|\rho_{id}:=\outerproduct{\psi_{id}}{\psi_{id}} is the ideal noise-free state, ρerr\rho_{err} is an error density matrix and η=(1ϵ)ν\eta=(1-\epsilon)^{\nu} is the probability that none of the gates have undergone errors. This probability actually [28, 23] approximates the fidelity F:=ψid|ρ|ψidF:=\expectationvalue{\rho}{\psi_{id}} given the noise model in Eq. 7 as

F=(1ϵ)ν+F=eξ+F+O(ϵ2/ν).F=(1-\epsilon)^{\nu}+\mathcal{E}_{F}=e^{-\xi}+\mathcal{E}_{F}+O(\epsilon^{2}/\nu). (9)

Here we approximate (1ξ/ν)ν=eξ+O(ϵ2/ν)(1-\xi/\nu)^{\nu}=e^{-\xi}+O(\epsilon^{2}/\nu) for small ϵ\epsilon and large ν\nu where ξ:=ϵν\xi:=\epsilon\nu is the circuit error rate as the expected number of errors in a circuit. In practice the approximation error F=ψid|ρerr|ψid\mathcal{E}_{F}=\expectationvalue{\rho_{err}}{\psi_{id}} is typically small and in the limiting case of white noise it decreases exponentially as F=1/d\mathcal{E}_{F}=1/d due to ρerr=Id/d\rho_{err}=\mathrm{Id}/d.

Assuming sufficiently deep, complex circuits, ref. [28] obtained an approximate bound for the commutator between the ideal and noisy quantum states as

[ρid,ρ]1const×eξξ/ν.\lVert[\rho_{id},\rho]\rVert_{1}\lessapprox\mathrm{const}\times e^{-\xi}\xi/\sqrt{\nu}. (10)

This bound confirms that as we increase the number of quantum gates ν\nu in a circuit but keeping the circuit error rate ξ\xi constant, the commutator norm decreases as 1/ν\propto 1/\sqrt{\nu} [28]. Furthermore, this function closely resembles to Eq. 1 which is a central aim of this work to explore.

Refer to caption
Figure 1: Simulating families of 10-qubit Strong entangling layer (SEL) ansatz circuits [32] at random gate parameters for an increasing number ν\nu of gates and per-gate depolarising error rates ϵ\epsilon. (a) the uniformity measure W(ν)W(\nu) of the error eigenvalues of the density matrix from Eq. 12 closely match the theoretical model (dashed lines) for random circuits and confirm that increasing the number of gates in random circuits scrambles local noise into global white noise. (b) the commutator norm C(ν)C(\nu) from Eq. 14 is significantly smaller in absolute value and decreases with a larger polynomial degree (steeper slope of the dashed lines) than the uniformity measure – this suggests that the dominant eigenvector of the density matrix ρ\rho approximately commutes with ρ\rho even when noise is not well described by white noise. The ϵ0\epsilon\rightarrow 0 simulations were approximated using ϵ=108\epsilon=10^{-8} (ϵ=107\epsilon=10^{-7}) when calculating WW (CC).

II.3 White noise in random circuits 

Random circuits have enabled quantum supremacy experiments using noisy quantum computers for two primary reasons: (a) the outputs of these circuits are hard to simulate classically and (b) they render local noise into global white noise [1], hence introducing only a trivial bias to the ideal probability distribution similarly as in Section I.1.

Ref [23] considered random circuits consisting of ss two-qubit gates, each of which undergoes two single qubit (depolarising) errors each with probability ϵ~\tilde{\epsilon} (assuming single-qubit gates are noiseless). We can relate this to our model by identifying the local noise after each two-qubit gate with the error event in Eq. 7 via the probability ϵ=1(1ϵ~)2=2ϵ~ϵ~2\epsilon=1-(1-\tilde{\epsilon})^{2}=2\tilde{\epsilon}-\tilde{\epsilon}^{2}. Ref [23] then established the fidelity F~\tilde{F} of the quantum state which one obtains from a noisy cross-entropy score as

F~=e2sϵ~±O(sϵ~2)=eξ±O(ϵξ).\tilde{F}=e^{-2s\tilde{\epsilon}\pm O(s\tilde{\epsilon}^{2})}=e^{-\xi\pm O(\epsilon\xi)}.

This coincides with our approximation from Eq. 9 up to an additive error in the exponent which, however, diminishes for low gate error rates. In the following we will thus assume FF~F\equiv\tilde{F}.

Measuring these noisy states in a the standard measurement basis {|j}j=1d\{|j\rangle\}_{j=1}^{d} produces a noisy probability distribution p~noisy(j)=j|ρ|j\tilde{p}_{noisy}(j)=\langle j|\rho|j\rangle. Ref. [23] established that this probability distribution rapidly approaches the white noise approximation p~wn=Fpid+(1F)punif\tilde{p}_{wn}=Fp_{id}+(1-F)p_{unif}. In particular, the total variation distance (via the l1l_{1} norm x1=i|xi|\lVert x\rVert_{1}=\sum_{i}|x_{i}|) between the two probability distributions is upper bounded as

12p~noisyp~wn1O(Fϵν)=O(eξξ/ν).\tfrac{1}{2}\lVert\tilde{p}_{noisy}-\tilde{p}_{wn}\rVert_{1}\leq O(F\epsilon\sqrt{\nu})=O(e^{-\xi}\xi/\sqrt{\nu}). (11)

This expression is formally identical to the bound on the commutator norm in Eq. 10; Indeed if the noise in the quantum state approaches a white noise approximation, it implies that the commutator norm must also vanish in the same order.

On the other hand, the reverse is not necessarily true as Eq. 11 is a stronger condition than Eq. 10 as the latter only guarantees that the dominant eigenvector approaches |ψ1|ψid\ket{\psi_{1}}\approx\ket{\psi_{id}} but does not imply anything about the eigenvalue distribution of ρ\rho or ρerr\rho_{err}.

III Numerical simulations

III.1 Target metrics

In the NISQ-era comprehensive error correction will not be feasible and thus hope is primarily based on variational quantum algorithms [11, 12, 13, 33, 34]. In this paradigm a shallow, parametrised quantum circuit is used to prepare a parametrised quantum state that aims to approximate the solution to a given problem, typically the ground state of a problem Hamiltonian. Due to its shallow depth the ansatz circuit is believed to be error robust and its tractable parametrisation allows to explore the Hilbert space near the solution. On the other hand, such circuits are structurally quite different than random quantum circuits and it was already raised in ref. [23] whether error bounds on the white noise approximation extend to these shallow quantum circuits.

We simulate such quantum circuits under the effect of local depolarising noise – while note that a broad class of local coherent and incoherent error models can effectively be transformed into local depolarising noise using, e.g., twirling techniques or randomised compiling [35, 36, 37, 38]. We analyse the resulting noisy density matrix ρ\rho by calculating the following two quantities. First, we quantify ‘closeness’ to a white noise state from Eq. 2 by computing uniformity measure WW as the l1l_{1}-distance between the uniform distribution and the non-dominant eigenvalues of the output state as

W:=12perrpunif1=12k=2d|λk1λ11d1|,W:=\frac{1}{2}\lVert p_{err}-p_{unif}\rVert_{1}=\frac{1}{2}\sum_{k=2}^{d}|\frac{\lambda_{k}}{1-\lambda_{1}}-\frac{1}{d-1}|, (12)

which only depends on spectral properties of the quantum state and can thus be computed straightforwardly. We show in 2 that WW is proportional to the trace distance from a white noise quantum state as

ρρwn1=(1λ1)W+w,\lVert\rho-\rho_{wn}\rVert_{1}=(1-\lambda_{1})W+\mathcal{E}_{w}, (13)

uo to a bounded error w\mathcal{E}_{w}. The uniformity measure WW thus determines the bias in estimating any traceless expected value as discussed in Section I.1.

Second, we calculate the commutator norm C\mathcal{E}_{C} from Section I.1 relative to 1λ11-\lambda_{1} as

C:=[ρid,ρ]11λ1=[ρid,ρerr]1+𝒪(q),C:=\frac{\lVert[\rho_{id},\rho]\rVert_{1}}{1-\lambda_{1}}=\lVert[\rho_{id},\rho_{err}]\rVert_{1}+\mathcal{O}(\mathcal{E}_{q}), (14)

which we relate to the commutator norm between the “error part” of the state ρerr\rho_{err} and the ideal quantum state ρid\rho_{id} in Lemma 1. In the following, we will refer to CC as the commutator norm – and recall that it determines the ultimate performance of purification-based error mitigation as discussed in Section I.1.

Refer to caption
Figure 2: XXX Hamiltonian: same simulations as in Fig. 1 but using 10-qubit HVA quantum circuits constructed for the XXX spin problem Hamiltonian. (a, c) at randomly chosen circuit parameters of the HVA we find the same conclusions as for random circuits in Fig. 1. (b) when the HVA circuit approximates the ground state of the Hamiltonian (VQE parameters) we find the noise in the circuit is not approximated well by white noise, i.e., the uniformity measure W(ν)W(\nu) is large and does not decrease as we increase ν\nu. (d) On the other hand, the commutator norm C(ν)C(\nu) is significantly smaller than W(ν)W(\nu) confirming that the the ideal quantum state approximately commutes with the noisy one. The ϵ0\epsilon\rightarrow 0 simulations were approximated using ϵ=108\epsilon=10^{-8} (ϵ=107\epsilon=10^{-7}) when calculating WW (CC).

III.2 Random states via Strong Entangling ansätze 

We first consider a Strong Entangling ansatz (SEA): it is built of alternating layers of parametrised single-qubit rotations followed by a series of nearest-neighbour CNOT gates as illustrated in Fig. 5 – and assume a local depolarising noise with probability ϵ\epsilon. We simulate random quantum circuits by randomly generating parameters |θk|2π|\theta_{k}|\leq 2\pi – note that these circuits are not necessarily Haar-random distributed and thus results in Section II.3 do not necessarily apply.

We simulate 1010-qubit circuits and in Fig. 1 (a) we plot the eigenvalue uniformity W(ν)W(\nu) while in Fig. 1 (b) we plot the commutator norm C(ν)C(\nu) for an increasing number ν\nu of quantum gates – all datapoints are averages over ten random seeds. These results surprisingly well recover the expected behaviour of random quantum circuits as for small error rates ϵ0\epsilon\rightarrow 0 both quantities W(ν)W(\nu) and C(ν)C(\nu) can be approximated by the function from Eq. 1 as we now discuss.

In Section II.3 we stated bounds of ref. [23] on the distance between p~noisy\tilde{p}_{noisy} and p~wn\tilde{p}_{wn}. Based on the assumption that these bounds also apply to the probability distributions pnoisy=ψk|ρ|ψkp_{noisy}=\bra{\psi_{k}}\rho\ket{\psi_{k}} and pwn:=ψk|ρwn|ψkp_{wn}:=\bra{\psi_{k}}\rho_{wn}\ket{\psi_{k}} we derive in 4 the approximate bound on the eigenvalue uniformity as

W=O(eξξ/ν1eξ).W=O\left(\frac{e^{-\xi}\xi/\sqrt{\nu}}{1-e^{-\xi}}\right).

Furthermore, by combining Eq. 14 and the bound in Eq. 10 we find that the commutator norm CC is similarly bounded by the same function. On the other, Fig. 1 (b) suggests that the commutator norm decreases with a larger polynomial degree and thus we approximate both W(ν)W(\nu) and C(ν)C(\nu) using the function

f(ν)=αξeξνβ(1eξ)=α/νβ+𝒪(ξ)f(\nu)=\alpha\frac{\xi e^{-\xi}}{\nu^{\beta}(1-e^{-\xi})}=\alpha/\nu^{\beta}+\mathcal{O}(\xi) (15)

where we fit the two parameters α\alpha and β\beta to our simulated dataset. The second equation above is an expansion for small circuit error rates ξ\xi as detailed in Section A.2.1. Indeed, in Fig. 1 (blue circles) for small ϵ0\epsilon\rightarrow 0 we observe a nearly linear function in the log-log plot in Fig. 1 and thus remarkably well recover the theoretical bounds with the polynomial power approaching b1/2b\rightarrow 1/2.

Refer to caption
Figure 3: TFI same simulations as in Fig. 1 but using 10-qubit HVA quantum circuits constructed for the TFI spin problem Hamiltonian. (a, c) at randomly chosen circuit parameters W(ν)W(\nu) decreases more slowly, in smaller polynomial order than random circuits – see text and see simulations with added layers of RzR_{z} gates in Fig. 6. (b) at the VQE parameters white noise is again not a good approximation, i.e., the uniformity measure W(ν)W(\nu) is large and does not decrease as we increase ν\nu. (d) the commutator norm C(ν)C(\nu) is smaller than W(ν)W(\nu) in absolute value by an order of magnitude. The ϵ0\epsilon\rightarrow 0 simulations were approximated using ϵ=108\epsilon=10^{-8} (ϵ=107\epsilon=10^{-7}) when calculating WW (CC).

Furthermore, comparing Fig. 1 (b, blue circles) and Fig. 1 (a, blue circles) suggest that the commutator norm has both a significantly smaller absolute value (smaller α\alpha) and decreases at a faster polynomial rate (larger beta) than the uniformity measure. In fact, the commutator norm is more than two orders of magnitude smaller than the uniformity measure which suggests that even when ρerr\rho_{err} is not approximated well by a white noise state it, nevertheless, almost commutes with the ideal pure state ρid\rho_{id}.

We finally consider how the absolute factor α\alpha depends on the number of qubits: we perform simulations at a small error rate ϵ0\epsilon\rightarrow 0 and fit our model function ανβ\alpha\nu^{\beta} to extract α(N)\alpha(N) for an increasing number of qubits. The results are plotted in Fig. 7 (e) and suggest that the prefactor α(N)\alpha(N) initially grows slowly but then saturates while note that a polylogarithmic depth is sufficient to reach anticoncentration [23].

III.3 Variational Hamiltonian Ansatz

Theoretical results guarantee that the SEL ansatz initialised at random parameters approach for an increasing depth unitary 2-designs thereby reproducing properties of random quantum circuits [39, 40]. It is thus not surprising that the model introduced in Section II.3 gives a remarkably good agreement between the SEL ansatz (dots on in Fig. 1) and genuine random circuits (fits as continuous lines in Fig. 1).

Here we consider the Hamiltonian Variational Ansatz (HVA) [41, 42] at more practical parameter settings: The HVA has the advantage that we can efficiently obtain parameters that increasingly better (as we increase the ansatz depth) approximate the ground state of a problem Hamiltonian – we will refer to these as VQE parameters. We also want to compare this circuit against random circuits and thus also simulate the HVA such that every gate receives a random parameter as detailed in Section B.1.

While the VQE parameter settings capture the relevant behaviour in practice as one approaches a solution, the random parameters are more relevant, e.g., at the early stages of a VQE parameter optimisation. Furthermore, as the circuit is entirely composed by Pauli terms in the problem Hamiltonian, the dimensionality of its dynamical Lie algebra is entirely determined by the problem Hamiltonian in contrast to an exponentially growing algebra of the SEL ansatz [24].

III.4 Heisenberg XXX spin model

We first consider a VQE problem of finding the ground state of the 1-dimensional XXX spin-chain model. We construct the HVA ansatz from Section III.3 for this problem Hamiltonian as a sum XXX=0+1\mathcal{H}_{\text{XXX}}=\mathcal{H}_{0}+\mathcal{H}_{1} as

0=k=1NΔkZk,1=k=1N[XkXk+1+YkYk+1+ZkZk+1].\mathcal{H}_{0}=\sum_{k=1}^{N}\Delta_{k}Z_{k},\,\mathcal{H}_{1}=\sum_{k=1}^{N}[X_{k}X_{k+1}{+}Y_{k}Y_{k+1}{+}Z_{k}Z_{k+1}].

The Pauli operators XXXX, YYYY and ZZZZ determine couplings between nearest neighbour spins in a 1-D chain and we choose them to be of unit strength. Furthermore, ZkZ_{k} are local on-site interactions |Δk|1|\Delta_{k}|\leq 1 that were generate uniformly randomly such that the Hamiltonian has a non-trivial ground state.

First, we simulate the HVA ansatz for N=10N=10 qubits with randomly generated circuit parameters as |θk|2π|\theta_{k}|\leq 2\pi and plot results for an increasing number of quantum gates in Fig. 2 (a, c). We a find similar behaviour for the eigenvalue uniformity W(ν)W(\nu) as with random SEL circuits in Fig. 1 (a) and obtain a reasonably good fit for ϵ0\epsilon\rightarrow 0 using our model function from Eq. 15. The commutator norm in Fig. 2 (c) is again significantly smaller in magnitude than the uniformity measure and decreases faster with a higher polynomial order similarly to as with the random SEL ansatz in Fig. 1 (b) .

Second, in Fig. 2 (b,d) we simulate the ansatz at the VQE parameters that approximate the ground state. Since the ansatz parameters become very small as one approaches an adiabatic evolution, it is not surprising that the output density matrix is not well-approximated by a white noise state: the uniformity measure is very large in Fig. 2 (b). The commutator norm in Fig. 2 (d) again, is significantly smaller than W(ν)W(\nu) and although it appears to slowly grow with ν\nu, it appears to decrease for ν\nu\rightarrow\infty. This agrees with observations of ref. [28] that the circuits need not be random for the commutator to be sufficiently small in practice.

Furthermore, in Fig. 7 (a, b) we investigate the dependence on NN and find that the prefactor α\alpha grows slowly and appears to saturate for N10N\geq 10.

Refer to caption
Figure 4: LiH same simulations as in Fig. 1 but using 6-qubit HVA quantum circuits constructed for a LiH molecular Hamiltonian. (a, c) at randomly chosen circuit parameters both W(ν)W(\nu) and C(ν)C(\nu) decrease as expected for random circuits due our randomised compiling strategy [43, 44]. (b) at the VQE parameters white noise is an increasingly bad approximation, i.e., the uniformity measure W(ν)W(\nu) increases as we increase ν\nu. (d) the commutator norm C(ν)C(\nu) is smaller than W(ν)W(\nu) in absolute value by 2 orders of magnitude. The ϵ0\epsilon\rightarrow 0 simulations were approximated using ϵ=108\epsilon=10^{-8} (ϵ=107\epsilon=10^{-7}) when calculating WW (CC).

III.5 TFI

In the next example we consider the transverse-field Ising (TFI) model TFI=0+1\mathcal{H}_{\text{TFI}}=\mathcal{H}_{0}+\mathcal{H}_{1} using constant on-site interactions hi=1h_{i}=1 and randomly generated coupling strengths |Ji|1|J_{i}|\leq 1 as

0=ihiXi,1=iJiZiZi+1.\mathcal{H}_{0}=-\sum_{i}h_{i}X_{i},\quad\mathcal{H}_{1}=-\sum_{i}J_{i}Z_{i}Z_{i+1}. (16)

We first simulate the HVA ansatz with random variational parameters in Fig. 3 (a, c). While at small error rates ϵ0\epsilon\rightarrow 0 Fig. 3 (a, blue) can be fitted well with our polynomial approximation form Eq. 15, we observe that the eigenvalue uniformity W(ν)W(\nu) in Fig. 3 (a, blue) decreases with a small polynomial degree.

Indeed, as the HVA ansatz is specific to a particular Hamiltonian, its dynamical Lie algebra may have a low dimensionality [24] resulting in a limited ability to scramble local noise into white noise; this explains why in Fig. 3 (a) the uniformity measure decreases more slowly, i.e., in a smaller polynomial order, than random circuits. For this reason, we additionally simulate in Fig. 6 the TFI-HVA ansatz but with adding RzR_{z} gates in each layer whose generator is not contained in the problem Hamiltonian. The increased dimensionality of the dynamic Lie algebra, indeed, improves scrambling as the white noise approximation is clearly better in Fig. 6 – while note that the increased dimensionality may also lead to exponential inefficiencies in training the circuit [24].

In stark contrast to the case of the uniformity measure W(ν)W(\nu), we find that the commutator norm in Fig. 3 (c, blue) decreases substantially for an increasing ν\nu despite the low dimensionality of the Lie algebra. This nicely demonstrates that a small commutator norm is a much more relaxed condition than white noise as the latter requires that the noise is fully scrambled in the entirety of the exponentially large Hilbert space. Finally, we simulate the TFI circuits at VQE parameters and find qualitatively the same behaviour as in the case of the XXX problem.

III.6 Quantum Chemistry: LiH

We consider a 6-qubit Lithium Hydride (LiH) Hamiltonian in the Jordan-Wigner encoding as a linear combination of non-local Pauli strings Pk{Id,X,Y,Z}NP_{k}\in\{\mathrm{Id},X,Y,Z\}^{\otimes N} as

LiH=k=1rhhkPk.\mathcal{H}_{LiH}=\sum_{k=1}^{r_{h}}h_{k}P_{k}. (17)

We construct the HVA ansatz by splitting this Hamiltonian into two parts with 0\mathcal{H}_{0} being composed of the diagonal Pauli terms in Eq. 17 while 1\mathcal{H}_{1} composed of non-diagonal Pauli strings.

Such chemical Hamiltonians typically have a very large number of terms with rh1r_{h}\gg 1 but a significant fraction only have small weights hkh_{k} thus the HVA would have a large number of gates with only very small rotation angles. For these reasons we construct a more efficient circuit whose basic building blocks are constructed using sparse compilation techniques [43, 44]: Each single layer in the HVA ansatz consists of gates that correspond to 100100 randomly selected terms of the Hamiltonian with sampling probabilities pk|hk|p_{k}\propto|h_{k}| proportional to the Pauli coefficients. This approach has the added benefit that it makes the circuit structure random as opposed to the fixed structures in Section III.4 and in Section III.5.

Results shown in Fig. 4 (a,c) agree with our findings from the previous sections: at randomly chosen circuit parameters the uniformity measure decreases according to Eq. 15; the commutator norm similarly decreases but in a higher polynomial order while its absolute value is smaller by at least an order of magnitude. In contrast, Fig. 4 (b) suggests that the errors are not well approximated by white noise with a large and non-decreasing W(ν)0.5W(\nu)\approx 0.5. Furthermore, Fig. 4 (b) again confirms that despite white noise is not a good approximation, the commutator norm is small in absolute value, i.e., 103\approx 10^{-3} in the practically relevant region. This guarantees a very good performance of the ESD/VD error mitigation techniques sufficient for nearly all practical purposes.

IV Discussion

Random quantum circuits—instrumental for demonstrating quantum advantage—are known to scramble local gate noise into global white noise for sufficiently long circuit depths [1]: general bounds have been proved on the approximation error which decrease as ν1/2\nu^{-1/2} as we increase the number ν\nu of gates in the random circuit [23].

In this work we consider shallow-depth, variational quantum circuits that are typical in practical applications of near-term quantum computers and answer the question: can variational quantum circuits scramble local gate noise into global depolarising noise? While the answer to this question is relevant for the fundamental understanding of noise processes in near-term quantum devices, it has significant implications in practice: the degree to which local noise is scrambled into white noise determines the performance of a broad class of error mitigation techniques that are of key importance to achieving value with near-term devices [20]. As such, we derive two simple metrics that bound performance guarantees: first, the uniformity measure WW characterises the performance of error mitigation techniques that assume global depolarising (white) noise [25]; second, the norm CC of the commutator between the ideal and noisy quantum states determines the performance of purification-based error mitigation techniques [21, 22] via bounds of ref. [28].

We perform a comprehensive set of numerical experiments to simulate typical applications of near-term quantum computers and analyse characteristics of noise based on the aforementioned two metrics. In all experiments in which we randomly initialise parameters of the variational circuits we semiquantitatively find the same conclusions. First, both metrics, the eigenvalue uniformity WW and the commutator norm CC are well described by our polynomial approximation from Eq. 15 for small gate error rates. Second, this confirms that, similarly to genuine random circuits, local errors get scrambled into global white noise with a polynomially decreasing approximation error as we increase the number of gates. Third, the commutator CC decreases at a higher polynomial rate and has a significantly, by 1-2 orders of magnitude, smaller absolute value in the practically relevant region than the eigenvalue uniformity WW. This confirms that purification based techniques are expected to have a superior performance compared to error mitigation techniques that, e.g., assume a global depolarising noise.

We then investigate the practically more relevant case when the ansatz circuits are initialised near the ground state of a problem Hamiltonian; in all cases we semiquantitatively find the same conclusions. First, the errors do not get scrambled into white noise and the approximation errors are large thus effectively prohibiting or at least significantly limiting the use of error mitigation techniques that assume global depolarising noise. Second, the commutator norm is quite small in absolute value, i.e., 102104\approx 10^{-2}-10^{-4} in the practically relevant region; Since the ansatz circuit prepares the ground state, the square of the commutator norm determines the performance of ESD/VD thus for all applications we simulated we expect a very good performance of the ESD/VD approach. Third, we identify strategies to improve scrambling of local noise into global white noise as we increase circuit depth: We find that inserting additional gates to a HVA that is otherwise not contained in the problem Hamiltonian increases the dimensionality of the dynamic Lie algebra and thus leads to a reduction of both metrics. We find that applying randomised compiling to these non-random, practical circuits also reduces both metrics.

While purification-based techniques [21, 22] have been shown to perform well on specific examples, the present systematic analysis of circuit noise puts these results into perspective and demonstrates the following: First, the superior performance of the ESD/VD technique is not necessarily due to randomness in the quantum circuits – albeit, in deep and random circuits its performance is further improved. Second, while some error mitigation techniques perform well on quantum circuits well-described by white noise [25, 26, 27], we identify various practical scenarios where a limited performance is expected.

The present work advances our understanding of the nature of noise in near-term quantum computers and helps making progress towards achieving value with noisy quantum machines in practical applications. As such, results of the present work will be instrumental for identifying design principles that lead to robust, error-tolerant quantum circuits in practical applications.

Data availability
Numerical simulation code is openly available in the repository: github.com/jfold/shallow-circuit-noise.

Acknowledgements

The authors thank Simon Benjamin for his support. The authors thank Richard Meister for instructions on pyQuEST and Arthur Rattew for multiple discussions on quantum simulations. All simulations in this work were performed using the simulation tools QuEST [45] and its Python interface pyQuEST [46]. B.K. thanks the University of Oxford for a Glasstone Research Fellowship and Lady Margaret Hall, Oxford for a Research Fellowship. B.K. derived analytical results and contributed to writing the manuscript. J.F. was supported by the William Demant Foundation [grant number 18-4438]. J.F. performed numerical simulations and contributed to writing the manuscript.

{quantikz}
\qw

& \gateR_z \gateR_y \gateR_z \ctrl1 \qw \targ \qw
\qw \gateR_z \gateR_y \gateR_z \targ \ctrl1 \qw \qw
\qw \gateR_z \gateR_y \gateR_z \qw \targ \ctrl-2 \qw

Figure 5: A single layer of the Strong Entangling Layers ansatz for three qubits: it first applies single-qubit gates RyRy, RzRz and RyRy on all qubits which is then followed by nearest neighbour CNOT gates.

Appendix A Derivation of Eq. 6

Recall that any quantum state can be transformed into a non-negative arrowhead matrix following Statement 1 from [28] as ρ~=F|ψ~idψ~id|+D+C\tilde{\rho}=F|\tilde{\psi}_{id}\rangle\langle\tilde{\psi}_{id}|+D+C with

ρ~=(FC2C3CdC2D2C3D3CdDd).\tilde{\rho}=\begin{pmatrix}F&C_{2}&C_{3}&\dots&C_{d}\\ C_{2}&D_{2}&&&\\ C_{3}&&D_{3}\\ \vdots&&&\ddots&\vdots\\ C_{d}&&&\dots&D_{d}\end{pmatrix}. (18)

We obtain the above matrix by applying a suitable unitary transformation ρ~:=UρU\tilde{\rho}:=U\rho U^{\dagger} such that |ψ~id:=U|ψid=(1,0,0)|\tilde{\psi}_{id}\rangle:=U|\psi_{id}\rangle=(1,0,\dots 0) while F,Ck,Dk0F,C_{k},D_{k}\geq 0 with k{2,3,,d}k\in\{2,3,\dots,d\} with dd denoting the dimension, and all other matrix entries are zero. Given the above arrowhead representation of a quantum state, one can analytically compute eigenvalues of the density matrix as roots of the following secular equation [28, 47]

P(x)=xF+k=2dCk2(Dkx)=0.P(x)=x-F+\sum_{k=2}^{d}\frac{C_{k}^{2}}{(D_{k}-x)}=0. (19)

With this we compute the deviation between dominant eigenvalue λ1\lambda_{1} and the fidelity as

λ1F=\displaystyle\lambda_{1}-F= k=2dCk2(λ1Dk)maxk(λ1Dk)1k=2dCk2\displaystyle\sum_{k=2}^{d}\frac{C_{k}^{2}}{(\lambda_{1}-D_{k})}\leq\max_{k}(\lambda_{1}-D_{k})^{-1}\sum_{k=2}^{d}C_{k}^{2}
\displaystyle\leq [ρid,ρ]2(2λ11)1,\displaystyle\lVert[\rho_{id},\rho]\rVert^{2}(2\lambda_{1}-1)^{-1}, (20)

where we have used that Dkλ1D_{k}\leq\lambda_{1} and that all summands are non-negative as Dk,Ck,λ10D_{k},C_{k},\lambda_{1}\geq 0, and in the second inequality we have used the series of matrix norms k=2Ck2=CHS2/2=[ρid,ρ]2\sum_{k=2}C_{k}^{2}=\lVert C\rVert_{HS}^{2}/2=\lVert[\rho_{id},\rho]\rVert^{2}_{\infty} as established in [28]. We have also introduced the abbreviation [ρid,ρ]\lVert[\rho_{id},\rho]\rVert given all pp-norms of the matrix [ρid,ρ][\rho_{id},\rho] are equivalent up to a constant factor. In particular, any pp-norm of the commutator can be computed as [ρid,ρ]p=21/pVar[ρ]\lVert[\rho_{id},\rho]\rVert_{p}=2^{1/p}\sqrt{\mathrm{Var}[\rho]} where we used the quantum mechanical variance Var[ρ]:=ψid|ρ2|ψidF2\mathrm{Var}[\rho]:=\expectationvalue{\rho^{2}}{\psi_{id}}-F^{2} as established in [28]. Furthermore, in the second inequality in Eq. 19 we have used that maxk(λ1Dk)1=(λ1D2)1(λ1λ2)1(2λ11)1\max_{k}(\lambda_{1}-D_{k})^{-1}=(\lambda_{1}-D_{2})^{-1}\leq(\lambda_{1}-\lambda_{2})^{-1}\leq(2\lambda_{1}-1)^{-1} by substituting the general inequality λ2(1λ1)\lambda_{2}\leq(1-\lambda_{1}) due to the fact that tr[ρ]=1\tr[\rho]=1.

By denoting the commutator norm as C\mathcal{E}_{C}, we can thus finally conclude that λ1FO(C)\lambda_{1}-F\in O(\mathcal{E}_{C}) as stated in Eq. 6.

A.1 Trace distance from white noise states

In this section we evaluate analytically the trace distance of any quantum state ρ\rho from the corresponding white noise state in Eq. 2 in terms of a distance between probability distributions.

Statement 1.

We can approximate the white noise-state in Eq. 2 in terms of the dominant eigengvalue λ1\lambda_{1} and the dominant eigenvector |ψ1\ket{\psi_{1}} of the quantum state as

ρwn=λ|ψ1ψ1|+(1λ1)Id/d+w,\rho_{wn}=\lambda\outerproduct{\psi_{1}}{\psi_{1}}+(1-\lambda_{1})\mathrm{Id}/d+\mathcal{E}_{w}, (21)

up to an approximation error w\mathcal{E}_{w} that is bounded via Eq. 23.

Proof.

We start by approximating the weight η\eta in Eq. 2 as ηFλ1\eta\approx F\approx\lambda_{1} via Eq. 9 as well as we approximate the dominant eigenvalue using Eq. 6 and then collect the approximation errors as

ρwn=λ|ψidψid|+(1λ1)Id/d+F+C+O(ϵ2/ν).\rho_{wn}=\lambda\outerproduct{\psi_{id}}{\psi_{id}}+(1-\lambda_{1})\mathrm{Id}/d+\mathcal{E}_{F}+\mathcal{E}_{C}+O(\epsilon^{2}/\nu).

We now use results in [28] for bounding the distance between the ideal and noisy quantum states as

|ψidψid||ψ1ψ1|1\displaystyle\lVert\outerproduct{\psi_{id}}{\psi_{id}}-\outerproduct{\psi_{1}}{\psi_{1}}\rVert_{1} =1ψid|ψ1\displaystyle=\sqrt{1-\langle\psi_{id}|\psi_{1}\rangle}
=1O(Cλ1λ2),\displaystyle=1-O\left(\frac{\mathcal{E}_{C}}{\lambda_{1}-\lambda_{2}}\right),

where C\mathcal{E}_{C} is the commutator norm from Eq. 6. We thus establish the approximation

ρwn=λ|ψ1ψ1|+(1λ1)Id/d+w,\rho_{wn}=\lambda\outerproduct{\psi_{1}}{\psi_{1}}+(1-\lambda_{1})\mathrm{Id}/d+\mathcal{E}_{w}, (22)

where we collect all approximation errors as

|w||F|+O(ϵ2/ν)+O[C(1+11λ2/λ1)].|\mathcal{E}_{w}|\leq|\mathcal{E}_{F}|+O(\epsilon^{2}/\nu)+O\left[\mathcal{E}_{C}(1+\frac{1}{1-\lambda_{2}/\lambda_{1}})\right]. (23)

Statement 2.

We define the eigenvalue uniformity as W:=12perrpunif1W:=\tfrac{1}{2}\lVert p_{err}-p_{unif}\rVert_{1} via the non-dominant eigenvalues of the density matrix perr:=(λ2,λ3,,λd)/(1λ1)p_{err}:=(\lambda_{2},\lambda_{3},\dots,\lambda_{d})/(1-\lambda_{1}). This metric is related to the trace distance from a white noise state (as in Eq. 4) as

ρρwn1=(1λ1)W+w,\lVert\rho-\rho_{wn}\rVert_{1}=(1-\lambda_{1})W+\mathcal{E}_{w}, (24)

where the approximation error w\mathcal{E}_{w} is stated in 1.

Proof.

We substitute the approximation of ρwn\rho_{wn} from Eq. 21 including the error term w\mathcal{E}_{w} and then we use the spectral decomposition of ρ\rho to obtain the trace distance as

ρρwn1=\displaystyle\lVert\rho-\rho_{wn}\rVert_{1}= k=2dλk|ψkψk|(1λ1)Id/d1+w\displaystyle\lVert\sum_{k=2}^{d}\lambda_{k}\outerproduct{\psi_{k}}{\psi_{k}}-(1-\lambda_{1})\mathrm{Id}/d\rVert_{1}+\mathcal{E}_{w}
=\displaystyle= 12k=2d|λk1λ1d|+w\displaystyle\frac{1}{2}\sum_{k=2}^{d}|\lambda_{k}-\frac{1-\lambda_{1}}{d}|+\mathcal{E}_{w} (25)
=\displaystyle= 1λ12perrpunif1+w.\displaystyle\frac{1-\lambda_{1}}{2}\lVert p_{err}-p_{unif}\rVert_{1}+\mathcal{E}_{w}. (26)

In the second equation we analytically evaluated the trace distance and thus in the third equation we rewrite the result in terms of perrp_{err} which is our “error probability” distribution as perr:=(λ2,λ3,,λd)/(1λ1)p_{err}:=(\lambda_{2},\lambda_{3},\dots,\lambda_{d})/(1-\lambda_{1}). ∎

Statement 3.

Alternatively to 2, if a quantum state admits the decomposition in Eq. 8 then we can state the trace distance without approximation as

ρρwn1=(1η)2pμpunif1.\lVert\rho-\rho_{wn}\rVert_{1}=\frac{(1-\eta)}{2}\lVert p_{\mu}-p_{unif}\rVert_{1}. (27)

This is directly analogous to the uniformity measure of the non-dominant eigenvalues of ρ\rho in 2, however, this expression quantifies the uniformity of the probability distribution pμp_{\mu} which are eigenvalues of the error density matrix ρerr\rho_{err}.

Let us assume the decomposition in Eq. 8. We find the following result via a direct calculation as

ρρwn1\displaystyle\lVert\rho-\rho_{wn}\rVert_{1} =(1η)ρerrId/d1\displaystyle=(1-\eta)\lVert\rho_{err}-\mathrm{Id}/d\lVert_{1}
=(1η)k=1dμk|ϕkϕk|Id/d1\displaystyle=(1-\eta)\lVert\sum_{k=1}^{d}\mu_{k}\outerproduct{\phi_{k}}{\phi_{k}}-\mathrm{Id}/d\lVert_{1}
=(1η)2k=1d|μk1/d|\displaystyle=\frac{(1-\eta)}{2}\lVert\sum_{k=1}^{d}|\mu_{k}-1/d|
=(1η)2pμpunif1\displaystyle=\frac{(1-\eta)}{2}\lVert p_{\mu}-p_{unif}\rVert_{1}

where we have used the spectral resolution of the error density matrix and then analytically evaluated the trace distance. Given ρerr\rho_{err} is a positive-semidefinite matrix with unit trace, its eigenvalues μk\mu_{k} form a probability distribution that we denote as pμp_{\mu}.

A.2 Upper bounding the uniformity measure

In this section we upper bound the uniformity measure based on the number of gates and error rates in a quantum circuit.

Statement 4.

We adopt the bounds of [23] in Eq. 11 for the distance between probability distributions measured in the standard basis 12p~noisyp~wn1\tfrac{1}{2}\lVert\tilde{p}_{noisy}-\tilde{p}_{wn}\rVert_{1} and assume the same bounds approximately apply to any measurement basis. Then, it follows that the uniformity measure from 2 is approximately bounded by the same bounds as

W=O(eξξ/ν1eξ)+O(w1λ1),W=O(\frac{e^{-\xi}\xi/\sqrt{\nu}}{1-e^{-\xi}})+O(\frac{\mathcal{E}_{w}}{1-\lambda_{1}}),

where the approximation error w\mathcal{E}_{w} is stated in 1.

Proof.

Let us consider measurements performed in the basis as the eigenvectors of the density matrix which yield probabilities as the eigenvalues as

pnoisy=ψk|ρ|ψk=(λ1,λ2,λd).p_{noisy}=\bra{\psi_{k}}\rho\ket{\psi_{k}}=(\lambda_{1},\lambda_{2}\dots,\lambda_{d}).

Measuring the white noise state in the same basis yields the following approximation of the probabilities using the error term from Eq. 21 as

pwn\displaystyle p_{wn} :=ψk|ρwn|ψk\displaystyle:=\bra{\psi_{k}}\rho_{wn}\ket{\psi_{k}}
=(λ1,1λ1d,1λ1d)+w.\displaystyle=(\lambda_{1},\frac{1-\lambda_{1}}{d}\dots,\frac{1-\lambda_{1}}{d})+\mathcal{E}_{w}.
Refer to caption
Figure 6: (left) TFI-HVA ansatz: same simulations as in Fig. 3 (a) but with added parametrised RzR_{z} gates after each layer. The additional gates increase the dimensionality of the dynamic Lie algebra which leads to a faster scrambling of local gate noise into white noise, e.g., the ϵ0\epsilon\rightarrow 0 curve is steeper than in Fig. 3 (a). See Appendix B for more details. (right) the dependence on the number of qubits shows a very similar trend as without the RzRz gates, i.e., compare to Fig. 7 (c).

The distance of the above two measurement probability distributions is then

12pnoisypwn1=(1λ1)W+w,\frac{1}{2}\lVert p_{noisy}-p_{wn}\rVert_{1}=(1-\lambda_{1})W+\mathcal{E}_{w},

where W=12perrpunif1W=\tfrac{1}{2}\lVert p_{err}-p_{unif}\rVert_{1} is our eigenvalue uniformity from 2. Under the assumption that the upper bound on the measurement probabilities 12p~noisyp~wn1\tfrac{1}{2}\lVert\tilde{p}_{noisy}-\tilde{p}_{wn}\rVert_{1} from Eq. 11 approximately holds for any measurement basis we can bound the eigenvalue uniformity as

W\displaystyle W =12(1λ1)pnoisypwn1+w1λ1\displaystyle=\frac{1}{2(1-\lambda_{1})}\lVert p_{noisy}-p_{wn}\rVert_{1}+\frac{\mathcal{E}_{w}}{1-\lambda_{1}}
O(F1λ1ϵν)+w1λ1\displaystyle\leq O\left(\frac{F}{1-\lambda_{1}}\epsilon\sqrt{\nu}\right)+\frac{\mathcal{E}_{w}}{1-\lambda_{1}}
=O(eξξ/ν1eξ)+O(w1λ1).\displaystyle=O\left(\frac{e^{-\xi}\xi/\sqrt{\nu}}{1-e^{-\xi}}\right)+O\left(\frac{\mathcal{E}_{w}}{1-\lambda_{1}}\right).

In the last equation we introduced the approximation of FF from Eq. 9 as well as the approximate dominant eigenvalue from Eq. 6. ∎

A.2.1 Expanding the upper bound

We now expand the upper bound from 4 for small ξ\xi as. More specifically, we consider the parametrised fit function from Eq. 15 and substitute the Taylor expansion eξ=1ξ+ξ2+e^{-\xi}=1-\xi+\xi^{2}+\dots as

αeξξ/νβ1eξ\displaystyle\alpha\frac{e^{-\xi}\xi/\sqrt{\nu}^{\beta}}{1-e^{-\xi}} =αeξνβξξξ2/2+\displaystyle=\alpha\frac{e^{-\xi}}{\nu^{\beta}}\frac{\xi}{\xi-\xi^{2}/2+\dots}
=αeξνβ11ξ/2+\displaystyle=\alpha\frac{e^{-\xi}}{\nu^{\beta}}\frac{1}{1-\xi/2+\dots}
=α1νβ1ξ+1ξ/2+\displaystyle=\alpha\frac{1}{\nu^{\beta}}\frac{1-\xi+\dots}{1-\xi/2+\dots}
=ανβ+O(ξ).\displaystyle=\frac{\alpha}{\nu^{\beta}}+O(\xi).
Refer to caption
Figure 7: Fit parameters α\alpha from Eq. 15 for an increasing number of qubits: The circuits in Figs. 1, 2, 3 and 4 at ϵ0\epsilon\rightarrow 0 were simulated for an increasing number of qubits and the curve from Eq. 15 was fitted.

A.3 Commutator norm

Lemma 1.

The commutators norms are approximately related as

[ρid,ρ]11λ1=[ρid,ρerr]1+q,\frac{\lVert[\rho_{id},\rho]\rVert_{1}}{1-\lambda_{1}}=\lVert[\rho_{id},\rho_{err}]\rVert_{1}+\mathcal{E}_{q}, (28)

up to the approximation error q\mathcal{E}_{q}.

Proof.

Using the decomposition from Eq. 8 we obtain

[ρid,ρ]1\displaystyle\lVert[\rho_{id},\rho]\rVert_{1} =[ρid,ηρid]+[ρid,(1η)ρerr]1\displaystyle=\lVert[\rho_{id},\eta\rho_{id}]+[\rho_{id},(1-\eta)\rho_{err}]\rVert_{1}
=(1η)[ρid,ρerr]1\displaystyle=(1-\eta)\lVert[\rho_{id},\rho_{err}]\rVert_{1}

We can approximate η=λ1+𝒪(F)+𝒪(C)\eta=\lambda_{1}+\mathcal{O}(\mathcal{E}_{F})+\mathcal{O}(\mathcal{E}_{C}) via Eq. 9 and Eq. 6 and obtain that

[ρid,ρ]11λ1=[ρid,ρerr]1+q.\frac{\lVert[\rho_{id},\rho]\rVert_{1}}{1-\lambda_{1}}=\lVert[\rho_{id},\rho_{err}]\rVert_{1}+\mathcal{E}_{q}. (29)

The error term can be obtained via the triangle inequality |q|[𝒪(FC)+𝒪(C2)]/(1λ1)|\mathcal{E}_{q}|\leq[\mathcal{O}(\mathcal{E}_{F}\mathcal{E}_{C})+\mathcal{O}(\mathcal{E}_{C}^{2})]/(1-\lambda_{1}). ∎

Appendix B Further details of numerical simulations

B.1 The SEL and HVA ansätze

The circuit structure of the SEL ansatz used in Fig. 1 is illustrated in Fig. 5: it consists of alternating layers of parametrised single-qubit rotations and a ladder of nearest-neighbour CNOT gates.

Let us now define the HVA ansatz. In particular, recall that the HVA ansatz is a discretisation of the adiabatic evolution

U(β¯,γ¯)=k=1νeiγk1eiβk0,U(\underline{\beta},\underline{\gamma})=\prod_{k=1}^{\nu}e^{-i\gamma_{k}\mathcal{H}_{1}}e^{-i\beta_{k}\mathcal{H}_{0}},

which is applied to the initial state as the ground state of the trivial Hamiltonian 0\mathcal{H}_{0}.

The individual evolutions are then trotterised such that a piece of time evolution eiγk1e^{-i\gamma_{k}\mathcal{H}_{1}} is broken up into products of evolution operators under the individual Hamiltonian terms as

eiγk1l=1rheiγkhlPl.e^{-i\gamma_{k}\mathcal{H}_{1}}\rightarrow\prod_{l=1}^{r_{h}}e^{-i\gamma_{k}h_{l}P_{l}}.

Above we utilised the decomposition of the non-trivial part of the Hamiltonian 1=l=1rhhlPl\mathcal{H}_{1}=\sum_{l=1}^{r_{h}}h_{l}P_{l} into Pauli strings Pl{Id,X,Y,Z}NP_{l}\in\{\mathrm{Id},X,Y,Z\}^{\otimes N}.

We set the circuit parameter as γk=k/ν\gamma_{k}=k/\nu and βk=1k/ν\beta_{k}=1-k/\nu, such that the circuit approximates a discretised adiabatic evolution between 0\mathcal{H}_{0} and 1\mathcal{H}_{1} – and we will refer to these as VQE parameters.

In the case of random parametrisation of the HVA ansatz, every gate implementing the evolution under a single Pauli string eiγkhlPle^{-i\gamma_{k}h_{l}P_{l}} is assigned a random parameter as eiθqPle^{-i\theta_{q}P_{l}} with |θq|2π|\theta_{q}|\leq 2\pi.

B.2 Inserting additional gates to the TFI ansatz

In Fig. 6 we repeated the same simulation as in Fig. 3 (a), i.e., using a HVA ansatz for the TFI spin model at random circuit parameters, but we appended to each layer a series of parametrised RzRz gates on each qubit. This guarantees that the dynamic Lie algebra generated by the Pauli terms of the TFI problem in Eq. 16 is expanded by the inclusion of Pauli Z operators. Increasing the circuit depth of the HVA ansatz thus leads to a faster increase of the dimensionality of the Lie algebra which demonstrably leads to a faster scrambling of local noise into global white noise, e.g., steeper slope of the ϵ0\epsilon\rightarrow 0 fit in Fig. 6 than in Fig. 3.

B.3 Scaling with the number of qubits

In Fig. 7 we simulate the same circuits as in Figs. 1, 2, 3 and 4 at error rates ϵ0\epsilon\rightarrow 0 and plot the fit parameter α\alpha—which is the prefactor in Eq. 15—for an increasing number of qubits. The results appear to confirm an asymptotically non-increasing trend confirming theoretical expectations of [23] for random circuits whereby α\alpha is constant bounded in terms of the number of qubits.

References

  • Arute et al. [2019] F. Arute, K. Arya, R. Babbush, D. Bacon, J. C. Bardin, R. Barends, R. Biswas, S. Boixo, F. G. Brandao, D. A. Buell, et al., Quantum supremacy using a programmable superconducting processor, Nature 574, 505 (2019).
  • Tillmann et al. [2013] M. Tillmann, B. Dakić, R. Heilmann, S. Nolte, A. Szameit, and P. Walther, Experimental boson sampling, Nature photonics 7, 540 (2013).
  • Biamonte et al. [2017] J. Biamonte, P. Wittek, N. Pancotti, P. Rebentrost, N. Wiebe, and S. Lloyd, Quantum machine learning, Nature 549, 195 (2017).
  • Jafferis et al. [2022] D. Jafferis, A. Zlokapa, J. D. Lykken, D. K. Kolchmeyer, S. I. Davis, N. Lauk, H. Neven, and M. Spiropulu, Traversable wormhole dynamics on a quantum processor, Nature 612, 51 (2022).
  • Kokail et al. [2019] C. Kokail, C. Maier, R. van Bijnen, T. Brydges, M. K. Joshi, P. Jurcevic, C. A. Muschik, P. Silvi, R. Blatt, C. F. Roos, and P. Zoller, Self-verifying variational quantum simulation of lattice models, Nature 569, 355 (2019).
  • Cao et al. [2019] Y. Cao, J. Romero, J. P. Olson, M. Degroote, P. D. Johnson, M. Kieferová, I. D. Kivlichan, T. Menke, B. Peropadre, N. P. Sawaya, et al., Quantum chemistry in the age of quantum computing, Chemical reviews 119, 10856 (2019).
  • McArdle et al. [2020] S. McArdle, S. Endo, A. Aspuru-Guzik, S. C. Benjamin, and X. Yuan, Quantum computational chemistry, Reviews of Modern Physics 92, 015003 (2020).
  • Bauer et al. [2020] B. Bauer, S. Bravyi, M. Motta, and G. K.-L. Chan, Quantum algorithms for quantum chemistry and quantum materials science, Chemical Reviews 120, 12685 (2020).
  • Motta and Rice [2022] M. Motta and J. E. Rice, Emerging quantum computing algorithms for quantum chemistry, WIREs Computational Molecular Science 12, e1580 (2022).
  • Preskill [2018] J. Preskill, Quantum computing in the nisq era and beyond, Quantum 2, 79 (2018).
  • Cerezo et al. [2021a] M. Cerezo, A. Arrasmith, R. Babbush, S. C. Benjamin, S. Endo, K. Fujii, J. R. McClean, K. Mitarai, X. Yuan, L. Cincio, et al., Variational quantum algorithms, Nature Reviews Physics 3, 625 (2021a).
  • Endo et al. [2021] S. Endo, Z. Cai, S. C. Benjamin, and X. Yuan, Hybrid quantum-classical algorithms and quantum error mitigation, Journal of the Physical Society of Japan 90, 032001 (2021).
  • Bharti et al. [2022] K. Bharti, A. Cervera-Lierta, T. H. Kyaw, T. Haug, S. Alperin-Lea, A. Anand, M. Degroote, H. Heimonen, J. S. Kottmann, T. Menke, W.-K. Mok, S. Sim, L.-C. Kwek, and A. Aspuru-Guzik, Noisy intermediate-scale quantum algorithms, Rev. Mod. Phys. 94, 015004 (2022).
  • Cerezo et al. [2022] M. Cerezo, G. Verdon, H.-Y. Huang, L. Cincio, and P. J. Coles, Challenges and opportunities in quantum machine learning, Nature Computational Science 2, 567 (2022).
  • van Straaten and Koczor [2021] B. van Straaten and B. Koczor, Measurement cost of metric-aware variational quantum algorithms, PRX Quantum 2, 030324 (2021).
  • Koczor and Benjamin [2022a] B. Koczor and S. C. Benjamin, Quantum analytic descent, Phys. Rev. Res. 4, 023017 (2022a).
  • Huang et al. [2020] H.-Y. Huang, R. Kueng, and J. Preskill, Predicting many properties of a quantum system from very few measurements, Nature Physics 16, 1050 (2020).
  • Boyd and Koczor [2022] G. Boyd and B. Koczor, Training variational quantum circuits with covar: Covariance root finding with classical shadows, Phys. Rev. X 12, 041022 (2022).
  • Chan et al. [2022] H. H. S. Chan, R. Meister, M. L. Goh, and B. Koczor, Algorithmic shadow spectroscopy, arXiv preprint arXiv:2212.11036  (2022).
  • Cai et al. [2022] Z. Cai, R. Babbush, S. C. Benjamin, S. Endo, W. J. Huggins, Y. Li, J. R. McClean, and T. E. O’Brien, Quantum error mitigation, arXiv preprint arXiv:2210.00921  (2022).
  • Koczor [2021a] B. Koczor, Exponential Error Suppression for Near-Term Quantum Devices, Phys. Rev. X 11, 031057 (2021a).
  • Huggins et al. [2021] W. J. Huggins, S. McArdle, T. E. O’Brien, J. Lee, N. C. Rubin, S. Boixo, K. B. Whaley, R. Babbush, and J. R. McClean, Virtual distillation for quantum error mitigation, Phys. Rev. X 11, 041036 (2021).
  • Dalzell et al. [2021] A. M. Dalzell, N. Hunter-Jones, and F. G. Brandão, Random quantum circuits transform local noise into global white noise, arXiv preprint arXiv:2111.14907  (2021).
  • Larocca et al. [2022] M. Larocca, P. Czarnik, K. Sharma, G. Muraleedharan, P. J. Coles, and M. Cerezo, Diagnosing barren plateaus with tools from quantum optimal control, Quantum 6, 824 (2022).
  • Vovrosh et al. [2021] J. Vovrosh, K. E. Khosla, S. Greenaway, C. Self, M. S. Kim, and J. Knolle, Simple mitigation of global depolarizing errors in quantum simulations, Phys. Rev. E 104, 035309 (2021).
  • Endo et al. [2018] S. Endo, S. C. Benjamin, and Y. Li, Practical quantum error mitigation for near-future applications, Phys. Rev. X 8, 031027 (2018).
  • Strikis et al. [2021] A. Strikis, D. Qin, Y. Chen, S. C. Benjamin, and Y. Li, Learning-based quantum error mitigation, PRX Quantum 2, 040330 (2021).
  • Koczor [2021b] B. Koczor, The dominant eigenvector of a noisy quantum state, New Journal of Physics 23, 123047 (2021b).
  • Koczor and Benjamin [2022b] B. Koczor and S. C. Benjamin, Quantum natural gradient generalized to noisy and nonunitary circuits, Phys. Rev. A 106, 062416 (2022b).
  • O’Brien et al. [2022] T. E. O’Brien, G. Anselmetti, F. Gkritsis, V. Elfving, S. Polla, W. J. Huggins, O. Oumarou, K. Kechedzhi, D. Abanin, R. Acharya, et al., Purification-based quantum error mitigation of pair-correlated electron simulations, arXiv preprint arXiv:2210.10799  (2022).
  • Jnane et al. [2022] H. Jnane, B. Undseth, Z. Cai, S. C. Benjamin, and B. Koczor, Multicore Quantum Computing, Phys. Rev. Appl. 18, 044064 (2022).
  • Schuld et al. [2020] M. Schuld, A. Bocharov, K. M. Svore, and N. Wiebe, Circuit-centric quantum classifiers, Physical Review A 101, 032308 (2020).
  • Koczor et al. [2020] B. Koczor, S. Endo, T. Jones, Y. Matsuzaki, and S. C. Benjamin, Variational-state quantum metrology, New J. Phys. 22, 083038 (2020).
  • Foldager et al. [2022] J. Foldager, A. Pesah, and L. K. Hansen, Noise-assisted variational quantum thermalization, Scientific reports 12, 1 (2022).
  • Silva et al. [2008] M. Silva, E. Magesan, D. W. Kribs, and J. Emerson, Scalable protocol for identification of correctable codes, Phys. Rev. A 78, 012347 (2008).
  • Magesan et al. [2012] E. Magesan, J. M. Gambetta, and J. Emerson, Characterizing quantum gates via randomized benchmarking, Phys. Rev. A 85, 042311 (2012).
  • Cai and Benjamin [2019] Z. Cai and S. C. Benjamin, Constructing smaller pauli twirling sets for arbitrary error channels, Sci. Rep. 9, 1 (2019).
  • Cai et al. [2020] Z. Cai, X. Xu, and S. C. Benjamin, Mitigating coherent noise using Pauli conjugation, npj Quantum Info. 6, 1 (2020).
  • McClean et al. [2018] J. R. McClean, S. Boixo, V. N. Smelyanskiy, R. Babbush, and H. Neven, Barren plateaus in quantum neural network training landscapes, Nature communications 9, 1 (2018).
  • Cerezo et al. [2021b] M. Cerezo, A. Sone, T. Volkoff, L. Cincio, and P. J. Coles, Cost function dependent barren plateaus in shallow parametrized quantum circuits, Nature communications 12, 1 (2021b).
  • Wecker et al. [2015] D. Wecker, M. B. Hastings, and M. Troyer, Progress towards practical quantum variational algorithms, Physical Review A 92, 042303 (2015).
  • Wiersema et al. [2020] R. Wiersema, C. Zhou, Y. de Sereville, J. F. Carrasquilla, Y. B. Kim, and H. Yuen, Exploring entanglement and optimization within the hamiltonian variational ansatz, PRX Quantum 1, 020319 (2020).
  • Campbell [2019] E. Campbell, Random compiler for fast hamiltonian simulation, Phys. Rev. Lett. 12310.1103/physrevlett.123.070503 (2019).
  • Ouyang et al. [2020] Y. Ouyang, D. R. White, and E. T. Campbell, Compilation by stochastic Hamiltonian sparsification, Quantum 4, 235 (2020).
  • Jones et al. [2019] T. Jones, A. Brown, I. Bush, and S. C. Benjamin, QuEST and high performance simulation of quantum computers, Scientific reports 9, 1 (2019).
  • Meister [2022] R. Meister, pyquest - a python interface for the quantum exact simulation toolkit (2022).
  • O’leary and Stewart [1990] D. O’leary and G. Stewart, Computing the eigenvalues and eigenvectors of symmetric arrowhead matrices, Journal of Computational Physics 90, 497 (1990).